Sei sulla pagina 1di 8

Available online at www.sciencedirect.

com

Acta Materialia 56 (2008) 387–394


www.elsevier.com/locate/actamat

Compressive deformation behavior at room temperature – 773 K


in Mg–0.2 mass%(0.035at.%)Ce alloy
a,*
Yasumasa Chino , Motohisa Kado b, Mamoru Mabuchi b

a
Materials Research Institute for Sustainable Development, National Institute of Advanced Industrial Science and Technology, 2266-98 Anagahora,
Shimo-shidami, Moriyama, Nagoya 463-8560, Japan
b
Department of Energy Science and Technology, Graduate School of Energy Science, Kyoto University, Yoshidahonmachi, Sakyo, Kyoto 606-8501, Japan

Received 15 June 2007; received in revised form 28 September 2007; accepted 28 September 2007
Available online 19 November 2007

Abstract

Ce is an efficient additive element because Ce addition makes Mg cold-workable. In the present work, compressive properties of
Mg–0.2 mass%(0.035 at.%)Ce alloy were investigated at room temperature – 773 K to understand the effects of Ce on deformation mech-
anisms. Ductility at room temperature was increased by Ce addition, but it decreased at 573 K. Thus, the very small Ce addition of
0.035 at.% strongly affected the mechanical properties of Mg. The unique behavior of the Mg–Ce alloy was related to homogeneous
deformation due to non-basal slips, although the c/a ratio of the Mg–Ce alloy was the same as that of pure Mg. It is therefore suggested
that the change in atomic binding states caused by Ce addition plays an important role in deformation mechanisms.
 2007 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

Keywords: Magnesium; Cerium; Mechanical properties; Shear bands; Recrystallization

1. Introduction It has been reported that high formability in Mg alloys is


attained at elevated temperatures [4–6]. However, it is
Lightweight materials can greatly contribute to a reduc- desirable to develop Mg alloys exhibiting high formability
tion in environmental load through light construction. In at/near room temperature from the point of view of form-
particular, Mg alloys have high potential for improving ing accuracy, cost and oxidation. To improve the formabil-
fuel efficiency and reducing CO2 emission because of their ity at/near room temperature, it is necessary to enhance the
high specific strength and stiffness [1]. In Mg, the (0 0 0 2) activity of the non-basal slips. In HCP metals, the c/a ratio
< 11 2 0 > basal slip takes place preferentially because affects the difference in CRSS between the basal slip and
the critical resolved shear stress (CRSS) for the basal slip the non-basal slips [3,7]; namely, a reduction in c/a ratio
is lower than those for the non-basal slips, that is, the pris- enhances the activity of the non-basal slips, where a and
matic and pyramidal slips [2]. Five independent slip sys- c are the lattice constants. According to previous works
tems are necessary to make a polycrystalline material [7–9], addition of Li, Y, La, Ce, Nd, Sm, Gd and Ca
undergo a general homogeneous deformation without pro- decreases the c/a ratio of Mg. In particular, Ce is one of
ducing cracks. However, the basal slip has only two inde- the interesting additive elements because its addition makes
pendent slip systems [3]. This gives rise to poor ductility Mg cold-workable [10]. However, adding too much of
and formability for polycrystalline Mg and its alloy, these elements causes poor ductility [11], in spite of a reduc-
thereby limiting their applications. tion in c/a ratio, because of an excessive solid-solution
strengthening mechanism. Hence, addition should be lim-
ited for the improvement of ductility. On the other hand,
*
Corresponding author. Tel.: +81 52 736 7461; fax: +81 52 736 7406. too little addition induces no solid-solution effects. Akhtar
E-mail address: y-chino@aist.go.jp (Y. Chino). et al. [12] showed that no solution-hardening is obtained

1359-6454/$30.00  2007 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.actamat.2007.09.036
388 Y. Chino et al. / Acta Materialia 56 (2008) 387–394

when the solute content is less than 0.025 at.%. Hence, in 1.7 · 103 s1. In addition, specimens with dimensions of
the present paper, 0.035 at.%Ce, which corresponds to 5 · 5 · 6 mm were prepared for observation of the speci-
0.2 mass%, is added to Mg, and its compressive properties men surface. The specimens were deformed to e = 5% at
are investigated at room temperature – 773 K. In addition, 573 K and the surfaces of planes parallel to the compres-
the effects of Ce on deformation mechanisms are discussed sion axis were observed using a scanning electron micro-
by comparing the compressive properties and microstruc- scope (SEM). For the compression tests at elevated
ture of the Mg–Ce alloy with those of pure Mg. temperatures, the specimens required 1.8 · 103 s to reach
equilibrium at the test temperature prior to straining. The
2. Experimental procedure temperature variation during the tests was less than
±1 K. Microstructure was observed using an optical micro-
Mg–0.2 mass%Ce alloy and pure Mg ingots were pre- scope and a transmission electron microscope (TEM). By
pared using an induction furnace in a graphite crucible the Schulz reflection method, the (0 0 0 2) pole figures were
under Ar atmosphere. The microstructures of the ingots examined using the compression axis cross section. The
are shown in Fig. 1. The grains were very coarse and of data were normalized by the (0 0 02) pole figure of a powder
the order of several hundred micrometers – a few millime- Mg datum. Lattice parameters for the a- and c-axes were
ters. Energy dispersive X-ray spectroscopy (EDS) measure- determined by XRD analysis using the ingots, that were
ments showed that there were very few precipitates and no annealed at 703 K for 7.2 · 104 s.
segregation in the Mg–Ce alloy.
Specimens with dimensions of /5 · 6 mm were cut from 3. Results
the ingots. Compression tests were carried out at room
temperature – 773 K with an initial strain rate of The true stress–true strain curves at room temperature
for the pure Mg and Mg–Ce alloy are shown in Fig. 2.
The Mg–Ce alloy exhibited a larger strain to failure than
the pure Mg. Note that a very small Ce addition of
0.035 at.% increased the ductility at room temperature.
Assuming that edge dislocations operate mainly during
deformation, an increase in yield stress by the solid-solu-
tion strengthening mechanism related to the misfit strain
can be given by
pffiffiffi
Dr ¼ 2mljej3=2 c1=2 ; ð1Þ
where Dr is the increase in yield stress caused by solid-solu-
tion strengthening, m is the Taylor orientation factor
(m = 3 for Mg [13]), l is the shear modulus
(l = 16.6 GPa for Mg [14]), e is the misfit strain and c is
the concentration of solid-solution. From Eq. (1), an
increase in yield stress caused by the 0.2 mass%Ce addition
is calculated to be 71 MPa, assuming that e = (rCe  rMg)/

Fig. 1. Microstructures of the specimens prior to the testing: (a) pure Mg Fig. 2. The true stress–true strain curves at room temperature for Mg and
and (b) Mg–0.2 mass%Ce alloy. Mg–0.2 mass%Ce alloy.
Y. Chino et al. / Acta Materialia 56 (2008) 387–394 389

rMg, where rCe and rMg are the atomic radii of Ce ence in tensile twins between the pure Mg and the Mg–
(1.83 · 1010 m [14]) and Mg (1.60 · 1010 m [14]), respec- Ce alloy.
tively. However, the flow stress of the Mg–Ce alloy was The microstructures of the region near the fracture sur-
almost the same as or slightly lower than that of the pure face of the specimens deformed to failure at room temper-
Mg. It is of interest to note that no solid-solution strength- ature are shown in Fig. 4, where Fig. 4a is the pure Mg and
ening was attained by the addition of Ce. This is probably Fig. 4b is the Mg–Ce alloy. Fracture is often accompanied
attributed to the activity of mobile screw dislocations [15– by the formation of the local deformation band in pure Mg
18]. Not only the compressive test, but also the tensile test [23]. Inspection of Fig. 4a reveals that local deformation
at room temperature, showed that the Mg–0.2 mass%Ce bands were formed and cracks were developed along the
alloy exhibited a larger strain to failure than the pure deformation bands in the pure Mg. However, such local
Mg, but almost the same flow stress as the pure Mg [19]. deformation bands were not observed in the Mg–Ce alloy
Therefore, it is conclusively demonstrated that a very small (Fig. 4b). The local deformation bands consist of a large
amount of Ce addition enhances the ductility, not the number of compressive twins with narrow banded mor-
strength, of Mg at room temperature. phology [10,24]. The compressive twins were observed in
The microstructures of the specimens deformed to the Mg–Ce alloy as well as the pure Mg, but the twins were
e = 2% at room temperature are shown in Fig. 3, where distributed more homogeneously in the Mg–Ce alloy.
Fig. 3a is for the pure Mg and Fig. 3b is for the Mg–Ce Therefore, it is likely that the deformation is more homoge-
alloy. It is known that f1 0 1 2g tensile twins are generated neous for the Mg–Ce alloy than for the pure Mg, resulting
easily during compressive tests in Mg [20–22]. Many tensile in the suppression of cracking and higher ductility for the
twins were observed in both the pure Mg and the Mg– Mg–Ce alloy.
0.2 mass%Ce alloy. It appears that there was little differ-

Fig. 4. Microstructures of the region near the facture surface of


Fig. 3. Microstructures of the specimens deformed to e = 2% at room the specimens deformed to failure at room temperature: (a) pure Mg and
temperature: (a) pure Mg and (b) Mg–0.2 mass%Ce alloy. (b) Mg–0.2 mass%Ce alloy.
390 Y. Chino et al. / Acta Materialia 56 (2008) 387–394

The (0 0 0 2) pole figures of the specimens deformed to


e = 10% at room temperature are shown in Fig. 5, where
Fig. 5a is the pure Mg specimen and Fig. 5b is the Mg–
Ce alloy. In general, the basal slip occurs preferentially
and the grains have a tendency to rotate in a direction
where the basal planes are perpendicular to the compres-
sion axis [7,25,26]. As expected, in the pure Mg, the basal
planes were intensively oriented in a direction perpendicu-
lar to the compression axis. The same trend was observed
in the Mg–Ce alloy, but the texture intensity of the Mg–
Ce alloy was about one-third that of the pure Mg. Barnett
et al. [10] also reported a reduction in texture intensity
caused by Ce addition. It is of interest to note that there
were other two peaks besides the peak perpendicular to
the compression axis in the Mg–Ce alloy. This suggests that
not only the basal slip, but also other deformation mecha-
Fig. 6. The true stress–true strain curves at 573 K for Mg and Mg–
nisms such as the non-basal slips are active for the Mg–Ce 0.2 mass%Ce alloy.
alloy. The reduction in (0 0 0 2) texture intensity gives rise to
the improvement of press formability [19].
The true stress–true strain curves at 573, 673 and 773 K
for the pure Mg and the Mg–0.2 mass%Ce alloy are shown
in Figs 6–8, respectively, where the tests were stopped when
the specimens were deformed to 50% without failure. At
573 K, the pure Mg was deformed to e = 50% without fail-
ure; however, the Mg–Ce alloy was fractured at e = 40%.
Also, the flow stress of the Mg–Ce alloy was higher than
that of the pure Mg, as shown in Fig. 6. Thus, the deforma-

Fig. 7. The true stress–true strain curves at 673 K for Mg and Mg–
0.2 mass%Ce alloy.

Fig. 5. The (0 0 0 2) pole figures of the specimens deformed to e = 10% at Fig. 8. The true stress–true strain curves at 773 K for Mg and Mg–
room temperature. 0.2 mass%Ce alloy.
Y. Chino et al. / Acta Materialia 56 (2008) 387–394 391

tion characteristics at 573 K were significantly different


from those at room temperature. However, the difference
in flow stress between the pure Mg and the Mg–Ce alloy
decreased by 773 K, as shown in Fig. 8.
Previous works [27,28] showed that creep resistance or
high-temperature strength is increased by the addition of
Ce due to the precipitation of Mg9Ce. However, the precip-
itates were not found in the Mg–0.2 mass%Ce alloy inves-
tigated in the present work. Therefore, the higher flow
stress at elevated temperatures for Mg–Ce is not attributed
to precipitation.

4. Discussion

It is known that the f1 0


1 2g tensile twins and f1 01 1g or

f3 03 2g compressive twins are formed at the early and later
stage during compression tests at room temperature in Mg
and its alloys, respectively [3,6,26,29,30]. Because the com-
pression twins induce a large shear deformation within the
twin volume [29], crack initiation often takes place at their
boundaries [31] and their localization causes premature
failure. In fact, in the pure Mg, local deformation bands
consisting of the compressive twins were observed and
cracks were developed along the bands. In the Mg–Ce
alloy, however, the deformation was homogeneous and
the formation of local deformation bands was suppressed.
An investigation of the stretch formability of rolled Mg–
0.2 mass%Ce alloy at room temperature and 433 K showed
that enhancement of stretch-formability by the small Ce
addition is attributed to the activity of the non-basal slip
such as the prismatic slip [19]. Thus, improvement of duc-
tility at room temperature by Ce addition is attributed to
the enhancement of the non-basal slips. However, the Fig. 9. SEM images of the surface of specimens deformed to e = 5% at
573 K: (a) pure Mg and (b) Mg–0.2 mass%Ce alloy.
Mg–Ce alloy exhibited poorer ductility at 573 K than the
pure Mg. This cannot be explained by the enhancement
of the non-basal slips due to Ce addition. slip, is activated at room temperature. Reed-Hill et al.
SEM images of the surface of specimens deformed to [36] showed that prismatic slip occurred at 83–298 K,
e = 5% at 573 K are shown in Fig. 9, where Fig. 9a is the whereas the pyramidal slip was activated at 423–559 K.
pure Mg and Fig. 9b is the Mg–Ce alloy. In the pure Hence, the non-basal slip observed at 573 K in the Mg–
Mg, straight lines were observed. Choi et al. [32] showed Ce alloy is probably the pyramidal slip.
that parallel slip-banding was observed in the surface Enhancement of the non-basal slip by the addition of Ce
grains with very few slip systems with high Schmid factors. is responsible for the improvement of ductility at room
Therefore, it is clear from Fig. 9a that the basal slip mainly temperature. However, the decrease in ductility caused by
occurred in the pure Mg. In the Mg–Ce alloy, however, the the addition of Ce found at 573 K cannot be explained
lines were wavy. The wavy lines result from the cross slips by the enhancement of the non-basal slip. Also, from
of dislocations related to the non-basal slips [33]. EDS measurements, there were very few precipitates and
TEM images of the specimens deformed to e = 5% at no segregation in the Mg–Ce alloy specimen deformed at
573 K are shown in Fig. 10, where Fig. 10a is the pure 573 K as well as the specimen prior to testing.
Mg and Fig. 10b is the Mg–Ce alloy. Dislocations in the Mechanical properties of Mg strongly depend on the
pure Mg were roughly parallel to each other; however, grain size, particularly at elevated temperatures [37]
there were many dislocations that were not parallel in the because grain-boundary sliding (GBS) occurs easily at ele-
Mg–Ce alloy. Similar results were obtained in the pure vated temperatures in Mg [38]. This is because of the large
Mg and Mg–Al alloy [34,35]. The results of the SEM and grain boundary diffusion coefficient of Mg [38]. It is known
TEM observations indicate that the non-basal slips are that the diffusion coefficient of Ce in Mg is lower than the
more active in the Mg–Ce alloy than in the pure Mg at self-diffusion coefficient of Mg [39]. Therefore, the unique
573 K as well as at room temperature. Koike et al. [17] deformation behavior at 573 K is suggested to be associ-
showed that the prismatic Æaæ, in addition to the basal Æaæ ated with diffusion-related events.
392 Y. Chino et al. / Acta Materialia 56 (2008) 387–394

Fig. 10. TEM images of the specimens deformed to e = 5% at 573 K: Fig. 11. Microstructures of the specimens deformed to e = 5% at 573 K:
(a) pure Mg and (b) Mg–0.2 mass%Ce alloy. (a) pure Mg and (b) Mg–0.2 mass%Ce alloy.

The microstructures of the specimens deformed to however, no recrystallized fine grains were observed.
e = 5% at 573 K are shown in Fig. 11, where Fig. 11a is GBS relaxes the stress concentration [41]. As shown
the pure Mg and Fig. 11b is the Mg–Ce alloy. Tensile twins in Fig. 6, the strain-hardening coefficient for the pure
were observed, but recrystallized fine grains were not found Mg was lower than that for the Mg–Ce alloy. This
in both the pure Mg and the Mg–0.2 mass%Ce alloy, indi- is attributed to the GBS due to the recrystallized fine
cating that the difference in flow stress between the pure grains for the pure Mg. Thus, the higher flow stress
Mg and the Mg–Ce alloy is not attributed to GBS at least and poor ductility for the Mg–Ce alloy are attributed
until 5%. Inspection of Fig. 6 reveals that the flow stress at to suppression of DRX as well as the low diffusion
e = 5% at 573 K for the Mg–Ce alloy was about twofold due to Ce addition.
that for the pure Mg. Therefore, the higher flow stress Sitdikov et al. [33] showed that mechanisms of DRX in
for the Mg–Ce alloy is likely to be due to the low diffusion Mg depend on the temperature, namely, twin DRX at 293–
coefficient of Ce in Mg. 623 K, continuous DRX at 523–773 K, and discontinuous
The microstructures of the specimens deformed to a DRX at more than 773 K. As shown in Fig. 11, many twins
large strain at 573 K are shown in Fig. 12, where were observed in the specimens deformed at 573 K. Hence,
Fig. 12a is the pure Mg at e = 50% and Fig. 12b is twin DRX must mainly occur in the deformed specimens.
the Mg–Ce alloy at e = 33%. Note that recrystallized Numerous twins were locally generated at a late stage dur-
fine grains were observed along the grain boundaries ing the compression test in the pure Mg (Fig. 4). The twins
in the pure Mg. Ion et al. [40] showed that deforma- in the local deformation bands induce DRX in the pure
tion becomes macroscopically inhomogeneous at a large Mg. On the other hand, the deformation was more homo-
strain at temperatures below 600 K and is confined to geneous for the Mg–Ce alloy than for the pure Mg, which
shear zones, which are fine-grained regions formed by is due to the active non-basal slip for the Mg–Ce alloy.
dynamic recrystallization (DRX). In the Mg–Ce alloy, Therefore, the homogeneous deformation due to the
Y. Chino et al. / Acta Materialia 56 (2008) 387–394 393

Fig. 12. Microstructures of the specimens deformed to large strain at Fig. 13. Microstructures of the specimens deformed to e = 50% at 773 K:
573 K: (a) pure Mg at e = 50% and (b) Mg–0.2 mass%Ce alloy at e = 33%. (a) pure Mg and (b) Mg–0.2 mass%Ce alloy.

The present work demonstrated that the Mg–Ce alloy


enhancement of non-basal slips is responsible for the sup- exhibits higher ductility at room temperature than the pure
pression of DRX by the addition of Ce. Mg. Recently, Barnett et al. [10] reported that Mg–
The microstructures of the specimens deformed to 0.2 mass%Ce alloy is considerably more rollable than pure
e = 50% at 773 K are shown in Fig. 13, where Fig. 13a Mg and AZ31 alloy at room temperature. The unique char-
is the pure Mg and Fig. 13b is the Mg–Ce alloy. It can acteristics of the Mg–Ce alloy are attributed to the
be seen that DRX occurred in both the pure Mg and enhancement of non-basal slips by Ce addition. However,
the Mg–Ce alloy, and the grain size of the Mg–Ce alloy the measurement of the lattice constants showed that the
was almost the same as that of the pure Mg, indicating c/a ratio of the Mg–0.2 mass%(0.035 at.%)Ce alloy is
that the addition of Ce has no effect on DRX at 773 K. 1.6247, which is the same as that of the pure Mg
The mechanism of DRX at 773 K is not twin DRX, but (=1.6244). Note that the very small Ce addition had no
discontinuous DRX [33]. Discontinuous DRX is closely effect on the c/a ratio, but it had an effect on the enhance-
related to dislocations rather than twins. The difference ment of non-basal slips.
in CRSS between the basal slip and the non-basal slips The range of a very large stress field caused by lattice
decreases with increasing temperature, and the CRSS of strain around a dislocation core, which cannot be analyzed
the basal slip is close to those of the non-basal slips. by the theory of elasticity, is generally about 5b [42], where
Therefore, the non-basal slip is sufficiently active at b is the Burgers vector. The atomic radius of Ce is relatively
773 K for the pure Mg as well as the Mg–Ce alloy, and larger than that of Mg, and an atom of Ce may affect the
DRX occurs in a similar manner in both the pure Mg field within 5b or more. The present investigation demon-
and the Mg–Ce alloy. This results in the same strain-hard- strated that only 0.035 at.%Ce affects the mechanical prop-
ening coefficient at 773 K in the pure Mg and the Mg–Ce erties of Mg. If 0.035 at.%Ce affects all Mg atoms, the
alloy, as shown in Fig. 8. range of a field affected by an atom of Ce is about 7b, that
394 Y. Chino et al. / Acta Materialia 56 (2008) 387–394

is, about 2 nm. Therefore, it is suggested that because an 4. The c/a ratio of the Mg–Ce alloy was the same as that of
atom of Ce has an effect up to a distance of 2 nm, only the pure Mg. The change in atomic binding states
0.035 at.%Ce addition has a distinguishable effect on the caused by the addition of Ce is suggested to play a vital
mechanical properties of Mg. All additive elements do role in the enhancement of non-basal slips.
not necessarily have such strong effects on the deformation
mechanisms caused by the small addition. Tegart [43] sug- References
gested that the valency differences play an important part
in the solid-solution strengthening of Mg. Hence, not only [1] Hakamada M, Furuta T, Chino Y, Chen Y, Kusuda H, Mabuchi M.
the atomic size effect, but also the electron state between Energy 2007;32:1352.
the atoms should be considered to understand the effects [2] Yoshinaga H, Horiuchi R. Trans JIM 1963;4:1.
[3] Yoo MH. Metall Trans A 1981;12:409.
of small Ce addition. Our preliminary ab initio calculations
[4] Watanabe H, Mukai T, Kohzu M, Tanabe S, Higashi K. Acta Mater
showed that an atom of Ce affects the electron density 1999;47:3753.
states not only around the Ce atom, but also around Mg [5] Kim WJ, Chung SW, Chung CS, Kum D. Acta Mater 2001;49:
atoms near the Ce atom. Although a small addition of 3337.
0.035 at.% has no effect on the c/a ratio, it affects the [6] Matsubara K, Miyahara Y, Horita Z, Langdon TG. Acta Mater
atomic binding states not only between the Ce atom and 2003;51:3073.
[7] Agnew SR, Yoo MH, Tome CN. Acta Mater 2001;49:4277.
the Mg atom, but also between the Mg atoms. McDonald [8] Busk RS. Trans AIME 1950;188:1460.
[44] showed that the Mg–0.2 mass%Ce is softened by cold- [9] Hehmann F, Sommer F, Predel B. Mater Sci Eng A 1990;125:249.
working, but it is hardened by subsequent heating. Such an [10] Barnett MR, Nave MD, Bettles CJ. Mater Sci Eng A 2004;386:205.
anomalous behavior of Mg containing a small amount of [11] McDonald JC. Trans AIME 1941;138:179.
[12] Akhtar A, Teghtsoonian E. Acta Metall 1969;17:1339.
Ce may be related to the change in atomic binding states
[13] Suzuki M, Sato H, Maruyama K, Oikawa H. Mater Sci Eng A
caused by Ce addition. Further research is in progress to 1998;252:248.
understand the relationship between the change in atomic [14] Jpn Inst Metals, editor. Kinzoku data book, revised 3rd ed.
binding states caused by Ce addition and the mechanical [Japanese], Maruzen, Tokyo; 1993.
properties. It is suggested from the results in the present [15] Yoshinaga H, Horiuchi R. Trans JIM 1963;5:14.
paper that the change in atomic binding states caused by [16] Obara T, Yoshinaga H, Morozumi S. Acta Metall 1973;21:845.
[17] Koike J, Kobayashi T, Mukai T, Watanabe H, Suzuki M, Maruyama
Ce addition gives rise to the enhancement of non-basal K, Higashi K. Acta Mater 2003;51:2055.
slips, resulting in the unique mechanical properties of the [18] Galiyev A, Sitdikov O, Kaibyshev R. Mater Trans 2003;44:426.
Mg–Ce alloy. [19] Chino Y, Kado M, Mabuchi M, to be submitted for publication.
[20] Ball EA, Prangnell PB. Scripta Metall Mater 1994;31:111.
5. Conclusions [21] Kleiner S, Uggowitzer PJ. Mater Sci Eng A 2004;379:258.
[22] Wang YN, Huang JC. Acta Mater 2007;55:897.
[23] Klimanek P, Pötzsch A. Mater Sci Eng A 2002;324:145.
The mechanical properties of Mg–0.2 mass% [24] Couling SL, Pashak JF, Sturkey L. Trans ASM 1959;51:94.
(0.035 at.%)Ce alloy and pure Mg were investigated by [25] Wonsiewich BC, Backofen WA. Trans Metall Soc AIME
conducting compression tests at room temperature – 1967;239:1422.
[26] Brown DW, Agnew SR, Bourke MAM, Holden TM, Vogel SC, Tome
773 K. The results are summarized as follows:
CN. Mater Sci Eng A 2005;399:1.
[27] Mellor GA, Ridley RW. J Inst Met 1952–53;81:245.
1. The Mg–Ce alloy showed higher ductility at room tem- [28] Leontis TE, Murphy JP. Trans AIME 1946;166:295.
perature than the pure Mg. On the other hand, the flow [29] Koike J. Metall Mater Trans A 2005;36:1689.
stress of the Mg–Ce alloy was almost the same as that of [30] Yoo MH, Lee JK. Philos Mag A 1991;63:987.
the pure Mg. [31] Kucherov L, Tadmor EB. Acta Mater 2007;55:2065.
[32] Choi YS, Piehler HR, Rollett AD. Metall Mater Trans A
2. In the pure Mg, local deformation bands consisting of 2004;35A:513.
compressive twins were observed and cracks were devel- [33] Sitidikov O, Kaibyshev R. Mater Trans 2001;42:1928.
oped along the bands. In the Mg–Ce alloy, however, the [34] Vagarali SS, Langdon TG. Acta Metall 1981;29:1969.
deformation was more homogeneous and the formation [35] Vagarali SS, Langdon TG. Acta Metall 1982;30:1157.
[36] Reed-Hill RE, Robertson WD. Trans AIME 1957;209:496.
of local deformation bands was suppressed due to the
[37] Mabuchi M, Chino Y, Iwasaki H. Mater Trans 2003;44:490.
enhancement of non-basal slips. This is responsible for [38] Mabuchi M, Chino Y, Iwasaki H. Mater Trans 2002;43:2063.
the improvement of ductility at room temperature by [39] Fujikawa S. J Jpn Inst Light Met 1992;42:822.
Ce addition. [40] Ion SE, Humphereys FJ, White SH. Acta Metall 1982;30:1909.
3. On the other hand, the Mg–Ce alloy exhibited higher [41] Koike J, Ohyama R, Kobayashi T, Suzuki M, Maruyama K. Mater
Trans 2003;44:445.
flow stress and poorer ductility at 573 K than the pure
[42] Kato M. Introduction to the theory of dislocations. Tokyo: Shok-
Mg. This is attributed to the suppression of dynamic abo; 2003.
recrystallization and the low diffusion due to Ce [43] Tegart WJMcG. J Inst Met 1960;89:215.
addition. [44] McDonald JC. Trans Metall Soc AIME 1958;212:45.

Potrebbero piacerti anche