You are on page 1of 9

International Journal of Lightweight Materials and Manufacture 1 (2018) 256e264

Contents lists available at ScienceDirect

International Journal of Lightweight Materials and Manufacture


journal homepage: https://www.sciencedirect.com/journal/
international-journal-of-lightweight-materials-and-manufacture

Original Article

In-situ observations of recrystallization and microstructural evolution


in cerium-containing rolled magnesium alloys
ndez-Escobar, Carl J. Boehlert*
Ajith Chakkedath, David Herna
Department of Chemical Engineering and Materials Science, Michigan State University, East Lansing, MI, 48824, USA

a r t i c l e i n f o a b s t r a c t

Article history: The recrystallization and microstructural evolution of rolled Mg-2Zn-xCe (x ¼ 0.2 and 0.6 wt%) were
Received 6 April 2018 studied by combining annealing with in-situ electron backscatter diffraction (EBSD) analysis. Sequential
Received in revised form EBSD orientation maps of the same microstructural patch were obtained at increasing temperatures,
19 September 2018
298 K, 423 K, and from 473 K to 598 K in 25 K increments. In both alloys, the nucleation of new grains
Accepted 20 September 2018
Available online 29 September 2018
was first observed at 473 K, and recrystallization was complete by 573 K. It was found that greater Ce
additions led to a higher fraction of recrystallized grains and an overall finer grain size. New grains with a
misorientation angle-axis relationship of 60 about <1010>, 56 about <1120>, and 86 about <1120>
Keywords:
Recrystallization
were observed. Although the twin evolution during heating was not captured, the misorientation axis
Microstructure relationships along <1010>, <1120>, <1011> were characterized for the recrystallized grains. The
Magnesium alloys misorientation angles were uniformly distributed between 30 and 90 in both alloys. The results are
Texture discussed and compared with Mg-3Al-1Zn (wt%).
In-situ microscopy © 2018 The Authors. Production and hosting by Elsevier B.V. on behalf of KeAi Communications Co., Ltd.
This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-
nc-nd/4.0/).

1. Introduction processing has been demonstrated previously [6,13], an investiga-


tion on the underlying mechanisms responsible for this
The crystallographic texture of wrought magnesium (Mg) alloys effect during annealing is lacking. Explanations range from differ-
has a significant influence on the elongation-to-failure [1,2] and the ences in the activity of deformation mechanisms, such as slip and
anisotropic mechanical behavior [3]. The development of micro- twinning [7,14e16] and their influence on orientation changes due
structure and texture during sheet rolling of Mg alloys is known to to dynamic or static recrystallization during the rolling process
be dependent on the influence of specific alloying elements. [8,17]. While in binary Mg-RE alloys a distinct weakening of the
However, the underlying mechanisms responsible for the texture basal texture is observed, in ternary Mg-Zn-RE alloys a different
development in Mg alloys during annealing are not yet understood texture develops. In this texture, orientations with the c-axis par-
[4,5]. Many works have shown that rare-earth (RE) element addi- allel to the normal direction (ND) of the sheet almost fully vanish,
tions lead to the development of weaker textures during rolling and but a weak texture with the c-axis tilted towards the transverse
annealing, and in some cases strong alignment of basal planes direction (TD) remains. The significance of this development de-
parallel to the sheet plane are not observed [5e10]. Without the RE pends on the content of the RE [18]. It has been hypothesized that
additions, conventional Mg alloys tend to form strong texture this is an effect of competitive grain growth with different orien-
during wrought processing and retain that texture after annealing tations during recrystallization, however, this has not been directly
[11,12], which makes further processing difficult. Thus, the ability observed.
to control the crystallographic texture in wrought Mg alloys is of In-situ scanning electron microscopy (SEM) is a valuable tool to
commercial interest. study the evolution of microstructure under various loading con-
Although the effect of RE element addition or combinations ditions. The ability of modern SEMs to incorporate heating as-
with Zn on the texture development in Mg alloys during wrought semblies into the stage, together with the fast indexing capabilities
of fully-automated electron backscatter diffraction (EBSD) systems,
* Corresponding author. enables capturing the evolving microstructure during in-situ
E-mail address: boehlert@egr.msu.edu (C.J. Boehlert). heating experiments. A review of in-situ EBSD heating experi-
Peer review under responsibility of Editorial Board of International Journal of
ments was provided by Wright et al. [19]. Such experiments are
Lightweight Materials and Manufacture.

https://doi.org/10.1016/j.ijlmm.2018.09.004
2588-8404/© 2018 The Authors. Production and hosting by Elsevier B.V. on behalf of KeAi Communications Co., Ltd. This is an open access article under the CC BY-NC-ND
license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
A. Chakkedath et al. / International Journal of Lightweight Materials and Manufacture 1 (2018) 256e264 257

typically performed to understand the phase transformations


and/or recrystallization behavior as a function of temperature
and/or time. In-situ heating experiments have been performed to
study microstructural evolution and recrystallization in aluminum
alloys [20e24], copper [25], tantalum [26], zirconium alloys [26],
titanium [20,27], and steel [20,28]. Similar studies can be beneficial
to understand the microstructural evolution in wrought Mg alloys,
in which the crystallographic texture has a significant impact on the
elongation-to-failure [1,2] and the anisotropic mechanical behavior
[3]. In this study, a novel in-situ annealing technique was
developed to understand the mechanisms responsible for texture
weakening observed during wrought processing in two alloys
studied previously [18], Mg-2Zn-0.2Ce (wt.%) and Mg-2Zn-0.6Ce
(wt.%) {referred to as ZE20 and ZE21, respectively, according to
the ASTM standards for naming Mg alloys}.

2. Experimental methods

The measured composition of the ZE20 and ZE21 alloys were Fig. 1. Heating profile of the in-situ EBSD experiments.
Mg-1.9Zn-0.2Ce (wt.%) and Mg-2.0Zn-0.5Ce (wt.%), respectively.
The alloys were gravity cast into billets. Samples having dimensions
of 280  50  20 mm (length x width x thickness) were cut in The resulting EBSD data were analyzed using EDAX TSL OIM
preparation for rolling. A homogenization heat treatment of 15 h at Analysis (v6.1.3) software. After acquiring the raw data, post-
623 K was carried out prior to rolling. The rolling experiment was processing clean-up procedures were performed to remove erro-
performed using a rolling mill with a maximum load of 50 tons and neous data points formed due to un-indexed or inappropriately
a roll diameter of 400 mm. The rolling was carried out at 673 K at a indexed patterns. A single iteration of a “grain dilation” clean-up
rolling rate of 16 m/min, and consisted of a total of fourteen passes procedure was performed. It is noted that the input parameters
with an increasing degree of deformation, 4. For the first four for the clean-up procedures were selected based on the overall
passes 4 ¼ 0.2. For the next seven passes 4 ¼ 0.2. For the last three average confidence index value of the raw data in an effort to
passes 4 ¼ 0.3. The final thickness of the sheets was ~1.3 mm. After minimize the number of points modified. In the case of scans taken
each pass, the sheets were reheated for 20 min at 673 K. After the at temperatures less than 523 K, approximately 20e30% points in
final rolling pass, the sheets were air cooled. Sheets in this as-rolled the EBSD maps were modified during the clean-up procedure. For
condition were used for the in-situ annealing experiments. scans taken at temperatures greater than 523 K, approximately
Flat rectangular samples of 15  10 mm (length x width) 15e25% of the EBSD data points were changed during the clean-up.
were cut from the as-rolled sheets using a diamond saw. The This was because the quality of the EBSD indexing increased with
samples were mechanically polished using silicon carbide planar an increase in temperature as the microstructure consisted mainly
grinding papers through 800, 1200, 2400, and 4000 grit. The of newly recrystallized (‘strain-free’) grains. Furthermore, the EBSD
specimens were then finely polished through 6, 3 and 1 mm data were partitioned with a grain tolerance angle of 5 and a
diamond paste sequentially. Colloidal silica solution, with a minimum grain size of 2 pixels. The average grain size was calcu-
0.04 mm particle size, was used for the final mechanical pol- lated from the EBSD maps using the appropriate functions in the
ishing. Ethanol was used both as a lubricant and as a cleanser for EDAX TSL OIM Analysis software.
the specimen. To further improve the quality of the sample
surface for EBSD, the specimens were electropolished using a 3. Results
Struers TenuPol-5 double jet system and a solution of 30% nitric
acid and 70% methanol as an electrolyte. The temperature of the 3.1. Ex-situ analysis of microstructure and texture
electrolyte was kept below 248 K during electropolishing, which
was performed at a voltage of approximately 12 V and a corre- Fig. 2 shows the EBSD inverse pole figure (IPF) map of the
sponding current of ~12 mA. microstructure and the corresponding {0001} pole figures along
The experimental setup and the techniques used to heat and the normal direction, showing the texture in the as-rolled condition
characterize the same microstructural patch, of dimensions of and after annealing at 673 K for 60 s for ZE20 and ZE21. As shown
approximately 100  150 mm, were described in detail in Ref. [29]. in Fig. 2a and c, elongated grains along the rolling direction (RD)
The test temperature was controlled using a constant-voltage were observed in the as-rolled microstructures of both materials
power supply to a tungsten-based heating element placed below and a recrystallized microstructure was observed after annealing,
the sample. The temperature was monitored using a K-type ther- see Fig. 2b and d. It is noted that these images were taken from
mocouple spot-welded to the side of the sample. The EBSD data different sections of the material, and thus they do not represent
was collected using a 0.5 mm step size and a EDAX TSL (Mahwah, NJ) the same microstructural patches before and after annealing. The
OIM Data Collection software connected to a Tescan Mira3 SEM, average grain diameters for ZE20 and ZE21 after annealing at 673K
operated at 25 kV. The working distance was approximately 15 mm. for 60s were approximately 10 mm and 6 mm, respectively. Both of
The heating profile followed during the in-situ experiments is the as-rolled alloys exhibited a basal texture in which the c-axis
shown in Fig. 1. Each EBSD scan was collected for approximately tended to align close to ND but with a certain tilt to either RD or TD
50 min, where heating was performed at a rate of ~2.5 K/min, and (see Fig. 2a and c). A stronger basal texture was observed in ZE20
the temperature was stabilized for approximately 15 min before than ZE21. This is in agreement with previous findings that higher
starting the EBSD scan. It should be noted that for ZE21, an addi- Ce contents result in texture weakening [6,7]. The maximum
tional heating step, from 598 K to 623 K, was included with the texture intensity decreased in both materials after annealing (from
same heating rate and stabilization time. 13.7 to 6.5 times random in the as-rolled condition to 7.8 and 4.9
258 A. Chakkedath et al. / International Journal of Lightweight Materials and Manufacture 1 (2018) 256e264

Fig. 2. The EBSD IPF map of the microstructure and the texture in the form of {0001} pole figures along the normal direction for (a) as-rolled ZE20, (b) ZE20 after annealing, (c)
as-rolled ZE21, and (d) ZE21 after annealing. The annealing was carried out at 673K for 60s for both materials. The rolling direction is vertical. It is noted that the images shown
above were taken from different sections of the material, thus they do not represent the same microstructural patches before and after annealing.

times random after annealing in ZE20 and ZE21, respectively). Thus, of misorientation values greater than 1.5 within the grains in the
this annealing treatment helped weaken the texture. In order to areas analyzed for both the alloys as a function of annealing tem-
understand the mechanisms responsible for the texture weakening perature. For both alloys, after the annealing treatments performed
during annealing, an identical microstructural patch was moni- at temperatures greater than or equal to 573K, the grains were
tored during annealing using the in-situ experimental technique relatively free of the strain.
described in the experimental section. The orientation relationship of the new grains, formed during
each heating step, was investigated with respect to their neighbors.
3.2. In-situ analysis of recrystallization and microstructural Specifically, the misorientation angles and the corresponding
evolution misorientation axis across the newly-formed grain boundaries were
examined. The grain boundaries with misorientation angles greater
Fig. 3 shows the EBSD IPF map of the same microstructural patch than 15 were only considered in the analysis. Fig. 6 shows the dis-
and the corresponding texture in the form of {0001} pole figures, tribution of the number fraction of grain boundaries with specific
along the ND, depicting the evolution of the microstructure in rotation axis for the newly-formed grains in ZE20 (for heating steps
rolled ZE21 as a function of the annealing temperature. For com- between 423 and 548 K) and ZE21 (for heating steps between 423
parison, the corresponding EBSD IPF maps and {0001} pole figures and 598 K). Among the misorientation relationships observed be-
for ZE20 are provided in Ref. [29]. During the annealing process, tween the newly-formed grain boundaries, rotation axes about
new grains started to appear during the heating step from 423K to <1010>, <1120>, and <1011> were prevalent. Grain boundaries
473K in both alloys. At ~573 K, a completely recrystallized micro- with orientation relationships corresponding to f1012g extension
structure was observed for both alloys. For ZE20, a total of 59 new twinning (86 about <1120>), f1011g contraction twinning (56
grains were observed for the heating steps between 423 and 548 K. about <1120>), and ð1012Þ-ð0112Þ extension double twinning (60
For ZE21, a total of 112 new grains were observed for the heating about <1010>) were also observed. This was expected due to the
steps between 423 and 598 K. recovery and growth of the twins, which formed during the rolling
Fig. 4 shows the grain orientation spread obtained from the process. Fig. 7 shows the distribution of the number fraction of the
EBSD data from the microstructural patch analyzed for ZE21 as a observed misorientation axis relationships for the newly-formed
function of annealing temperature. As shown in Fig. 4, the grain grain boundaries in ZE20 and ZE21, sorted into <hki0> and <hkil>
orientation spread was less than 1.5 in almost all the area analyzed types. In this work, <hki0> is referred as a crystallographic family of
after the heat treatment at 573K, suggesting that the grains were directions in which the “l” coefficient equals zero (l ¼ 0). Overall, the
relatively free of the strain. Fig. 5 shows the fraction distribution of the misorientation axis across the newly-formed grain
A. Chakkedath et al. / International Journal of Lightweight Materials and Manufacture 1 (2018) 256e264 259

Fig. 3. EBSD IPF maps and the corresponding {0001} pole figures, along the ND, of the same microstructural patch depicting the microstructural evolution in rolled ZE21 as a
function of annealing temperature. Imax is the maximum intensity value observed in the pole figures. Black regions in the maps are un-indexed points. The rolling direction is
horizontal.
260 A. Chakkedath et al. / International Journal of Lightweight Materials and Manufacture 1 (2018) 256e264

Fig. 3. (continued).

boundaries maintained a proportion of approximately 70% <hki0> that the basal texture of AZ31B was strengthened by grain growth
and approximately 30% <hkil>. during annealing.
In the microstructural patch monitored during the in-situ
experiments, 159 and 300 new grain boundaries from 59 to 112
4. Discussion
newly-formed grains were observed in ZE20 and ZE21, respectively.
To obtain a more statistically meaningful dataset, a larger area from
It was observed that the texture intensity increased slightly for
the final microstructure (after the last annealing step) was EBSD
the initial heating step, from 298 K to 423 K maximum, see Figs. 3
analyzed for both alloys. Furthermore, the final microstructure of
and 8. The reason for this slight increase is not clear and is unex-
the rolled ZE20 and ZE21 was compared with a rolled conventional
pected as the original 298K microstructure was not significantly
Mg alloy, Mg-3Al-1Zn (AZ31) [29,36]. It is noted that the rolling
altered by the 423K heating. The recrystallization behavior of both
parameters and annealing conditions for AZ31 were different from
alloys followed similar trends. As shown in Fig. 3, new grains started
the Ce-containing alloys. Fig. 9 compares the EBSD IPF map in the
to appear at temperatures between 423 and 473 K, and as expected,
ND, the corresponding texture in the form of {0001} pole figures,
with the formation of recrystallized grains, the maximum texture
and the misorientation angle distribution in the final annealed
intensity in the microstructural patches gradually decreased. The
microstructure for ZE20, ZE21, and AZ31 [36]. The pole figures and
maximum texture intensity continued to decrease for annealing
the misorientation angle distribution were obtained from regions
temperatures up to 548 K (see Figs. 3 and 8). This is consistent with
containing approximately 2800, 1400, and 4800 grains in ZE20,
the literature, where a gradually weakening of the basal texture
ZE21, and AZ31, respectively. The ZE20 was annealed at tempera-
results during recrystallization during elevated-temperature
tures up to 598 K, whereas ZE21 was annealed at temperatures up
annealing. For example, the basal texture of partially recrystallized
to 623 K. Thus, assuming that recrystallization was complete by
AZ31 was slightly weaken after annealing at 250  C [30]. An
573 K in both alloys, ZE21 subsequently underwent a longer grain
as-rolled Mg-2.9Zn showed a gradually weakened basal texture
growth stage, such that a coarser grain size would be expected.
when annealed at 275  C [31] or 400  C [32]. An as-rolled Mg-0.3Zn-
However, the average grain sizes after the final annealing step in
0.1Ca and Mg-0.1Ca alloy showed a gradually weakening of the basal
rolled ZE20 and ZE21 were approximately 28 mm and 24 mm,
texture with an increasing volume fraction of recrystallized grains
respectively. The slightly smaller grain size of ZE21 was related to
[33]. Wu et al. [34] observed that the basal texture of as-rolled Mg-
the higher amount of RE content, which may have retarded the
1Gd weakened after annealing at 400  C for 30 min.
grain growth during annealing [7,15]. The average grain size in a
Overall, in both alloys, equiaxed and strain-free grains were
similarly rolled AZ31 microstructure after annealing was approxi-
observed after annealing at temperatures between 548 and 573K,
mately 16 mm, which may be due to the different processing history,
suggesting that recrystallization was complete by 573K (see Fig. 3).
annealing conditions, and composition of this alloy [12,36].
The maximum texture intensity increased for annealing tempera-
Nevertheless, a weaker texture was observed in ZE20 and ZE21
tures above 548 K (see Figs. 3 and 8). This was associated with the
after annealing, as compared to AZ31 (see Fig. 9). In addition, the
grain growth, which occurred after recrystallization, and therefore,
maximum texture intensity observed in ZE21 was smaller than that
fewer grains were present in the microstructural patches analyzed
of ZE20, suggesting that an increase in the Ce content assists
at higher temperatures. Similarly, Bhattacharyya et al. [35] found
A. Chakkedath et al. / International Journal of Lightweight Materials and Manufacture 1 (2018) 256e264 261

Fig. 4. The grain orientation spread maps of the ZE21 microstructural patch represented in Fig. 3 as a function of annealing temperature. The rolling direction is horizontal.

texture weakening in Mg alloys, which is consistent with previous solute concentration. Robson [43] explored the effect of solute
findings [13]. Solute drag, which restricts grain boundary motion, is addition on grain boundary segregation and mobility in Mg, and the
described by the interaction between solute atoms and high angle temperature and strain rate regimes in which dynamic recrystal-
grain boundaries during dynamic and static recrystallization [37]. lization would be suppressed as a result of solute drag were pre-
Alloying additions arrest the progression of recrystallization by dicted. Though the effect of solutes on boundary mobility is
imposing solute drag on grain boundaries or by precipitation reduced at higher temperatures [44], it is felt that the increased
(Zener) pinning [38e40]. The effect of solutes on boundary mobility concentration of Ce in the ZE21 alloy likely led to its finer grain size.
is well described, and solute segregation is known to have a sig- The misorientation angle distribution among the grain bound-
nificant effect on boundary mobility [41]. Boundary migration aries in AZ31 was concentrated around 30 , whereas, a uniform
decreases with increasing solute concentration [42], and the distribution of misorientation angles was observed in ZE20 and
boundary velocity is found to be inversely proportional to the ZE21 (see Fig. 9). Rotations about <1120>, <1010>, and <1011>
262 A. Chakkedath et al. / International Journal of Lightweight Materials and Manufacture 1 (2018) 256e264

Fig. 5. The fraction of misorientation values greater than 1.5 within the grains
analyzed during the annealing experiments as a function of annealing temperature for
both alloys. Fig. 8. Maximum texture intensity values for the microstructural patches analyzed
during the annealing experiments for both alloys.

were identified in ~76% and ~78% of all the newly-formed grain


boundaries after the annealing experiments in ZE20 and ZE21,
respectively (see Fig. 6). Fig. 10 shows the distributions of <1120>,
<1010>, and <1011> rotation axes after annealing for ZE20 and
ZE21, as compared to AZ31. A significantly different distribution of
<hki0> and <hkil> type rotation axis was observed for AZ31
compared to ZE20 and ZE21. Specifically, a significantly higher
proportion of <hkil> type rotation axis was observed in the
Ce-containing alloys compared to AZ31 (a number fraction of 0.16
and 0.18 for ZE20 and ZE21, respectively, compared to 0.06 for
AZ31). These differences might be due to the different type of
deformation systems activated during the rolling process, in which
the activity of the different type of slip systems could result in a
difference in the characteristics of the new grains formed in
the deformed regions. For example, RE additions are expected to
enhance the non-basal slip activity during deformation [1,16,45].
Fig. 6. Distribution of the number fraction of grain boundaries with specific rotation However, the distribution of the deformation modes was not
axis for the newly-formed grains in (a) ZE20 during the annealing experiments for investigated during rolling of the alloys in this study.
temperatures between 423 and 548 K and (b) ZE21 during the annealing experiments
for temperatures between 423 and 598 K.

Fig. 7. The distribution of the number fraction of the observed misorientation rotation axis for the newly-formed grain boundaries in ZE20 (blue) and ZE21 (red), sorted into <hki0>
(lightly shaded) and <hkil> (solid) types.
A. Chakkedath et al. / International Journal of Lightweight Materials and Manufacture 1 (2018) 256e264 263

Fig. 9. The EBSD IPF map in the ND, the corresponding texture in the form of {0001} pole figures, and the misorientation angle distribution in the final annealed microstructure in
rolled (a) ZE20, (b) ZE21, and (c) AZ31 [36]. The grain boundaries highlighted by white lines in the EBSD IPF correspond to <1120> and <1010> type, and the black ones correspond
to <1011> type rotation axis. Imax is the maximum texture intensity values observed. The EBSD data was post-processed using a clean-up procedure in which multiple iteration of
grain dilation with a grain tolerance angle of 5 and minimum grain size of 5 pixels was performed. Furthermore, the EBSD data were partitioned with a grain tolerance angle of 5
and a minimum grain size of 15e25 pixels.

5. Summary and conclusions employed to understand the effect of Ce additions (0.2e0.6 wt%) on
the recrystallization behavior of rolled Mg-2Zn alloys.
A novel in-situ experimental technique which involves anneal-
ing inside a SEM combined with EBSD analysis was developed and (a) A completely recrystallized microstructure with equiaxed and
strain-free grains were observed after the annealing step be-
tween 548 and 573K, suggesting that recrystallization was
complete by 573K in rolled ZE20 and ZE21.
(b) The average grain sizes after the final annealing step at 598K
and 623K in ZE20 and ZE21 were approximately 28 mm and
24 mm, respectively. This suggests that increased Ce additions
assists in the prevention of grain growth in Mg alloys.
(c) A weaker texture was observed in rolled ZE20 and ZE21 after
annealing compared to a conventional alloy, AZ31, that was
rolled and annealed. This suggests that Ce additions assist
in texture weakening in Mg alloys.

Acknowledgements

The funding for this research was supported by National Science


Foundation Division of Material Research (Grant No. DMR1107117)
Fig. 10. The distribution of the grain boundaries with <1120>, <1010>, and <1011>
through the Materials World Network program. Mr. Vahid Kha-
type rotation axis in ZE20 and ZE21 in the final annealed microstructure as compared demi, a PhD student at Michigan State University, is acknowledged
with rolled and annealed AZ31 [36]. for the assistance in the development of the in-situ heating stage
264 A. Chakkedath et al. / International Journal of Lightweight Materials and Manufacture 1 (2018) 256e264

setup used in this work. The technical assistance of Dr. Zhe Chen, a [21] P.J. Hurley, F.J. Humphreys, A study of recrystallization in single-phase
aluminium using in-situ annealing in the scanning electron microscope,
postdoctoral fellow at the University of California Santa Barbara, is
J. Microsc. 213 (3) (2004) 225e234.
also acknowledged. Drs Jan Bohlen, Sangbong Yi, and Dietmar [22] D. Mattissen, A. Wærø, D.A. Molodov, L.S. Shvindlerman, G. Gottstein, In-situ
Letzig of the Magnesium Innovation Centre MagIC, Helmholtz- investigation of grain boundary and triple junction kinetics in aluminiume10
Zentrum Geesthacht, Geesthacht, Germany are acknowledged for ppm magnesium, J. Microsc. 213 (3) (2004) 257e261.
[23] A. Lens, C. Maurice, J.H. Driver, Grain boundary mobilities during recrystalli-
providing the processed alloys used in this work. zation of AleMn alloys as measured by in situ annealing experiments, Mater.
Sci. Eng. A 403 (1) (2005) 144e153.
[24] Anne-Laure Helbert, W. Wang, F. Brisset, T. Baudin, R. Penelle, In situ EBSD
investigation of recrystallization in a partially annealed and cold-rolled
References aluminum alloy of commercial purity, Adv. Eng. Mater. 14 (1-2) (2012) 39e44.
[25] K. Mirpuri, H. Wendrock, S. Menzel, K. Wetzig, J. Szpunar, High temperature
behavior of Cu films studied in-situ by Electron Backscatter Diffraction,
[1] Y. Chino, K. Motohisa, M. Mamoru, Enhancement of tensile ductility and Microelectron. Eng. 76 (1) (2004) 160e166.
stretch formability of magnesium by addition of 0.2 wt%(0.035 at%) Ce, Mater. [26] N. Bozzolo, S. Jacomet, R.E. Loge , Fast in-situ annealing stage coupled with
Sci. Eng. 494 (1) (2008) 343e349. EBSD: a suitable tool to observe quick recrystallization mechanisms, Mater.
[2] A.A. Luo, R.K. Mishra, A.K. Sachdev, High-ductility magnesiumezincecerium Char. 70 (2012) 28e32.
extrusion alloys, Scripta Mater. 64 (5) (2011) 410e413.
[27] G.G.E. Seward, S. Celotto, D.J. Prior, J. Wheeler, R.C. Pond, In situ SEM-EBSD
[3] S. Yi, J. Bohlen, F. Heinemann, D. Letzig, Mechanical anisotropy and deep observations of the hcp to bcc phase transformation in commercially pure
drawing behaviour of AZ31 and ZE10 magnesium alloy sheets, Acta Mater. 58
titanium, Acta Mater. 52 (4) (2004) 821e832.
(2) (2010) 592e605. [28] H. Nakamichi, F.J. Humphreys, I. Brough, Recrystallization phenomena in an IF
[4] S. Mohapatra, J. Jain, Overview of static recrystallization in magnesium alloys, steel observed by in situ EBSD experiments, J. Microsc. 230 (3) (2008)
Adv. Mater. Process. 173 (4) (2015) 28e31.
464e471.
[5] L.W.F. Mackenzie, M.O. Pekguleryuz, The recrystallization and texture of [29] A. Chakkedath, C.J. Boehlert, D. Hernandez, J. Bohlen, S. Yi, D. Letzig,
magnesiumezincecerium alloys, Scripta Mater. 59 (6) (2008) 665e668.
C.J. Boehlert, In-situ EBSD technique characterizes microstructure evolution of
[6] J. Bohlen, M.R. Nürnberg, J.W. Senn, D. Letzig, S.R. Agnew, The texture and magnesium alloy, June, Adv. Mater. Process. 174 (6) (2016) 19e21.
anisotropy of magnesiumezincerare earth alloy sheets, Acta Mater. 55 (6) [30] X. Huang, K. Suzuki, Y. Chino, Different annealing behaviours of warm rolled
(2007) 2101e2112. Mg-3Al-1Zn alloy sheets with dynamic recrystallized microstructure and
[7] K. Hantzsche, J. Bohlen, J. Wendt, K.U. Kainer, S.B. Yi, D. Letzig, Effect of rare deformation microstructure, Mat. Sci. Eng. A Struct. 560 (2013) 232e240.
earth additions on microstructure and texture development of magnesium  Martin, M. Sanjari, E. Essadiqi, S. Yue, Texture weakening and
[31] S.A. Farzadfar, E.
alloy sheets, Scripta Mater. 63 (7) (2010) 725e730. static recrystallization in rolled Mg-2.9Y and Mg-2.9Zn solid solution alloys,
[8] T. Al-Samman, X. Li, Sheet texture modification in magnesium-based alloys by J. Mater. Sci. 47 (2012) 5488e5500.
selective rare earth alloying, Mater. Sci. Eng. A 528 (10) (2011) 3809e3822. [32] S.A. Farzadfar, E. Martin, M. Sanjari, E. Essadiqi, M.A. Wells, S. Yue, On the
[9] A. Chakkedath, J. Bohlen, S. Yi, D. Letzig, Z. Chen, C.J. Boehlert, The effect of Nd deformation, recrystallization and texture of hot-rolled Mg-2.9Y and Mg-
on the tension and compression deformation behavior of extruded Mg-1Mn
2.9Zn solid solution alloys-A comparative study, Mater. Sci. Eng. A 534
(wt pct) at temperatures between 298 K and 523 K (25 C and 250 C), (2012) 209e219.
Metall. Mater. Trans. A 45 (8) (2014) 3254e3274.
[33] Z.R. Zeng, Y.M. Zhu, S.W. Xu, M.Z. Bian, C.H.J. Davies, N. Birbilis, J.F. Nie,
[10] N. Stanford, Micro-alloying Mg with Y, Ce, Gd and La for texture mod- Texture evolution during static recrystallization of cold-rolled magnesium
ificationda comparative study, Mater. Sci. Eng. A 527 (2010) 2669e2677. alloys, Acta Mater. 105 (2016) 479e494.
[11] M.T. Perez-Prado, O.A. Ruano, Texture evolution during annealing of magne- [34] W.X. Wu, L. Jin, Z.Y. Zhang, W.J. Ding, J. Dong, Grain growth and texture
sium AZ31 alloy, Scripta Mater. 46 (2) (2002) 149e155. evolution during annealing in an indirectextruded Mg-1Gd alloy, J. Alloy.
[12] C.J. Boehlert, Z. Chen, I. Gutie rrez-Urrutia, J. Llorca, M.T. Pe
rez-Prado, In situ
Comp. 585 (2014) 111e119.
analysis of the tensile and tensile-creep deformation mechanisms in rolled [35] J.J. Bhattacharyya, S.R. Agnew, G. Muralidharan, Texture enhancement during
AZ31, Acta Mater. 60 (4) (2012) 1889e1904. grain growth of magnesium alloy AZ31B, Acta Mater. 86 (2015) 80e94.
[13] J. Bohlen, S. Yi, D. Letzig, K.U. Kainer, Effect of rare earth elements on the [36] Z. Chen, A Study of the Effects of Processing and Alloying on the Micro-
microstructure and texture development in magnesiumemanganese alloys structure and Deformation Behavior of Wrought Magnesium Alloys, PhD
during extrusion, Mater. Sci. Eng. A 527 (26) (2010) 7092e7098.
Thesis, Michigan State University, 2012.
[14] S.R. Agnew, M.H. Yoo, C.N. Tome, Application of texture simulation to [37] Elena Pereloma, David V. Edmonds (Eds.), Phase Transformations in
understanding mechanical behavior of Mg and solid solution alloys containing
Steels: Fundamentals and Diffusion-controlled Transformations, vol. 1,
Li or Y, Acta Mater. 49 (20) (2001) 4277e4289. Elsevier, 2012.
[15] J.P. Hadorn, K. Hantzsche, S. Yi, J. Bohlen, D. Letzig, J.A. Wollmershauser, [38] A. Rollett, F.J. Humphreys, G.S. Rohrer, M. Hatherly, Recrystallization and
S.R. Agnew, Role of solute in the texture modification during hot deformation Related Annealing Phenomena, Elsevier, 2004.
of Mg-rare earth alloys, Metall. Mater. Trans. A 43 (4) (2012) 1347e1362. [39] Yuqing Weng (Ed.), Ultra-fine Grained Steels, Springer Science & Business
[16] S. Sandlo €bes, S. Zaefferer, I. Schestakow, S. Yi, R. Gonzalez-Martinez, On the
Media, 2009.
role of non-basal deformation mechanisms for the ductility of Mg and MgeY [40] J.P. Hadorn, K. Hantzsche, Y. Sangbong, J. Bohlen, D. Letzig, S. Agnew, Effects of
alloys, Acta Mater. 59 (2) (2011) 429e439. solute and second-phase particles on the Texture of Nd-containing Mg alloys,
[17] N. Stanford, The effect of rare earth elements on the behaviour of magnesium- Metall. Mater. Trans. A 43 (4) (2012) 1363e1375.
based alloys: Part 2erecrystallisation and texture development, Mater. Sci. [41] C. Dimitrov, O. Dimitrov, F. Dworschak, The interaction of self interstitials
Eng. A 565 (2013) 469e475.
with undersized solute atoms in electron-irradiated aluminium, J. Phys. F Met.
[18] J. Bohlen, S. Yi, J. Victoria-Hernandez, G. Kurz, D. Letzig, K.U. Kainer, The in- Phys. 8 (6) (1978) 1031.
fluence of the combination of alloying elements on the microstructure and ments d’addition sur la
[42] C. Frois, M.O. Dimitrov, Influence de quelques ele
texture development during rolling and annealing of magnesium, in: The 10th recristallisation de l’aluminium tre s pur, Annales de Chimie 1 (1966).
International Conference on Magnesium Alloys and Their Applications, Jeju, [43] Joseph D. Robson, Effect of rare-earth additions on the texture of wrought
Korea, 2015, pp. 625e636. magnesium alloys: the role of grain boundary segregation, Metall. Mater.
[19] Stuart I. Wright, Matthew M. Nowell, A Review of in Situ EBSD studies. Trans. A 45 (8) (2014) 3205e3212.
Electron Backscatter Diffraction in Materials Science, Springer US, 2009,
[44] Paul Gordon, R.A. Vandermeer, Mechanism of boundary migration in recrys-
pp. 329e337. tallization, Trans. Metall. Soc. AIME 224 (5) (1962) 917.
[20] G.G.E. Seward, D.J. Prior, J. Wheeler, S. Celotto, D.J.M. Halliday, R.S. Paden, [45] Nicole Stanford, Dale Atwell, Matthew R. Barnett, The effect of Gd on the
M.R. Tye, High-temperature electron backscatter diffraction and scanning recrystallisation, texture and deformation behaviour of magnesium-based
electron microscopy imaging techniques: in-situ investigations of dynamic alloys, Acta Mater. 58 (20) (2010) 6773e6783.
processes, Scanning 24 (5) (2002) 232e240.

You might also like