Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Journal of Alloys and Compounds 686 (2016) 1008e1016

Contents lists available at ScienceDirect

Journal of Alloys and Compounds


journal homepage: http://www.elsevier.com/locate/jalcom

Evolution of the cold-rolling and recrystallization textures in


FeNiCoAlNbB shape memory alloy
Huadong Fu a, *, Huimin Zhao a, Shilei Song a, Zhihao Zhang a, Jianxin Xie a, b, **
a
Institute for Advanced Materials and Technology, University of Science and Technology Beijing, Beijing, 100083, Peoples Republic of China
b
Beijing Laboratory of Metallic Materials and Processing for Modern Transportation, Beijing, 100083, Peoples Republic of China

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

Article history: The evolution of cold-rolling and recrystallization textures in the newly-developed ferrous shape
Received 21 May 2016 memory alloy (FeNiCoAlNbB) was investigated and the improving mechanism on superelasticy of the
Received in revised form severe cold-rolled alloy was discussed. A weaker copper rolling texture ({112}111) was formed in
25 June 2016
FeNiCoAlNbB alloy at relatively low rolling reductions (80%); the copper rolling texture transformed to
Accepted 27 June 2016
Available online 29 June 2016
the Goss {110}001 and brass {110}112 orientations through twinning and dislocation slipping with
the rolling reductions of 90%e95%; signicantly enhanced rolling texture of a strong brass orientation
was obtained with the rolling reduction of 98.5%. The 98.5% cold-rolled FeNiCoAlNbB alloy after solution
Keywords:
Rolling
treatment at 1220  C for 1 h with strong {hk0}001 recrystallization texture, followed by aging for 96 h
Recrystallization at 600  C exhibited good superelasticity of 3.2% with residual strain only 0.7%, and the tensile strength
Texture was approximately 960 MPa. Compared with the non-superelasticity in the as-forged FeNiCoAlNbB alloy,
Shape memory the considerably improving superelasticity in this alloy mainly attributes to the formation of strong
Superelasticity favorable textures and the suppression of grain boundary precipitation.
2016 Elsevier B.V. All rights reserved.

1. Introduction Centering on this found, wide attentions and much research have
recently been aroused and launched [11e15]. Afterwards, Kai-
Shape memory alloys (SMAs), as one kind of intelligent func- numas group developed a new shape memory alloy FeNi-
tional materials with both shape memory effect (SME) and super- CoAlNbB alloy with more cost-efcient and low-melting element
elasticity (SE) [1e3], can be divided into three groups: NiTi-based Nb instead of Ta, and this alloy also showed superelasticity over
[4], Cu-based [5] and Fe-based [6]. NiTi-based alloys are the most 5% [16].
well-known SMAs owing to their excellent shape memory prop- Relevant studies on FeNiCo-based SMAs fabricated by
erties and corrosion resistance [7,8]. However, their low cold- rolling recrystallizing have conrmed that an excellent supere-
workability and high processing cost have limited the further lastic performance can be obtained only with cold-rolling reduction
large-scale application. In contrast, with high strength, excellent above 98% followed by advisable solution and aging treatments.
cold workability and low cost etc., the newly-developed Fe-based The internal machnism of this critical reduction is not clear at
shape memory alloys enjoy a broad application prospect in industry present. Therefore, it is indicated that rolling reduction signicantly
[9]. inuences the microstructure and superelasticity of the alloy. To
Particularly the thermoelastic martensitic transformation in reveal the improving mechanism on the superelastic properties of
FeNiCoAlTaB alloy discovered by Kainuma et al. [10], the alloy FeNiCo-based alloys with severe rolling deformation, clarifying the
showed huge superelastic strain, which was almost twice the rolling and recrystallization texture evolution in these alloys is
maximum superelastic strain obtained in the NiTi-based alloy. fundamentally important.
Thus, in the present paper, the effects of rolling deformation on
microstructure and texture after cold-rolling and annealing in
* Corresponding author.
FeNiCoAlNbB alloy were investigated.
** Corresponding author. Institute for Advanced Materials and Technology, Uni- The reasons of excellent superelasticity at severe rolling defor-
versity of Science and Technology Beijing, Beijing, 100083, Peoples Republic of mation were also discussed, which could offer guidance for efcient
China. manufacture of the alloy.
E-mail addresses: hdfu@ustb.edu.cn (H. Fu), jxxie@mater.ustb.edu.cn (J. Xie).

http://dx.doi.org/10.1016/j.jallcom.2016.06.273
0925-8388/ 2016 Elsevier B.V. All rights reserved.
H. Fu et al. / Journal of Alloys and Compounds 686 (2016) 1008e1016 1009

Table 1 Table 2
Chemical composition of the FeNiCoAlNbB alloy (wt%). Euler angles and Miller indices for common texture components in FCC metals.

Fe Ni Co Al Nb B Texture components Symbols Miller indices Euler angles ( )

41.93 30.45 18.29 5.11 4.21 0.01 f1 F f2

Cube (C) - {001}100 45 0 45


Rotated Cube (Rt-C) {001}110 0/90 0 45
2. Experimental procedures Goss (G) A {110}001 90 90 45
Goss/Brass (G/B) {110}115 74 90 45
Brass (B) : {110}112 55 90 45
Pure iron (99.9 wt%), pure nickel (99.9 wt%), pure cobalt (99.9 wt A < {110}111 35 90 45
%), pure aluminum (99.9 wt%), pure niobium (99.9 wt%) and fer- Rotated Goss (Rt-G) ; {011}011 0 90 45
roboron (17.5 wt% boron content, iron and boron content 99.95 wt Rotated Cu (Rt-Cu) B {112}011 0 35 45
S C {123}634 59 37 63
%) were used as raw materials. Referring to the research results
Copper (Cu) {112}111 90 35 45
given by Omori et al. [16], the Fe-28Ni-17Co-11.5Al-2.5Nb-0.05B
(at.%) alloy ingot was rst induction melted and casted in a vac-
uum furnace. Its chemical composition is shown in Table 1.
The as-cast ingot was then homogenized and forged in the eld emission gun electron source (FEG) with the voltage of 20 kV.
temperature range of 1100e1200  C to both reduce the casting Superelasticity experiments were performed in cyclic loading/
defects and improve the homogeneity of microstructure. The unloading tension at room temperature (~20  C) with increasing
20 mm-thick as-forged specimens were homogenized at 1200  C applied strain until the specimen fractured. Each cycle of defor-
for 30 min, followed by water quenching. Then, the alloys were mation value is 1%. The dog-bone-shaped tension specimens with
cold-rolled with the reductions of 60%, 80%, 90%, 95% and 98.5%, gauge dimensions of 0.3 mm  4 mm  20 mm were examined
without intermediate annealing. The rolling reduction can be along rolling direction on an MTS810 universal material testing
described as (h0h1)/h0  100%, where h0 and h1 are the machine with the strain rate of 1  103 s1.
thicknesses of the sheets before and after cold rolling, respectively.
Subsequently, the cold-rolled sheets were solution-treated at 3. Results
1220  C for 1 h to obtain recrystallized microstructure, followed by
water quenching to retain g-single phase structure, and then aged 3.1. Microstructure and texture evolution in cold-rolled alloy
at 600  C for 96 h.
The rolled and recrystallized specimens in various reductions Fig. 2 shows the microstructure of the as-forged and cold-rolled
were cut out for texture measurement with an area of 24 mm alloy with various rolling reductions of 60%,80%,90%,95% and 98.5%.
(RD)  14 mm by D5000 X-ray diffraction device. The {111}, {200}, It can be seen that the microstructure of the as-forged specimen
{113}, and {220} incomplete pole gure data were used to calculate was dominated by uniform equiaxed grains with an average size of
the orientation distribution functions (ODFs) using the Bunge ~100 mm. After over 60% of the rolling deformation, grains were
method. Fig. 1 shows locations of common texture components of long rolled and increasing rolling deformation intensied the grain
FCC metals in f2 0 , 45 , and 65 sections of the Euler space. deformation and renement. Severely fragmented and much ner
Table 2 gives the Euler angles and the Miller indices for common microstructure developed in the 98.5% cold-rolled alloy sheets with
texture components in FCC metals. The evolution of texture during the blurred grain boundaries.
deformation and recrystallization can be understood from the Fig. 3 shows respective ODFs (f2 0 , 45 , and 65 sections) for
f2 0 , 45 and 65 sections of the respective ODFs and the cold-rolled alloy sheets by different cold-rolling reductions. The
important orientations appearing in the different ODF sections are intensity of the texture in alloy before rolling (Fig. 3(a)) is about 1
summarized for ease of understanding. and the textures are approximated by the random orientation
The microstructure and texture of the cold-rolled and recrys- distribution, or nearly random texture. Presence of Cu component
tallized specimens were characterized using an electron back- is noticed in the f2 45 section of the 60%e80% cold-rolled
scatter diffraction (EBSD) system (Oxford Instruments, UK) specimens with almost unchanged orientation density (Fig. 3(b)
attached to a scanning electron microscope (SEM) equipped with and (c)). It is noted that the 90% cold-rolled texture is dominated by
newly-developed nearly G/B orientation instead of gradually absent

Fig. 1. Locations of common texture components of FCC metals in f2 0, 45, and 65 sections.
1010 H. Fu et al. / Journal of Alloys and Compounds 686 (2016) 1008e1016

Fig. 2. Microstructures of FeNiCoAlNbB alloy with different cold-rolling reductions (a) 0% (as-forged); (b) 60%; (c) 80%; (d) 90%; (e) 95%; (f) 98.5%.

Cu orientation, f(g)max 11.4, 13.3 in Fig. 3(d) and (e). With further recrystallization texture evolution in subsequent annealed alloy
increasing rolling reduction deformation up to 98.5%, a strong with initial different cold-rolling reductions.
rolling brass texture ({110}112) developed and apparently It can be seen that no obvious typical recrystallization texture in
strengthened (f(g)max 26.1) in Fig. 3(f). 80% cold-rolled alloy after solution-treated. A relatively weak
The cold-rolling textures of FCC metals are conveniently repre- recrystallized Goss texture ({110}001) presents in the alloy with
sented by orientation densities f(g) along the a (110//ND) and t rolling reduction of 90%. A slow increase of the orientation density
bers as shown in Fig. 4. It can be seen that the cold-rolling textures displays when rolling reduction up to 95%. Furthermore, the f(g)
are characterized by strong aber (G through B). Copper ({112} max of recrystallization texture reached 37.2 with the rolling
111), brass ({110}112) and Goss ({110}001) textures are three reduction of 98.5%, which is much higher than 11.3 obtained in the
major components in FeNiCoAlNbB cold-rolling texture (Fig. 4). At alloy with 95% rolling deformation. This peak intensity value ob-
lower rolling reduction (80%), the textures consist of major cop- tained at (5, 35, 0) and (5, 55, 0) in Euler space, and corresponding
per texture orientation, minor brass ({110}112) and Goss orien- Miller indices are (023)[100] and (032)[100] orientations. Mean-
tations ({110}001). At higher rolling reduction (90%), the while, an increasing intensity of Goss texture component is also
proportion of brass and Goss texture components signicantly in- obtained with the increase of deformation.
creases. Moreover, with increasing the cold-rolling reduction, the Therefore, cold-rolling deformation less than 90% and recrys-
density of Goss texture ascends rstly and descends at last, pre- tallization at 1220  C for 1 h does not give rise to a signicant
senting the trend of the transition from the Goss to the brass favorable recrystallization texture and a strong preferred 100
texture. recrystallized orientation only exists in the alloy with severe rolling
deformation about 98.5% and recrystallization at 1220  C for 1 h.
3.2. Recrystallization texture evolution in annealed alloy Fig. 6 shows the orientation densities f(g) along the a (110//
ND) and t bers for the annealed FeNiCoAlNbB alloy. It is noted that
Referring to the results of Omori et al. [16], the cold-rolled the intensity of dominant rolling brass texture ({110}112)
FeNiCoAlNbB sheets need to be solution-treated at 1220  C for decrease drastically, while the recrystallized (023)[100], (032)[100]
1 h to ensure nal shape memory properties. Fig. 5 shows the orientations and Goss texture ({110}001) rise apparently. This
H. Fu et al. / Journal of Alloys and Compounds 686 (2016) 1008e1016 1011

Fig. 3. ODFs (f2 0 , 45 , and 65 sections) for rolling textures of FeNiCoAlNbB alloy with different cold-rolling reductions (a) 0% (as-forged), (b) 60%, (c) 80%, (d) 90%, (e) 95%, (f)
98.5%.
1012 H. Fu et al. / Journal of Alloys and Compounds 686 (2016) 1008e1016

Fig. 4. Orientation densities f(g) along (a) a ber (110//ND) and (b) t ber for FeNiCoAlNbB alloy with different cold-rolling reductions.

implies that the brass rolling texture changes into the strong 100 main deformation mechanisms. As for rolling reduction 80%, alloy
ber texture after recrystallization. deformation is characterized by normal dislocation slipping at
active {111}110 systems, consisting of (-11-1)[110],(1-1-1)
3.3. Microstructure and superelasticity of the alloy fabricated by [110],(111)[10-1] and (111)[01-1] calculated by RC (Relaxed Con-
rolling recrystallization straints) model [17]. The contributions of these slip systems cause
{112} plane parallels to rolling plane, 111 direction parallel to
In order to further clarify the microstructural and textural rolling plane, thus forming the relatively stable copper texture
discrepancy before and after the cold rolling, EBSD scans and ({112}111).
analysis were acquired from the as-forged alloy and the 98.5% cold- With the increase of rolling reduction, dislocation slips are
rolled and recrystallized alloy. Quasi-colored orientation maps of impeded by the piling up of dislocations. Then the alloy with low or
microstructure, inverse pole gures and grain boundary misori- medium SFE, such as FeNiCoAlNbB alloy tends to twinning defor-
entation of the alloys are shown in Fig. 7 and Fig. 8. mation. Copper-oriented ({112}111) grains undergo mechanical
In the as-forged specimens, the microstructure of equiaxed twinning easily rotate to corresponding twinning oriented position
grains with an average grain size of ~200 mm is observed with high {552} 115. The twinning deformation changes the grain orienta-
fraction of high-angle boundaries and a large amount of annealing tion signicantly, so that the dislocation slips proceed again. The
twinning grains, as shown in Fig. 7. The as-forged alloy mainly metastable {552} 115 orientation then ows to Goss orientation
exhibits random distribution of grain orientation and grain ({110}001), which at last turn to stable brass texture ({110}112)
boundary misorientation, except the twinning boundaries. as shown in Fig. 4(b). In the Goss to brass texture transition, a pair of
The alloy fabricated by 98.5% cold-rolled and then annealed at shear stresses with contrary symbols can be resolved, contributing
1220  C for 1 h presents a strong 100 ber texture, with an to overcome the probable difculties of grain orientation aggre-
average grain size of 200e500 mm. Low-angle boundary in this gation on the brass orientation. Meanwhile, due to the special
alloy becomes the main grain boundary type with a proportion of symmetrical positions of the brass texture, brass oriented crystals
~53%. twinning orientation may in another symmetrical position of brass
The as-forged specimen and the 98.5% cold-rolled specimen are orientation, or rotate to other unstable orientation position and
both solution-treated at 1220  C for 1 h, followed by water eventually return to the brass orientation.
quenched to maintain austenite phase, and nally aged at 600  C From the above, with the increase of rolling deformation, rstly
for 96 h to create the g0 phase precipitate [10]. The superelasticity active slip systems cause the aggregation of the relatively stable
was determined by cyclic tensile test at room temperature with copper texture ({112}111); continuing deformation leads to me-
increment strain of 1%. The tensile stress-strain curves as shown in chanical twinning from copper orientation ({112}111) towards
Fig. 9. (t is total strain, r is residual strain, se is superelastic strain, {552} 115; further deformation results in that {552} 115 orien-
e is elastic strain). tation nally ows to stable brass orientation ({110}112) through
By comparison between the two curves in Fig. 9, the as-forged Goss orientation ({110}001), the process schematic diagram is
specimen shows a high yield stress and total strain of less than shown in Fig. 10.
1% without superelasticity. However, in the 98.5% cold-rolled The formation of recrystallization texture is related to the
specimen, the superelastic response is clearly observed with the spontaneous change of microstructure, and the trend direction of
total strain of 5%, residual strain of 0.7%, recoverable strain of 3.2% the change is lowering the total system energy. On the basis of the
and tensile strength of ~960 MPa. strain-energy release-maximization theory (SERM) developed by
Lee et al. [18], the strain energy release in recrystallized grains can
4. Discussion be maximized when the absolute maximum internal-stress direc-
tion (AMSD) is parallel to the minimum Youngs modulus direction
4.1. Analysis on cold-rolling and recrystallization texture evolution (MYMD). The MYMD of FeNiCoAlNbB alloy with FCC structure is
001 orientation. Therefore, the recrystallized grains collect in
Polycrystalline deformation gives rise to the considerable grain {hk0} 001 recrystallized textures. Moreover, the study by Engler
rotations toward particular orientations inside alloys, thus various et al. [19] indicated that the brass orientation ({110}112) in cold-
rolling texture types are closely related to main mechanisms of rolled CueMn alloys transformed to the Goss orientation ({110}
rolling deformation. 001) after recrystallization. Hjelen et al. [20] reported that the
In the FeNiCoAlNbB alloy with relatively low stacking fault en- Goss oriented grains in aluminum appeared to be nucleated and
ergy (SFE), dislocation slip and mechanical twinning are the two grow from the deformation brass texture, further conrming the
H. Fu et al. / Journal of Alloys and Compounds 686 (2016) 1008e1016 1013

Fig. 5. ODFs (f2 0 , 45 and 65 sections) for recrystallization textures of FeNiCoAlNbB alloy with different cold-rolling reductions (a) 0% (as-forged), (b) 60%, (c) 80%, (d) 90%, (e)
95%, (f) 98.5%.
1014 H. Fu et al. / Journal of Alloys and Compounds 686 (2016) 1008e1016

Fig. 6. Orientation densities f(g) along (a) a ber (110//ND) and (b) t ber for annealed FeNiCoAlNbB alloy with different cold-rolling reductions.

Fig. 7. Quasi-colored orientation maps of microstructure, inverse pole gures and grain boundary misorientation of the as-forged alloy after solution treatment (RD: Rolling di-
rection, TD: Transverse direction, ND: Normal direction).

fact that the Goss texture was closely related to the brass defor- expected in 110 and 111 directions are about 4.1% and 2.1%,
mation texture. The experimental results in the present study respectively, and the strain in the 100 direction has a maximum
coincide with the theoretical analysis and the experimental results value of ~8.7%.
of Lee and Englers studies. The nal recrystallization texture of the In this paper, the severe cold-rolled recrystallized FeNi-
FeNiCoAlNbB alloy is {hk0}001 texture (including {230}001 CoAlNbB alloy possesses dominant strong {hk0}001textures
texture and Goss texture). ((023)[100] and (032)[100] texture and Goss texture), and these
textures play a key role in the improvement of the superelasticity in
4.2. Increasing mechanism on superelasticity of 98.5% cold the alloy.
rolled recrystallized alloy The x-ray diffraction patterns of the as-forged and cold-rolled
FeNiCoAlNbB alloy after solution-aging treatment are shown in
From the tensile stress-strain curves in Fig. 9, the non- Fig. 11. It is shown that the as-forged and cold-rolled alloy after
superelastic as-forged alloy undergoing 98.5% cold-rolling and solution-aging treatment have the same phase composition of g
annealing, exhibits 5.0% total strain with 3.2% recovery strain. The phase (austenitic phase), a0 phase (BCT or BCC structure), g0 phase
main reasons for this huge improvement of superelastic perfor- ((Ni,Fe,Co)3(Al,Nb) phase with a L12 structure) and b phase (NiAl-
mance are associated with the formation of the strong favorable ordered bcc phase with a B2 structure). In addition, as-forged and
textures and the suppression of precipitation at low-energy grain cold-rolled alloys after solution-aging treatment differ greatly in
boundaries. grain boundary structures, as shown in Fig. 12. Considering of the
The superelasticity of the polycrystalline shape memory alloy is microstructure, XRD patterns and of the alloys in this study and the
strongly dependent on grain orientation owing to martensitic previous study [6], it is indicated that b-NiAl-ordered bcc phase
variant selection. On the basis of phenomenological theoretical precipitates along the grain boundaries in the matrix phase, which
calculation displayed by Kainuma et al. [10], the superelastic strains are marked by arrows in Fig. 12. In the cold-rolled alloy, the fraction
H. Fu et al. / Journal of Alloys and Compounds 686 (2016) 1008e1016 1015

Fig. 8. Quasi-colored orientation maps of microstructure, inverse pole gures and grain boundary misorientation of the 98.5% cold-rolled recrystallized alloy.

of low-angle boundaries (about 53%) with low-energy could be up to 98.5%, the alloy is characterized by the highly enhanced
lifted by strong recrystallization textures, which can improve the dominant brass texture, leading to the formation of strong
mechanical properties by suppression of the grain boundary pre- (023) [100], (032) [100] and Goss textures after
cipitation of the brittle b phase. The as-forged alloy with a random recrystallization.
orientation distribution, where most of the grain boundaries are (2) The 98.5% cold-rolled FeNiCoAlNbB alloy after solution
composed of high-angle boundaries covered by the brittle b phase, treatment at 1220  C for 1 h and then aging treatment at
often fractures before showing superelasticity.

5. Conclusions

(1) At relatively lower cold rolling reduction (80%), a weak


copper texture appeared in the FeNiCoAlNbB alloy. Stronger
brass and Goss textures were formed in the alloy with cold-
rolling reduction of 90e95%. With the cold-rolling reduction

Fig. 9. Tensile stress-strain curves of the as-forged alloy and the 98.5% cold-rolled
specimen after the same solution-aging treatment. Fig. 10. Orientation rotation process in cold-rolled FeNiCoAlNbB alloy.
1016 H. Fu et al. / Journal of Alloys and Compounds 686 (2016) 1008e1016

Fig. 11. X-ray diffraction patterns of the (a) as-forged and (b) cold-rolled FeNiCoAlNbB alloy after solution-aging treatment.

Fig. 12. Grain boundary structures of the (a) as-forged and (b) cold-rolled FeNiCoAlNbB alloys after solution-aging treatment (b phase marked by arrows).

600  C for 96 h exhibits good superelasticity with the [7] Song Y, Chen X, Dabade V, et al. Enhanced reversibility and unusual micro-
structure of a phase-transforming material. Nature, 502: 85e88.
superelasticity of 3.2%, residual strain of 0.7%, and tensile
[8] Chluba C, Ge W, DeMiranda R L, et al. Ultralow-fatigue shape memory alloy
strength of ~960 MPa. Compared with the non-superelastic lm. Science, 348: 1004e1007.
as-forged alloy, this improvement of superelastic perfor- [9] J.M. Jani, M. Leary, A. Subic, et al., A review of shape memory alloy research,
mance is chiey associated with the formation of strong application and opportunities, Mater. Des. 56 (2014) 1078e1113.
[10] Y. Tanaka, Y. Himuro, R. Kainuma, et al., Ferrous polycrystalline shape-
favorable textures and the suppression of grain boundary memory alloy showing huge superelasticity, Science 327 (2010) 1488e1490.
precipitation. [11] Y. Geng, M. Jin, W. Ren, et al., Effects of aging treatment on martensitic
transformation of Fe-Ni-Co-Al-Ta-B alloys, J. Alloys Compd. 577 (2013)
S631eS635.
Acknowledgements [12] F. Borza, N. Lupu, V. Dobrea, et al., Tailoring the magnetic properties of new
Fe-Ni-Co-Al-(Ta,Nb)-B superelastic rapidly quenched microwires, J. Appl.
This work was supported by the Major States Basic Research Phys. 117 (2015) 17E512.
[13] J. Ma, B. Kockar, A. Evirgen, et al., Shape memory behavior and ten-
Development Program (973 Program) of China under contract sionecompression asymmetry of a FeNiCoAlTa single-crystalline shape
number 2011CB606300, National Natural Science Foundation of memory alloy, Acta Mater. 60 (2012) 2186e2195.
China (No. 51504023), Fundamental Research Funds for the Central [14] P. Kroo, M.J. Holzweissig, T. Niendorf, et al., Thermal cycling behavior of an
aged FeNiCoAlTa single-crystal shape memory alloy, Scr. Mater. 81 (2014)
Universities (No. FRF-TP-15-051A2) and State Key Lab of Advanced
28e31.
Metals and Materials Foundation (2014-Z06). [15] J. Ma, B.C. Hornbuckle, I. Karaman, et al., The effect of nanoprecipitates on the
superelastic properties of FeNiCoAlTa shape memory alloy single crystals,
Acta Mater. 61 (2013) 3445e3455.
References
[16] T. Omori, S. Abe, Y. Tanaka, et al., Thermoelastic martensitic transformation
and superelasticity in Fe-Ni-Co-Al-Nb-B polycrystalline alloy, Scr. Mater. 69
[1] K. Otsuka, C.M. Wayman (Eds.), Shape Memory Materials, Cambridge Uni- (2013) 812e815.
versity Press, 1998. [17] J. Hirsch, K. Lucke, Mechanism of deformation and development of rolling
[2] D. Stoeckel, Shape memory actuators for automotive applications, Mater. Des. textures in polycrystalline F.C.C. metals-II. Simulation and interpretation of
11 (1990) 302e307. experiments on the basis of taylor-type theories, Acta Metall. 36 (1988)
[3] D.C. Dunand, P. Mullner, Size effects on magnetic actuation in Ni-Mn-Ga shape 2833e2904.
memory alloys, Adv. Mater. 23 (2011) 216e232. [18] D.N. Lee, Strain energy release maximization model for evolution of recrys-
[4] R. Mirzaeifar, R. DesRoches, A. Yavari, et al., A micromechanical analysis of the tallization textures, Int. J. Mech. Sci. 42 (2000) 1645e1678.
coupled thermomechanical superelastic response of textured and untextured [19] O. Engler, Recrystallization textures in copper-manganese alloys, Acta Mater.
polycrystalline NiTi shape memory alloys, Acta Mater. 61 (2013) 4542e4558. 49 (2001) 1237e1247.
[5] S.M. Ueland, Y. Chen, C.A. Schuh, Oligocrystalline shape memory alloys, Adv. [20] J. Hjelen, R. rsund, E. Nes, On the origin of recrystallization textures in
Funct. Mater. 22 (2012) 2094e2099. aluminum, Acta Metall. Mater. 39 (1991) 1377e1404.
[6] T. Omori, K. Ando, M. Okano, et al., Superelastic effect in polycrystalline
ferrous alloys, Science 333 (2011) 68e71.

You might also like