Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

This article was downloaded by: [Georgetown University]

On: 02 February 2015, At: 13:38


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered
office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Philosophical Magazine
Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/tphm20

Effect of solute atoms on fracture


toughness in dilute magnesium alloys
a a a
Hidetoshi Somekawa , Tadanobu Inoue & Kaneaki Tsuzaki
a
Research Center for Strategic Materials, National Institute for
Materials Science, 1-2-1 Sengen, Tsukuba, Ibaraki, 305-0047,
Japan.
Published online: 16 Sep 2013.

To cite this article: Hidetoshi Somekawa, Tadanobu Inoue & Kaneaki Tsuzaki (2013) Effect of solute
atoms on fracture toughness in dilute magnesium alloys, Philosophical Magazine, 93:36, 4582-4592,
DOI: 10.1080/14786435.2013.838008

To link to this article: http://dx.doi.org/10.1080/14786435.2013.838008

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the
“Content”) contained in the publications on our platform. However, Taylor & Francis,
our agents, and our licensors make no representations or warranties whatsoever as to
the accuracy, completeness, or suitability for any purpose of the Content. Any opinions
and views expressed in this publication are the opinions and views of the authors,
and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content
should not be relied upon and should be independently verified with primary sources
of information. Taylor and Francis shall not be liable for any losses, actions, claims,
proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or
howsoever caused arising directly or indirectly in connection with, in relation to or arising
out of the use of the Content.

This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden. Terms &
Conditions of access and use can be found at http://www.tandfonline.com/page/terms-
and-conditions
Philosophical Magazine, 2013
Vol. 93, No. 36, 4582–4592, http://dx.doi.org/10.1080/14786435.2013.838008

Effect of solute atoms on fracture toughness in dilute magnesium


alloys
Hidetoshi Somekawa*, Tadanobu Inoue and Kaneaki Tsuzaki

Research Center for Strategic Materials, National Institute for Materials Science, 1-2-1 Sengen,
Tsukuba, Ibaraki 305-0047, Japan
(Received 15 April 2013; accepted 21 August 2013)
Downloaded by [Georgetown University] at 13:38 02 February 2015

The effect of alloying elements on the toughness and the fracture behaviour
was investigated on seven kinds of Mg-0.3 at.% X (X = Ag, Al, Ca, Pb, Sn, Y
and Zn) alloys with a grain size of 3–5 μm. The fracture toughness and
fracture behaviour in magnesium alloys were closely related to the segregation
energy. The Mg–Al and –Zn alloys that had small segregation energy showed
high toughness and ductile fracture in most regions, while the Mg–Ca alloy
with large segregation energy exhibited low toughness and intergranular
fracture. These different tendencies resulted from solute segregation at grain
boundaries (GBs). The change in the lattice parameter ratio was the influential
material parameter regardless of whether the GB embrittlement was for
enhancement or suppression.
Keywords: magnesium alloys; segregation; fracture toughness; solute atom

1. Introduction
Magnesium and its alloys, which are the lightest among all the conventional alloys,
have great potential as structural materials in industry. However, their mechanical prop-
erties, i.e. strength and toughness, must be sufficient to satisfy both the reliability and
safety requirements. Two types of methods are used to develop magnesium alloys with
high strength and high toughness properties. The first method is grain refinement. Since
magnesium has a large Taylor factor, its Hall–Petch slope is larger than that in the other
metallic materials [1], which indicates that the strength increases considerably with finer
grain structures [2]. The grain refinement also tends to decrease and prevent twinning
formation [3], although the interface between the matrix and the twin boundary
becomes the crack-propagation path site for magnesium during the fracture toughness
tests [4].
The other method for enhancing these properties is the addition of solid solution
atoms. In addition to the classical results using single crystals [5], the first principal
calculation [6] has also shown that the addition of solute atoms with a large misfit size
is effective for increasing the strength of magnesium alloys. Furthermore, the stress to
form the twinning, i.e. f1 0 1 2g-type twinning, is increased with the addition of such
solute elements [7]. Indeed, while the toughness in pure magnesium with an average

*Corresponding author. Email: somekawa.hidetoshi@nims.go.jp

Ó 2013 Taylor & Francis


Philosophical Magazine 4583

grain size of 5 μm is around 15 MPam1/2 [8], the toughness of the conventional/


commercial magnesium alloys, such as Mg–Al–Zn (AZ31) and Mg–Zn–Zr (ZK60)
alloys, improves to around 25 and 30 MPam1/2, respectively [9]. The combination of
using both the fine-grained structures and solid solution alloying has the potential to
enhance these properties even further in the magnesium alloys. However, neither the
toughness in the binary alloys nor the role of alloying elements on toughness has been
clarified yet. Therefore, the effect of alloying elements on toughness was investigated
using several kinds of magnesium binary alloys with fine-grained structures in this
study.

2. Experimental and numerical procedures


Seven kinds of Mg–0.3 at.% X (X = Ag, Al, Ca, Pb, Sn, Y and Zn) binary alloys were
Downloaded by [Georgetown University] at 13:38 02 February 2015

used in this study. They were produced by casting, and were solution-treated at a
temperature of 773 K for 7.2 h. Then, they were extruded into a plate shape at several
temperatures, which are listed in Table 1, to refine the microstructures and to control the
texture so that the basal plane is parallel to the extrusion direction. The microstructures
of the extruded alloys were observed by transmission electron microscopy (TEM) and
electron back-scattered diffraction (EBSD).
The fracture toughness was evaluated by the J-integral method according to the
ASTM E1890. The specimen that was used for investigating the fracture toughness con-
sisted of a three-point bending sample with a width of 10 mm, a thickness of 5 mm and
a length of 44 mm. It was machined directly from the extruded plate, and a V-notch
was made normal to the extruded direction. The fracture surface was observed by a
scanning electron microscope (SEM) after the toughness test in the ASTM E399. The
deformed microstructures near the crack tip were also observed by the EBSD method in
the Mg–Al and Mg–Ca alloys. The deformed samples were obtained, when the tough-
ness tests were discontinued as the load reached the maximum value.
The surface segregation energy in the specific magnesium alloys was obtained using
the molecular dynamics (MD) simulation. The surface segregation energy is defined as
the difference in the total energy between the solute atoms existing in the bulk solid
solution state, Esol, and in the free surface of the basal or the non-basal planes, Ebasal
and Enon-basal. The surface energy in the basal plane is reported to be lower than that in

Table 1. The results of fracture toughness tests in the magnesium binary alloys.

T, K JIC, MPam r, nm [14] X0, at.% [13,14] T0, K [13,14] Fracture type
Mg–Al 458 0.0081 0.143 11.5 710 D
Mg–Ag 493 0.0055 0.144 3.8 745 D+I (I > D)
Mg–Ca 498 0.0018 0.197 0.8 789 D+I (I >> D)
Mg–Sn 448 0.0046 0.158 3.4 834 D+I (I > D)
Mg–Pb 418 0.0057 0.175 7.8 739 D+I (I > D)
Mg–Y 563 0.0040 0.181 3.5 840 D+I (I > D)
Mg–Zn 458 0.0099 0.137 2.4 613 D
where T is the extrusion temperature, JIC is the fracture toughness, r is the atomic radius [14], X0 is the
maximum solubility [13,14] and T0 is the eutectic temperature [13,14]. D and I are the ductile fracture and the
intergranular fracture, respectively.
4584 H. Somekawa et al.

the non-basal plane in magnesium [10]. Since there is a possibility that the solute atoms
exist in the basal or the non-basal planes, the average surface segregation energy that
ignores the effect of texture is determined using the follow equation:

   
1 Esol  Ebasal Esol  Enon-basal
ES ¼ þ ð1Þ
2 Abasal Anon-basal

where A is the area in the basal or the non-basal planes. The total energies of Esol, Ebasal
and Enon-basal in magnesium alloys are obtained from the MD simulation. The simula-
tion was performed at a temperature of 298 K under an NTP ensemble and 1 atm. The
generalized embedded-atom method (GEAM) potential was used for magnesium and its
alloys [11].
Downloaded by [Georgetown University] at 13:38 02 February 2015

3. Results
3.1. Fracture toughness
Typical initial microstructures of the extruded alloys are shown in Figure 1. The
average grain size, d, of these alloys is 3–5 μm. The misorientation angles are analysed
in the image quality maps, and the grain structures of the extruded alloys are found to
consist of mostly high-angle grain boundaries (GBs). The texture distribution from the
EBSD analysis is included in each figure. The peak intensities of all the alloys tend to
accumulate at the basal plane due to the wrought process. This feature has also been
observed in the extruded magnesium alloys [12]. The precipitate particles are not con-
firmed in any of the TEM images, even in the Mg–Ca alloy despite having a maximum
solubility of 0.82 at.% [13]. This is reasonable since the alloying level of 0.3 at.% is
below the solubility limit of each alloy at each extrusion temperature.

Figure 1. Typical microstructures of the extruded alloys by TEM and EBSD observations: (a)
Mg–Ca, (b) Mg–Y, (c) Mg–Ag, (d) Mg–Pb, (e) Mg–Al, (f) Mg–Zn and (g) Mg–Sn alloys. Black
and yellow lines in Figure (c–g) indicate the high-angle and low-angle GBs with the
misorientation angles of 15° 6 h and 3° 6 h < 15°, respectively.
Philosophical Magazine 4585

The results of the fracture toughness tests using the J-integral method are
summarized in Table 1 for all of the extruded alloys. This table includes the atomic
radii of each element [14]. The alloying elements affect the toughness. The Mg–Zn
alloy and the Mg–Ca alloy have the highest and the lowest toughness, respectively,
among the alloys. The toughness is influenced by the microstructural factors, such as
the average grain size, the dispersion of second phase/precipitate particles and texture
distribution; however, the microstructural observations (e.g. Figure 1) show that all of
the alloys have similar grain sizes and texture feature, and do not contain these
particles. Therefore, these microstructural factors are not considered in this study.
Figure 2 shows the fracture surfaces of (a) Mg–Al and (b) Mg–Y alloys observed
by SEM after the toughness tests. These figures show that the type of fracture is influ-
enced by the alloying elements. The Mg–Al alloy, which has a very high toughness
among the present alloys, shows a ductile fracture resulting from the void formation.
Downloaded by [Georgetown University] at 13:38 02 February 2015

However, the dominant fracture in the Mg–Y alloy is not the ductile fracture. The size
of a step in the fracture region is around 5 μm, which corresponds to the grain size.
Although all the micrographs are not shown here, the SEM observations for the other
alloys show similar tendencies. The ductile fracture can be observed in small regions in
the alloys with a low toughness, but most of the regions show non-ductile fractures.
Table 1 presents the type of fracture that occurs in all of the alloys.
Typical deformed microstructures near the crack tip are shown in Figure 3(a), (c)
Mg–Al and (b), (d) Mg–Ca alloys. Figure 3(a–d) show the inverse pole figures image

Figure 2. SEM observation of the fracture surface after the toughness tests in the ASTM
E399: (a) Mg–Al and (b) Mg–Y alloys.
4586 H. Somekawa et al.
Downloaded by [Georgetown University] at 13:38 02 February 2015

Figure 3. Typical deformed microstructures near the crack tip using EBSD observation: (a), (c)
Mg-Al and (b), (d) Mg-Ca alloys. Black and white arrows indicate the f1 0  1 2g twinning
confirmed by EBSD analysis and the crack-propagation route, respectively.

and the image quality maps obtained by the EBSD method, respectively. The EBSD
observations of the fine-grained alloys, i.e. several μm sizes, could not identify any
deformed microstructures such as the GBs and twin boundaries clearly due to the severe
strains around the crack tip. Therefore, we substituted for the alloy ones with d = 10 μm,
which were prepared by static annealing after the extrusion. Figure 3 shows that a small
amount of f1 0  1 2g twins, which is marked by the arrows, can be observed in both
alloys, since the grain size of around 10 μm is not fine enough to prevent the twinning
formation. However, the amount of twinning is too small to affect the fracture
behaviour of becoming crack-propagation sites. The biggest concern in this observation
is that the alloying elements produce a change in the crack-propagation path, which is
marked by the white arrows in Figure 3(c) and (d). The cracks propagate into the grain
interior in the Mg–Al alloy. The blunt in the crack tip of this alloy is also larger than
that of the Mg–Ca alloy, and thus indicates a higher toughness. On the contrary, the
GBs are the crack-propagation sites in the Mg–Ca alloy, i.e. occurrence of intergranular
fracture, which is the type of fracture in the alloys with a low toughness (in Figure 2
(b)). These results reveal that the crack-propagation behaviour is closely related to the
difference in the toughness of the magnesium alloys.
Philosophical Magazine 4587

3.2. GB/surface segregation energy


Many papers have reported that the intergranular fracture is concerned with the GB
embrittlement due to the impurity/solute atom segregation. For example, the addition of
phosphorus or sulphur atom into the iron system and nickel system alloys [15–17], and
the addition of alkaline atoms into the aluminium system alloys enhance the occurrence
of GB embrittlement [18]. On the contrary, in the iron system alloys, the carbon or
boron atom is reported to play a role in the enhancement for GB cohesion [17]. Rice
and Wang have investigated the segregation effects on GB embrittlement/cohesion from
a thermodynamic viewpoint [19]. They showed that the difference in the binding energy
between the GB segregation energy, ΔEgb, and the surface segregation energy, ΔEs, is
critical in determining whether the impurity/solute atom is an embrittlement or a
cohesion enhancer. A brief explanation of their model is as follows. When the surface
segregation energy is lower than the GB segregation energy, DEgb  DEs [ 0 (note that
Downloaded by [Georgetown University] at 13:38 02 February 2015

both ΔEgb and ΔEs are negative), the impurity/solute atoms at the surface are more
stable, and thus would reduce the GB cohesion, i.e. causing the enhancement of GB
embrittlement. On the other hand, the opposite, DEgb  DEs \ 0, indicates the enhance-
ment of GB cohesion. Hence, hereafter, we consider the impact of solute atoms on GB
embrittlement/cohesion in the magnesium alloys.
To the best of our knowledge, there have not been any reports on the influence of
alloying elements on these segregation energies in the magnesium alloys. Thus, the
well-known free energy balance and the MD simulation are used to obtain a rough esti-
mation of the energies and to understand the segregation behaviours. The following free
energy balance is obtained by considering the saturated GBs and a diluted solution of
solute atoms in the matrix [20]:

cgb ¼ c0gb þ CðDHseg  RT ln X Þ ð2Þ

where γ0gb and γgb are the non-segregated and the segregated GB energies, respectively.
The values of RT lnX and ΔHseg are the segregation entropy and the enthalpy terms,
respectively; Γ is the GB excess (=7  1018/m2 [21]); R is the gas constant; T is the
working (extrusion) temperature (listed in Table 1); and X is the bulk solute concentra-
tion (= 0.3 at.% in this study). The difference between the entropy and the enthalpy is
assumed to be the GB segregation energy; DEgb ¼ DHseg  RT ln X . However, in the
present calculation, it is difficult to obtain the segregation enthalpy in each binary sys-
tem due to the lack of data, and thus, the following empirical relation is used [22];

DHseg  DGseg ¼ RT0 ln X0  ð10  6Þ ð3Þ

where X0 is the terminal solute solubility and T0 is the eutectic temperature. The values
for X0 and T0 are quoted from the binary phase diagrams in the references [13,14] and
presented in Table 1.
The results of the GB segregation energies obtained using Equations (2) and (3) are
shown in Figure 4. Before moving to the discussion, the validity of GB segregation
energy is confirmed by the other calculation method; the well-known McLean model,
which is the function of solute concentration [23]. The detailed calculation is shown in
Appendix B. As compared with the GB segregation energies obtained by the McLean
4588 H. Somekawa et al.
Downloaded by [Georgetown University] at 13:38 02 February 2015

Figure 4. The relationship between the fracture toughness, JIC, and the GB and surface
segregation energy in the magnesium binary alloys.

model and Equations (2) and (3), they show similar values and ranges. Thus, the use of
empirical equations is sufficient for understanding the effect of solute atoms on GB
segregation behaviour, in this study. This figure reveals that the difference in energy
between enthalpy and entropy has been shown to be negative, DEgb ¼ DHseg 
RT ln X \ 0, in all of the present alloys. This result indicates that all the present solute
atoms tend to favour solute segregation at the GBs. Recent experimental studies have
also reported that the segregation of solute atoms at the GBs was observed by the
TEM-EDX analysis in the magnesium binary alloys [24,25].
Figure 4 also indicates the relationship between the toughness (listed in Table 1)
and the GB and surface segregation energies. The surface segregation energy could only
be calculated for the Mg–Al, –Ag and –Pb alloys, since the present GEAM potential is
not appropriate for the other solute atoms in the magnesium alloys. The difference in
segregation energies, DE ¼ DEgb  DEs , are calculated to be 1001, 942 and 608 mJ/m2
for Mg–Ag, –Pb and –Al alloys, respectively. The present calculation based on the three
alloys indicates the two important behaviours from thermodynamic viewpoints; (i) the
surface segregation energy is lower than the GB segregation energy, DEgb  DEs [ 0,
and (ii) the alloy with the larger difference in segregation energies, ΔE, has lower
toughness, JIC: Mg–Al > –Pb > –Ag. Hence, we have determined that the alloy with
large GB segregation energy has large surface segregation energy. Based on this
assumption, we can say that all of the present solute atoms play a role in enhancing not
for GB cohesion but for GB embrittlement. In addition, the toughness and fracture
behaviours are influenced by the segregation energies despite some scattering in the
estimated values that resulted from using the empirical equations and the experimental
data in Figure 4. The Mg–Ca alloy with large segregation energies shows low tough-
ness with intergranular fracture due to the enhancement of GB embrittlement. On the
other hand, the alloys with small segregation energies, such as the Mg–Al alloy, have
high toughness and display the grain interior crack-propagation behaviour because of
the prevention for GB embrittlement. The solute atom segregations, such as for calcium
and yttrium atoms, in magnesium alloys are more GB embrittled as compared to those
in solute atoms, e.g. aluminium and zinc atoms.
Philosophical Magazine 4589

4. Discussion
We will consider what the influential material parameter is for the segregation effect on
GB embrittlement in the magnesium binary alloys. The elements that have a larger dif-
ference in their atomic radius between their solute and solvent atoms tend to segregate,
because the solute expulsion from the bulk reduces the elastic strain energy [23,26].
Table 1 shows that the calcium, zinc and yttrium atoms have larger atomic size differ-
ences than that of the magnesium atom (= 1.60  1010 m [14]). We can assume that
the addition of these atoms to magnesium promotes the segregation and the occurrence
of intergranular fracture. However, the Mg–Zn alloy has a higher toughness than that of
the Mg–Ca and –Y alloys and does not exhibit intergranular fracture. This discrepancy
occurs since these atoms exist in a solid solute state. The atomic size of solute and
solvent atoms in a solid solution state differs from that in a pure state due to the
electronic interaction. The addition of solute atoms and/or a concentration of these
Downloaded by [Georgetown University] at 13:38 02 February 2015

atoms lead to a change in the lattice parameters along the a-axis and the c-axis in the
HCP crystal. The variation in the lattice parameter c/a ratio with the additional concen-
tration has been obtained for several kinds of binary alloys by the X-ray analysis
[27,28]. The lattice parameter for the Mg–Ca alloy was obtained by the X-ray measure-
ment as shown in Figure A1, since there is not much literature for this alloy.
The relationship between the toughness and the absolute value of the change in the
lattice parameter ratio is described in Figure 5. There is a good relationship between the
change in the lattice parameter ratio and the toughness. The Mg–Ca alloy that shows a
large change in its lattice parameter c/a ratio exhibits low toughness resulting in
enhancing GB embrittlement, while the Mg–Zn alloy with a small change in the c/a
ratio displays high toughness. The addition of solute atoms, which produces a large
change in the lattice parameter ratio, causes a large misfit at the GBs, and hence, these
GBs are assumed to become the origin of fracture. In addition, this result is shown that
the solute atom, which has a large or small lattice parameter c/a ratio, enhances or
suppresses the GB segregation, and then increases or decreases the tendency for GB
embrittlement, respectively. A similar tendency has demonstrated that the yttrium atom
has a much stronger occurrence of segregation to GBs than the zinc atom in the

Figure 5. The relationship between the fracture toughness, JIC, and the difference in lattice
parameter c/a ratio with concentration in the magnesium binary alloys.
4590 H. Somekawa et al.

magnesium alloys [24]. On the other hand, from the viewpoint of electronic structure,
the covalent-like bonds enhance the GB embrittlement in metallic materials because of
the limitation of the atom mobility [29,30]. It is interesting that the charge densities of
the Mg–Y bonds are reported to be very high due to the covalent-like bonds, but their
bond strength becomes weakened and the bond characteristic is changed with the addi-
tion of aluminium and zinc atoms by the first principle calculation [31]. In the future, it
is necessary to understand the effect of alloying elements on the lattice parameter ratio
and the bond characteristic in the magnesium alloys. However, the present result is
worth addressing that the influential material parameter for enhancing or preventing GB
embrittlement is a change in the lattice parameter c/a ratio in the magnesium alloys.

5. Summary
Downloaded by [Georgetown University] at 13:38 02 February 2015

The effect of alloying elements on toughness was investigated using several kinds of
magnesium binary alloys with fine-grained structures in this study. The following results
were obtained.

(1) The alloying elements affect the toughness. The Mg–Zn alloy and the Mg–Ca
alloy have the highest and the lowest toughness, respectively, among the alloys.
These different tendencies resulted from solute segregation at GBs.
(2) The fracture toughness and fracture behaviour in magnesium alloys were closely
related to the segregation energy. The alloys with large segregation energies
show low toughness with intergranular fracture due to the enhancement for GB
embrittlement. On the other hand, the alloys with small segregation energies
have high toughness and display the grain interior crack-propagation behaviour
because of the prevention for GB embrittlement.
(3) The change in the lattice parameter c/a ratio was the influential material
parameter for enhancing or preventing GB embrittlement in the magnesium
alloys.

Acknowledgement
The authors are grateful to Dr Y. Osawa and Ms R. Komatsu (National Institute for Materials
Science) for his casting and her technical help, respectively. This work was partially supported by
the JSPS Grant-in-Aid in No. 23760675.

References

[1] K. Kubota, M. Mabuchi and K. Higashi, J. Mater. Sci. 34 (1999) p.2255.


[2] R. Armstrong, I. Codd, R.M. Douthwait and N.J. Petch, Philos. Mag. 7A (1962) p.45.
[3] M.A. Meyers, O. Vohringer and V.A. Lubarda, Acta Mater. 49 (2001) p.4025.
[4] H. Somekawa, A. Singh and T. Mukai, Philos. Mag. Lett. 89 (2009) p.2.
[5] A. Akhtar and E. Teghtssoonian, Philos. Mag. 25A (1972) p.897.
[6] J.A. Yasi, L.G. Hector and D.R. Trinkle, Acta Mater. 58 (2010) p.5704.
[7] H. Somekawa and T. Mukai, Mater. Sci. Eng. A561 (2013) p.378.
[8] H. Somekawa and T. Mukai, Scripta Mater. 53 (2005) p.1059.
Philosophical Magazine 4591

[9] M.M. Avedesian and H. Baker (eds.), ASM Specialty Handbook, Magnesium and Magne-
sium Alloys, ASM International, Materials Park, OH, 1999.
[10] M. Baskes and R.A. Johnson, Model. Simul. Mater. Sci. Eng. 2 (1994) p.147.
[11] X.W. Zhou, R.A. Johnson and H.N.G. Wadley, Phys. Rev. B 69 (2004) p.144113.
[12] M.R. Barnett, Z. Keshavarz, A.G. Beer and D. Atweel, Acta Mater. 52 (2004) p.5093.
[13] T.B. Massalski, Binary Alloy Phase Diagrams, 2nd ed., ASM International, Materials Park,
OH, 1990.
[14] W.F. Gale and T.C. Totemeier, Smithells Metals Reference Book, 8th ed., Elsevier, Oxford,
2004.
[15] R.A. Mulford, C.J. McMahon, D.P. Pope and H.C. Feng, Metall. Mater. Trans. A7 (1976)
p.1183.
[16] R. Wu, A.J. Freeman and G.B. Olson, Science 265 (1994) p.376.
[17] M. Yamaguchi, M. Shima and H. Kaburaki, Science 307 (2005) p.393.
[18] K. Horikawa, S. Kuramoto and M. Kanno, Acta Mater. 49 (2001) p.3981.
Downloaded by [Georgetown University] at 13:38 02 February 2015

[19] J.R. Rice and J.S. Wang, Mater. Sci. Eng. A107 (1989) p.23.
[20] J. Weissmuller, Nanostruct. Mater. 3 (1993) p.261.
[21] P. Choi, M. Silva, U. Klement, T. Al-Kassab and R. Kirchheim, Acta Mater. 53 (2005)
p.4473.
[22] M.P. Seah, J. Phys. F. 10 (1980) p.1043.
[23] D. McLean, Grain Boundaries in Metals, Oxford Press, London, 1957.
[24] J.P. Hadorn, K. Hantzsche, S. Yi, J. Bohlen, D. Letzing, J.A. Wollmershauser and S.R.
Agnew, Metall. Mater. Trans. A43 (2012) p.1347.
[25] T. Mukai, K. Hono, H. Somekawa, T. Honma, Magnesium alloy exhibiting high strength
and high ductility and method for production thereof, US patent publication, No.
US7871476B2.
[26] F.L. Williams and D. Nason, Surf. Sci. 45 (1974) p.377.
[27] R.S. Busk, J. Metals. 188 (1950) p.1460.
[28] S. Miura, S. Imagawa, T. Toyoda, K. Ohkubo and T. Mohri, Mater. Trans. 49 (2008) p.952.
[29] R. Haydock, J. Phys. C. 14 (1988) p.3807.
[30] L. Goodwin, R.J. Need and V. Heine, Phys. Rev. Lett. 60 (1988) p.2050.
[31] K. Chen and K. Boyle, Metall. Mater. Trans. A40 (2009) p.2751.

Appendix A

Three kinds of Mg–X at% Ca (X = 0.04, 0.3 and 0.6) alloys were used to obtain the lattice param-
eter c/a ratio according to the concentration for this system alloy. These alloys were prepared by
casting, and then were solution-treated at a temperature of 773 K for 2 h for the Mg–0.04Ca and
Mg–0.3Ca alloys and for 24 h for the Mg–0.6Ca alloy. The chemical compositions of calcium,
which were determined by the inductively coupled plasma mass spectrometry, were 0.06 wt.%
(=0.04 at.%), 0.47 wt.% (=0.30 at.%) and 0.94 wt.% (=0.57 at%). The X-ray diffraction (XRD)
measurements were performed using Cu-Kα radiation in these solution-treated alloys. The
measurement angle spacing, accelerating voltage and current were 0.02°, 40 kV and 300 mA,
respectively. The lattice spacing, d, in HCP metals was given by the following equation:

1 4 sin2 h 4 h2 þ hk þ k 2 l 2
¼ ¼  þ 2 ða1Þ
d2 k2 3 a2 c

The results of the XRD measurement from these alloys are shown in Figure A1. Using Equation
(a1) and λ = 0.1542 nm, the c/a ratios for the Mg–0.04Ca, 0.3Ca and 0.6Ca alloys were calculated
4592 H. Somekawa et al.
Downloaded by [Georgetown University] at 13:38 02 February 2015

Figure A1. Typical results of the XRD measurement from these three kinds of alloys.

to be 1.6243, 1.6249 and 1.6253, respectively. Although the analysed samples have only the three
different compositions, the concentration dependence on the lattice parameter c/a ratio was
determined to be 1.88 for the Mg–Ca alloy.

Appendix B

The simple model for describing the equilibrium segregation activity in the metallic materials has
been developed by McLean, as follows [23]:
 
Xgb X DEgb
¼ exp  ðb1Þ
1  Xgb 1  X kT

where Xgb is the solute atom concentration at GBs and k is the Boltzmann’s constant. The
previous paper has observed that the yttrium atoms are segregated at the GBs in the Mg–0.3 at.%
Y alloy, which is produced by the same method, i.e. extrusion. In addition, the concentration of
yttrium atoms at GBs and the matrix is measured to be about 0.9 and 0.3 at.%, respectively, by
TEM-nanoEDX analysis [25]. Using these measured values and the above Equation (b1), the GB
segregation energy, ΔEgb, is estimated to be 5.8 kJ/mol in the Mg–Y alloy. On the other hand,
the GB segregation energy is obtained to be 6.2 kJ/mol using Table 1 and equation (3) in the
text.

You might also like