You are on page 1of 11

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/319192320

In situ observation on temperature dependence of martensitic transformation


and plastic deformation in superelastic NiTi shape memory alloy

Article  in  Materials & Design · August 2017


DOI: 10.1016/j.matdes.2017.08.037

CITATIONS READS

37 323

4 authors, including:

Yao Xiao Pan hai Zeng


Ruhr-Universität Bochum Wuhan University
19 PUBLICATIONS   179 CITATIONS    48 PUBLICATIONS   334 CITATIONS   

SEE PROFILE SEE PROFILE

Li Ping Lei
Tsinghua University
75 PUBLICATIONS   683 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Lightweight Structure View project

All content following this page was uploaded by Yao Xiao on 28 November 2019.

The user has requested enhancement of the downloaded file.


Materials and Design 134 (2017) 111–120

Contents lists available at ScienceDirect

Materials and Design

journal homepage: www.elsevier.com/locate/matdes

In situ observation on temperature dependence of martensitic


transformation and plastic deformation in superelastic NiTi shape
memory alloy
Yao Xiao a, Pan Zeng a,⁎, Liping Lei a, Yanzhi Zhang b
a
Department of Mechanical Engineering, Tsinghua University, Beijing 100084, China
b
Institute of Materials, China Academy of Engineering Physics, Mianyang 621900, China

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• The deformation mode of NiTi transits


from localization to delocalization as
temperature increases.
• Plasticity occurs together with martens-
itic transformation at the band front and
its amount increases with temperature.
• Plastic deformation dominates over re-
sidual martensite as the source of strain
irreversibility in superelastic NiTi.

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

Article history: In situ digital image correlation (DIC) and in situ X-ray diffraction (XRD) are applied to investigate the effect of
Received 3 May 2017 temperature on martensitic transformation and plastic deformation in superelastic NiTi shape memory alloy.
Received in revised form 30 July 2017 Via in situ DIC, two well-known deformation modes of NiTi are identified at various temperatures:
Accepted 17 August 2017
(A) localized forward and reverse transformations with little residual strain (b1%); (B) localized forward trans-
Available online 18 August 2017
formation and homogenous reverse transformation with considerable residual strain (N 1%). As temperature in-
Keywords:
creases from 25 °C to 120 °C, the mechanical response of NiTi gradually transits from Type A to Type B. We verify
NiTi shape memory alloy that plastic strain accumulates concurrently as the traverse of the front of localized deformation band. Via in situ
Phase transformation XRD observation, we conclude that it is material plasticity rather than retained martensite that plays a dominant
Plasticity role in the irreversibility of NiTi. The experimental results provide both macroscopic and lattice level scenarios to
Localized deformation understand the temperature dependence of complicated thermomechanical coupling and plasticity in
superelastic NiTi.
© 2017 Elsevier Ltd. All rights reserved.

1. Introduction dentistry and medical fields [1] due to the unique shape memory effect
(SME) and superelasticity (SE) originating from thermoelastic phase
Near-equiatomic NiTi shape memory alloy (SMA) holds great poten- transformation between austenite (cubic structure, B2) and martensite
tial for a broad range of applications in engineering, aeronautics, (monoclinic structure, B19′) [2–4]. It is well established that under ten-
sion, stress-induced transformation of superelastic NiTi leads to local-
⁎ Corresponding author. ized deformation, featured by the coexistence of several localized
E-mail address: zengp@mail.tsinghua.edu.cn (P. Zeng). deformation bands (LDBs) [5–19]. More recently, Zheng et al. [20]

http://dx.doi.org/10.1016/j.matdes.2017.08.037
0264-1275/© 2017 Elsevier Ltd. All rights reserved.
112 Y. Xiao et al. / Materials and Design 134 (2017) 111–120

reported that cyclic band nucleation/annihilation accelerates the micro- heating stage was installed on the Empyrean diffractometer
structure degradation of NiTi and deteriorates the fatigue performance. (PANalytical Inc.) equipped with PIXcel3D detector and with CuKα radi-
These findings further emphasize the importance of understanding the ation at 40 kV and 40 mA. The X-ray was focused on the gauge section of
localized deformation of superelastic NiTi. Apart from macroscopic the sample by laser positioning. The projection of the X-ray on the spec-
investigation, substantial efforts have also been made to elucidate imen's surface was a rectangle with width of about 1 mm and length of
the underlying microstructural mechanisms of localized deformation about 12 mm. The specimen was quasi-statically loaded at the strain
in superelastic NiTi through optical microscopy [21], electron rate of 1.8 × 10−4 s−1 with loading axis perpendicular to the direction
backscattered diffraction (EBSD) [22,23] and X-ray diffraction (XRD) of the incident beam. During X-ray scanning, the grip displacement
[24–27]. was fixed. The scanning range of the specimen was 35° b 2θ b 50° or
Despite of the abundant investigation on superelastic NiTi at room 35° b 2θ b 95°, and the step size was 0.013°.
temperature, the temperature dependence of the mechanical behavior
of superelastic NiTi, however, has not been studied in sufficient detail. 3. Experimental results and discussion
At the temperature above martensite desist temperature (Md), stress-
induced martensite is suppressed. Twinning and dislocation slip domi- 3.1. Overview of in situ DIC observation
nate in the deformation of NiTi, leading to the accumulation of plasticity
[28]. At the temperature between autensite finish temperature (Af) and Fig. 1 shows the applied stress-applied strain response of NiTi at var-
Md, our recent studies [12,13] have shown that the deformation mode ious temperatures. The applied stress is calculated from the ratio be-
of superelastic NiTi undergoes significant change with the elevating tween the applied force and the original cross section area. The
temperature. NiTi exhibits well-known localized deformation and supe- applied strain represents the spatial average of the strain in the gauge
rior reversibility in the relatively low temperature regime, nevertheless section from DIC results. Fig. 2 presents the temperature dependence
NiTi demonstrates localized forward transformation and homogenous of residual strain, nucleation stress and austenitic elastic modulus. The
reverse transformation with considerable residual strain in the relative- nucleation stress of LDB is defined as the maximum stress prior to the
ly high temperature regime. It is widely accepted that the reversible onset of the first LDB, and it is linearly fitted with respect to temperature
strain is due to elastic deformation and stress-induced transformation, in Fig. 2(b). The temperature dependence of the nucleation stress is 5.60
while the irreversible mechanisms are commonly attributed to plastic MPa/°C, in good accordance with the reported results ranging from
deformation and the formation of retained martensite [28–32]. Šittner 5.3 MPa/°C to 8.5 MPa/°C [34,39–43]. The elastic modulus of austenite,
et al. [33] pointed out that the concurrence of plastic deformation and defined as the initial stiffness by fitting the first 0.5% linear response,
stress-induced transformation leads to the elimination of localized de- is shown in Fig. 2(c). Basically, the elastic modulus of austenite displays
formation, whereas the systematic microscopic investigation is missing. increasing trend with elevating temperature, i.e. from 66 GPa at 25 °C to
The objective of this work is to examine the mechanical behavior of 95 GPa at 120 °C, and gradually gets saturated.
superelastic NiTi between Af and Md. The study presents an in-depth in- Figs. 3 to 5 and Figs. S1 to S7 in Supplementary material illustrate
sight into the unusual thermomechanical properties of superelastic sets of full-field DIC contours and the corresponding central line longi-
NiTi. By employing in situ digital image correlation (DIC) and in situ tudinal strain profile at the representative applied strains. In light of
X-ray diffraction (XRD), we acquired the local strain profile, the evolu- the previous research [12,13], two typical deformation modes of NiTi
tion of LDB and the lattice-level data from macroscopic and microscopic are identified: (A) both forward and reverse transformations are local-
perspectives. Another contribution of the work is to provide dependable ized and with little residual strain (b 1%); (B) forward transformation
data which can be used to calibrate and establish the constitutive model is localized while reverse transformation is homogeneous and with
accounting for the temperature dependence of mechanical response of considerable residual strain (N 1%). At temperatures between 25 °C
NiTi, hitherto limited at room temperature and without considering and 55 °C, NiTi exhibits Type A mechanical response. The specimen
the interaction between transformation and plastic deformation in
NiTi [34–38]. In the following, the material properties and experimental
methods are introduced in Section 2. Macroscopic DIC observation, mi-
croscopic XRD observation and discussions are provided in Section 3.
Conclusions are drawn in Section 4.

2. Materials and experiments

The material used in this study was commercially available NiTi


polycrystalline sheets (0.5 mm thickness, 55.82 wt% Ni, Memry Corp.).
The austenite finish temperature Af is 3.3 °C [13], which means that
the specimen is austenitic at and above room temperature. For in situ
DIC observation, the sheet was electro-discharge machined into dog-
bone shape with gauge width of 7 mm and gauge length of 40 mm. Be-
fore the tensile test, a random speckle pattern was painted by spraying
white background and black spot on the surface of the specimen. The
specimen was placed in the environmental chamber and was
strained/unloaded by the testing machine (Shimazu AG-X universal
testing machine) at the strain rate of 2 × 10−4 s−1 to ensure isothermal
condition. A digital camera (Daheng Image, DHSV1410FM) was used to
record the surface morphology with the frequency of 1 Hz and the spa-
tial resolution of captured images was 1392 × 1040 pixels. The spatial
accuracy of DIC results is up to 1 μm, which can meet the requirement
of the experiment [12,13].
For the in situ XRD experiment, the dog-bone tensile specimen with
gauge width of 1.8 mm and gauge length of 14 mm was electro- Fig. 1. Applied stress-applied strain response of NiTi. The onset points of LDB are marked
discharge machined from the sheet. A custom-made mini-tensile/ by arrows.
Y. Xiao et al. / Materials and Design 134 (2017) 111–120 113

Fig. 2. Temperature dependence of (a) residual strain, (b) nucleation stress of LDB (the dashed line is the linearly fitted curve) and (c) elastic modulus of austenite.

undergoes an initial stiff and nearly linear branch during which austen- to loading. As can be seen in the strain profiles in Fig. 3 and Figs. S1 to S3,
ite deforms elastically and homogeneously. At the end of the linear significant strain jump at the LDB front is clearly noticed.
branch, local maximum stress is reached in the applied stress-applied For the cases tested at and above 75 °C, NiTi demonstrates Type B
strain curve and LDB tends to nucleate at one of the end parts of the mechanical response. The specimen undergoes localized deformation
gauge section. Following the stress drop, the stress plateau enables during loading, characterized by the nucleation and propagation of
LDB to propagate at essentially constant load and another LDB may ini- LDB. During unloading, the material undergoes homogenous deforma-
tiate at the other end part of the gauge section. When LDB covers the tion. For the tests at 75 °C and 85 °C, the lower stress plateau and the
whole gauge section, the applied stress takes an abrupt upturn. During transition knee associated with reverse transformation are less distinct,
unloading, the material is subjected to the mechanical evolution reverse compared with the tests at and below 55 °C. When the temperature

Fig. 3. Representative full-field DIC contours and the corresponding central line Fig. 4. Representative full-field DIC contours and the corresponding central line
longitudinal strain profile at various applied strains at 25 °C. longitudinal strain profile at various applied strains at 65 °C.
114 Y. Xiao et al. / Materials and Design 134 (2017) 111–120

3.2. Overview of in situ XRD observation

In the in situ XRD test, the deformation of the specimen is under dis-
placement control. The applied strain value is the nominal strain calcu-
lated from the ratio between the grip displacement and the gauge
length. Upon loading, the strains at which the XRD measurements are
done are selected as: 0% (the reference state), 0.7% (the elastically de-
formed state), 1.5%, 3%, 5% or 6% (within upper stress plateau), 7.5% or
8% (after the termination of the upper stress plateau). Upon unloading,
the strain at which XRD measurement is performed decreases from 6.5%
to the residual strain, with the interval of 1.5% or 1%. The XRD profiles at
25 °C, 45 °C, 65 °C, 85 °C and 105 °C are shown in Figs. 6 to 8. The diffrac-
tion pattern is smoothed by B-spline and indexed by austenitic B2 phase
(a0 = 3.015 Å) and martensitic B19′ phase (a = 2.882 Å, b = 4.123 Å, c =
4.626 Å, β = 97°) [2,44]. The diffraction pattern is further normalized by
the peak intensity of B2(110) at the reference state under the given
temperature.
Upon loading, NiTi first undergoes austenitic elastic deformation
with the decrease of B2(110) peak intensity. When the strain exceeds
0.7%, the commencement of forward transformation is evidenced by
the emergence of B19′ peaks. Basically, martensite with lower symme-
try possesses lower structure factors than austenite with higher sym-
metry [2,45]. Due to the weak diffraction of martensite, the B19′ peaks
at 2θ N 50° are not remarkable and even cannot be accurately indexed
at temperatures higher than 45 °C. Therefore, in this paper, primary at-
tention is paid to the XRD profile with 2θ ranging from 35° to 50°. For
Fig. 5. Representative full-field DIC contours and the corresponding central line tests at 25 °C and 45 °C, B2 (110) peak splits into three B19′ peaks, i.e.
longitudinal strain profile at various applied strains at 120 °C. B19′ (110), B19′ (020) and B19′ (111). B19′ (020) peak undergoes
rapid increment since the beginning of transformation, while B19′
(110) peak or B19′ (111) peak does not become distinct until the elon-
exceeds 85 °C, reverse transformation plateau eventually vanishes and gation increases to 5%. For tests at 65 °C, B2 (110) peak splits into two B19′
the unloading branch becomes linear, similar to the performance of or- peaks, i.e. B19′ (110) and B19′ (020). For tests at 85 °C and 105 °C, only
dinary elastic-plastic metal. the transition between B2 (110) and B19′ (020) can be witnessed.
It should be noted that the LDB evolution of NiTi at 65 °C differs from When the highest applied strain is reached, for test at 25 °C, only B19′
that of NiTi at other temperatures. During loading, the LDB front gradu- peaks are detectable, which is a signature of the completion of
ally gets bent with the ongoing deformation, which is a salient feature of transformation, nevertheless, for tests at and above 45 °C, B2(110)
Type B mechanical response; during unloading, although LDB can be peak and B19′ peaks coexist, which is an indicator of partial
recognized, the strain gradient at the LDB front is much less than the transformation.
cases with Type A mechanical response, manifesting more homogenous Upon unloading, for test at 25 °C, NiTi initially deforms in martensitic
deformation. Therefore, the mechanical response at 65 °C is the tran- phase. When the applied strain is below 5%, reverse transformation ini-
sient response between Type A and Type B. tiates accompanied with the emergence of B2 peak. As unloading

Fig. 6. Representative XRD profiles at various applied strains at 25 °C. Left column: normalized XRD profile within 35° b 2θ b 50°; Right column: selected XRD profile within 35° b 2θ b 95°.
Y. Xiao et al. / Materials and Design 134 (2017) 111–120 115

Fig. 7. Representative XRD profiles at various applied strains at 45 °C. Left column: Normalized XRD profile within 35° b 2θ b 50°; Right column: Selected XRD profile within 35° b 2θ b 95°.

progresses, B19′ (110) peak first disappears, followed by B19′ (111) peak, unloading process, which is consistent with the previous observation
and finally B19′ (020) peak vanishes. For tests at and above 45 °C, at room temperature [5,7,9]. For the cases tested above 75 °C, the
unloading gives rise to simultaneous reverse transformation, evidenced band angle varies with elongation. At first the band inclination obtained
by the sharp increment of B2 (110) peak and the decrement of B19′ from the strain map lies between 55° and 60°. As elongation increases to
peaks. The difference of the XRD profile evolution results from different 3%, the increasing band inclination makes LDB front get more and more
deformation modes and transformation mechanisms at various tempera- horizontal. Eventually, LDB front becomes horizontally arc-shape at rel-
tures [27]. After complete unloading, all B19′ peaks, regardless of testing atively high applied strains, e.g. frames 4 in Fig. 4 and Figs. S4 to S7. From
temperature, revert back to original B2 peaks, i.e. B2 (110), B2 (211) the view of continuum mechanics, the inclination angle of LDB can be
and B2 (220), and no residual B19′ can be figured out in the profile. related to the transformation surface and the associative flow rule (See
Supplementary material). At temperatures between 25 °C and 65 °C,
3.3. Evolutions of LDB and plastic deformation von Mises criterion can be applied to predict the band angle with
theoretical value (θvM) of 54.74° [5,11]. At temperatures above 75 °C,
At temperatures between 25 °C and 65 °C (Figs. 3 and S1 to S3), the deviation from θvM implies the variation of the principal strain rates,
straight LDB front and nearly constant band inclination ranging from stemming from the complicated stress state [11] and the severe plastic
52° to 57° are observed in the strain map throughout the loading- accumulation upon deformation.

Fig. 8. Representative normalized XRD profiles at various applied strains at (a) 65 °C, (b) 85 °C and (c) 105 °C.
116 Y. Xiao et al. / Materials and Design 134 (2017) 111–120

Additional experiments were conducted to check the interaction be-


tween LDB evolution and plastic deformation at three typical tempera-
tures, i.e. 75 °C, 95 °C and 120 °C. The specimens were unloaded as the
length of LDB was about half of the gauge length. Fig. 9 displays sets of
full-field DIC contours and the corresponding central line longitudinal
strain profile at the representative applied strains. Frame 2 corresponds
to the largest deformation state of the specimen, so its strain profile is
identical to the maximum strain distribution. LDB separates NiTi into
Zone I (the region once covered by LDB and with maximum strain
around 7.5%), Zone II (the region without being covered by LDB), and
Zone III (the transient region between Zone I and Zone II). As shown
in Fig. 9, Zone I broadens while Zone II shrinks with increasing deforma-
tion. Zone III shifts along the loading axis and its width keeps nearly un-
changed. The strain variation in residual strain profile (frame
4) indicates the concurrence of transformation and plastic deformation.
The relation between the maximum strain (ɛmax) and residual strain
(ɛres) is illustrated in Fig. 10. All the correlation coefficients (R2) are well
above 0.95. Although we cannot confirm the interdependency of ɛmax Fig. 10. Relation between the maximum strain (data from frames 2 in Fig. 9) and residual
and ɛres, we can at least conclude that ɛmax and ɛres are positively related strain (data from frames 4 in Fig. 9) at 75 °C, 95 °C and 120 °C.
variables. Furthermore, we verify that higher temperature elicits more
severe superelasticity degradation. On one hand, as elucidated in
Fig. 10, the fitted slope of dɛres/dɛmax increases sharply from 0.142 to uncover the underlying mechanism. In this temperature regime, we
0.412 as temperature increases from 75 °C to 120 °C. On the other also note that forward transformation is localized while reverse trans-
hand, as plotted in Fig. 2(a), more drastic residual strain accumulation formation is homogeneous. It can be attributed to the fact that LDB
is clearly identified with elevating temperature, increasing from 0.1% front moves through the virgin microstructure during loading while it
(25 °C) to 3.3% (120 °C), which is in agreement with the previous obser- moves through the plastically deformed microstructure during
vation [41]. unloading.
At temperatures higher than 75 °C, the nucleation and propagation
of LDB are witnessed during forward transformation with the absence 3.4. Interaction between the defect and plastic deformation
of stress plateau, which contradicts common perception of the mechan-
ics of LDB propagation [5,8,9,17,33]. We infer that the intense difference Since the unloaded specimen may only contain little retained mar-
of the mechanical responses within and out of LDB leads to the disap- tensite which X-ray diffractometer fails to resolve, the primary source
pearance of the stress plateau. Further investigation is desired to of irreversibility is the accumulation of material plasticity [25,27,29].

Fig. 9. Representative full-field DIC contours and the corresponding central line longitudinal strain profile at various applied strains at 75 °C, 95 °C and 120 °C. The gauge sections are
separated into Zone I (the region once covered by LDB and with maximum strain around 7.5%), Zone II (the region without being covered by LDB), and Zone III (the transient region
between Zone I and Zone II). The specimens are unloaded at about 4% (frame 2) when LDB covers about half of the gauge section.
Y. Xiao et al. / Materials and Design 134 (2017) 111–120 117

from the XRD profile. D and ɛm are obtained by linearly fitting B2(110),
B2(211) and B2(220) peaks with Eq. (1).
The approximate estimation of the dislocation density ρ holds [47]:

pffiffiffiffiffiffi
ρ¼ 6π εm =bD ð2Þ

where b is the length of the Burgers vector. For body-centered cubic


austenite, b equals to 1/2a0 ⟨111⟩ [27]. Generally, the calculated value
of dislocation density obtained from XRD method is not strictly precise.
In the XRD method, theoretical concepts, instrumental parameters,
crystallographic defects and textural effects will affect the calculated
dislocation density [48]. Nevertheless, comparisons of XRD method
with other alternative methods for measuring dislocation densities
indicate that the results obtained from XRD are fairly reasonable [47,
48], so the calculated value of dislocation density is used for qualitative
analysis. At 25 °C, the dislocation density before deformation (ρo)
and that after deformation (ρd ) are essentially identical at about
Fig. 11. The evolution of FWHM with respect to temperature. The data obtained from
1.5 × 1014 m− 2. At 105 °C, ρd is 1.12 ± 0.02 × 1015 m− 2, an order
virgin NiTi and deformed NiTi are symbolled with up triangle and down triangle,
respectively. Error bars are included, but are usually smaller than the symbol size. of magnitude higher than ρd at 25 °C. The reported ρd required for
delocalization of cyclically deformed NiTi ranges from 8.2 × 1014 m−2
to 2.1 × 1015 m−2 [25,27]. In the paper, ρd of the transient case (tested
The evolution of full width half maximum (FWHM), which scales with at 65 °C) is 6.55 ± 0.17 × 1014 m− 2, in general agreement with the
the density of defects introduced by plastic deformation in grain do- previous results.
mains [46,47], is shown in Fig. 11. For all the virgin specimens, FWHM
of a particular crystallographic plane keeps at a constant level. A single 3.5. Evolution of microstructure
loading-unloading procedure at 25 °C has limited effect on the accumu-
lation of defects, which is in accordance with little plastic deformation at Since the penetration depth of X-ray is around tens of micrometers
25 °C. Deforming at elevating temperature leads to the increment of and the deformation of the specimen may be heterogeneous, the lattice
FWHM, indicating the accumulation of defects or residual strain occurs strain obtained from the XRD profile is the average lattice strain of the
in NiTi. surface, and we are able to analyze the evolution of the microstructure
The concurrence of localized transformation and plastic deforma- qualitatively. In some macroscopic studies, martensite volume fraction
tion, which is shown in Fig. 9, is further discussed from the microscopic (MVF, fM) yields the ratio between local strain and the strain jump at
view. As proposed by Benafan et al. [28,32] and Ezaz et al. [30], the LDB front, presuming that material undergoes complete transformation
transformation-induced plasticity is more ready to be triggered in aus- within Zone I and transformation does not occur within Zone II [49].
tenite. The nucleation and build-up of slip in austenite are required to This hypothesis, however, is quite controversial since microscopic ob-
accommodate the rather high transformation strain upon traversing servations have shown that fM within LDB ranges from 60% to 100%, de-
martensite interfaces [2,28,30]. Elevating temperature gives rise to pending on loading condition as well as loading history [21,24,25,50].
lower yield stress [2] while higher transformation stress due to Besides, the strain jump at LDB front consists of both plastic strain and
Clausius-Clapeyron relation [40], which will greatly facilitate the defect transformation strain. Therefore, in this work, fM is calculated from the
generation in NiTi. Several dislocation systems of austenite, such as XRD results [25,27,50]:
⟨100⟩{011} system [29], ⟨100⟩{001}system [30], {114} compound twin-
ning [28,29] etc., have been experimentally observed. As calculated by P
I HKL;M
Ezaz et al. [30], ⟨100⟩{011} system will be activated at 2.6 GPa, and fM ¼ P P ð3Þ
I HKL;M þ Ihkl;A
⟨111⟩ type slip is expected to be activated over 5 GPa. It is worth noting
that the maximum applied stress in the experiments is no N1 GPa,
let alone the resolved shear stress. Therefore, without extra stress, the where IHKL,M and Ihkl,A denote the integrated intensities of martensite
intense accumulation of defects or plasticity cannot be triggered, just with plane HKL and austenite with plane hkl, respectively. XRD profiles
as what happens in Zone II in Fig. 9. For material within Zone III, the are deconvoluted with Lorentz function to separate the diffraction
strain gradient at the front of macroscopic LDB and the misfit at micro- peaks. Due to the weak diffraction of martensite at 2θ N 50°, only B2
scopic austenite-martensite interface provides additional stress concen- (110), B19′ (110), B19′ (020) and B19′ (111) are taken into account in
tration. Therefore, intense plastic accumulation occurs concurrently Eq. (3). The maximum fM (denoted as fM,max) tends to decrease with in-
with the traverse of LDB front. For material within Zone I, the vanishing creasing temperature. At 25 °C, NiTi is subjected to complete transfor-
of LDB front weakens the extent of stress concentration. Hence the dis- mation, i.e. fM,max = 1. At temperatures higher than 25 °C, the density
location defect will not experience drastic increment and a relatively of defects increases with elevating temperature and the accompanying
constant plastic strain is attained within Zone I. internal stress impedes NiTi undergoing complete transformation
Williamson-Hall equation is applied to analyze the diffraction line- even at the end of the upper stress plateau, e.g. fM,max is only 0.37 at
broadening [46]: 105 °C. From DIC results, the end of the upper stress plateau is the
signature that LDB fully covers the gauge section. Therefore,
β cosθ=λ ¼ 1=D þ Cεm sinθ=λ ð1Þ transformation within the propagating LDB is not completed except
for the case at 25 °C. Furthermore, it is straightforward for us to
extrapolate that other loading conditions associated with significant
β is the integral width, θ is the diffraction angle, λ is the wavelength defect generation, such as mechanical training and thermal cycling,
of the X-ray, D is the average crystallite size, ɛm is the micro strain, and C will also induce incomplete martensitic transformation and
is the constant concerning the nature of inhomogeneous micro strain. In eventually originates in the decrease of reversible transformation
the paper, C is set as 4 [27] and λ equals to 1.5418 Å. β and θ are obtained strain and superelasticity deterioration [25,29,50].
118 Y. Xiao et al. / Materials and Design 134 (2017) 111–120

The lattice strain of a particular crystallographic plane hkl is denoted The evolution of ɛB2(110) is shown in Fig. 12. At 25 °C, the lattice strain
as ɛhkl and is obtained from the following formula: keeps nearly unchanged during transformation. As temperature in-
creases, the evolution of ɛB2(110) traces a nearly closed hysteresis. Upon
dhkl −d0;hkl loading, B2(110) first undergoes elastic deformation when MVF equals
εhkl ¼ ð4Þ
d0;hkl to zero. After the commencement of forward transformation, the lattice
strain decreases mildly with increasing applied strain. A sharp turn is no-
where d0,hkl and dhkl are the d-spacings of the plane hkl before loading ticeable at the end of loading. At temperatures lower than 65 °C, the evo-
and under applied stress, respectively. The lattice strain of austenitic lution of the lattice strain upon unloading is characterized with the
B2(110) is denoted as ɛB2(110). One should note that ɛB2(110) decreases transition knee, and the overall evolution is reverse to that upon loading.
with elongation because the diffraction plane is perpendicular to the At temperatures higher than 65 °C, the lattice response is submitted to lin-
tensile direction and the lattice contracts in the transverse direction early unloading. The change of the lattice response implies the transition
owing to Poisson's effect. of the deformation mode in NiTi. The residual lattice strain is b 0.05%,

Fig. 12. The evolution of ɛB2(110) at (a) 25 °C, (b) 45 °C, (c) 65 °C, (d) 85 °C and (e) 105 °C. Error bars are included, but are usually smaller than the symbol size.
Y. Xiao et al. / Materials and Design 134 (2017) 111–120 119

indicating NiTi essentially transforms back to its original lattice state after [15] Y. Xiao, P. Zeng, L. Lei, Grain size effect on mechanical performance of nanostruc-
tured superelastic NiTi alloy, Mater. Res. Express 4 (2017), 035702.
complete unloading. [16] Y. Liu, Y. Liu, J.V. Humbeeck, Luders-like deformation associated with martensite re-
orientation in NiTi, Scr. Mater. 39 (1998) 1047–1055.
4. Conclusions [17] D. Favier, H. Louche, P. Schlosser, L. Orgeas, P. Vacher, L. Debove, Homogeneous and
heterogeneous deformation mechanisms in an austenitic polycrystalline Ti-50.8 at.%
Ni thin tube under tension. Investigation via temperature and strain fields measure-
In situ digital image correlation and in situ X-ray diffraction are ap- ments, Acta Mater. 55 (2007) 5310–5322.
plied to investigate the effect of temperature on martensitic transforma- [18] V. Grolleau, H. Louche, V. Delobelle, A. Penin, G. Rio, Y. Liu, D. Favier, Assessment of
tension–compression asymmetry of NiTi using circular bulge testing of thin plates,
tion and plastic deformation in superelastic NiTi shape memory alloy.
Scr. Mater. 65 (2011) 347–350.
From the results obtained in the present study, several conclusions [19] V. Delobelle, G. Chagnon, D. Favier, T. Alonso, Study of electropulse heat treatment of
can be made. cold worked NiTi wire: from uniform to localised tensile behavior, J. Mater. Process.
Technol. 227 (2016) 244–250.
[20] L. Zheng, Y. He, Z. Moumni, Investigation on fatigue behaviors of NiTi polycrystalline
(1) Two typical deformation modes of NiTi are identified: (A) both strips under stress-controlled tension via in-situ macro-band observation, Int. J.
forward and reverse transformations are localized and with little Plast. 90 (2017) 116–145.
residual strain (b1%); (B) forward transformation is localized [21] L.C. Brinson, I. Schmidt, R. Lammering, Stress-induced transformation behavior of a
polycrystalline NiTi shape memory alloy: micro and macromechanical investiga-
while reverse transformation is homogeneous and with consid- tions via in situ optical microscopy, J. Mech. Phys. Solids 52 (2004) 1549–1571.
erable residual strain (N 1%). As temperature increases from 25 [22] S. Mao, X. Han, Z. Zhang, M. Wu, The nano-and mesoscopic cooperative collective
°C to 120 °C, the mechanical response of NiTi gradually transits mechanisms of inhomogenous elastic-plastic transitions in polycrystalline TiNi
shape memory alloys, J. Appl. Phys. 101 (2007) 103522.
from Type A to Type B.
[23] S. Mao, J. Luo, Z. Zhang, M. Wu, Y. Liu, X. Han, EBSD studies of the stress-induced B2-
(2) Type A deformation is realized mainly by reversible martensitic B19′ martensitic transformation in NiTi tubes under uniaxial tension and compres-
transformation, while Type B deformation is accomplished by sion, Acta Mater. 58 (2010) 3357–3366.
martensitic transformation alongside drastic plastic deformation. [24] M.L. Young, M.X. Wagner, J. Frenzel, W.W. Schmahl, G. Eggeler, Phase volume frac-
tions and strain measurements in an ultrafine-grained NiTi shape-memory alloy
Plastic deformation occurs together with martensitic transforma- during tensile loading, Acta Mater. 58 (2010) 2344–2354.
tion at the propagating LDB front and its amount increases with [25] E. Polatidis, N. Zotov, E. Bischoff, E.J. Mittemeijer, The effect of cyclic tensile loading
increasing temperature. on the stress-induced transformation mechanism in superelastic NiTi alloys: an in-
situ X-ray diffraction study, Scr. Mater. 100 (2015) 59–62.
(3) The density of defects remaining in superelastic NiTi after a single [26] P. Sedmák, J. Pilch, L. Heller, J. Kopeček, J. Wright, P. Sedlák, M. Frost, P. Šittner,
loading-unloading procedure increases with increasing tempera- Grain-resolved analysis of localized deformation in nickel-titanium wire under ten-
ture. Plastic deformation dominates over residual martensite as sile load, Science 353 (2016) 559–562.
[27] N. Zotov, M. Pfund, E. Polatidis, A.F. Mark, E.J. Mittemeijer, Change of transformation
the source of strain irreversibility in superelastic NiTi. mechanism during pseudoelastic cycling of NiTi shape memory alloys, Mater. Sci.
Eng. A 682 (2017) 178–191.
[28] O. Benafan, R.D. Noebe, S.A. Padula, A. Garg, B. Clausen, S. Vogel, R. Vaidyanathan,
Acknowledgements Temperature dependent deformation of the B2 austenite phase of a NiTi shape
memory alloy, Int. J. Plast. 51 (2013) 103–121.
[29] R. Delville, B. Malard, J. Pilch, P. Sittner, D. Schryvers, Transmission electron micros-
The work is supported by National Natural Science Foundation of copy investigation of dislocation slip during superelastic cycling of Ni–Ti wires, Int. J.
China (No. 51671176). Plast. 27 (2011) 282–297.
[30] T. Ezaz, J. Wang, H. Sehitoglu, H.J. Maier, Plastic deformation of NiTi shape memory
alloys, Acta Mater. 61 (2013) 67–78.
Appendix A. Supplementary data
[31] O. Benafan, R.D. Noebe, S.A. Padula, D.J. Gaydosh, B.A. Lerch, A. Garg, G.S. Bigelow, K.
An, R. Vaidyanathan, Temperature-dependent behavior of a polycrystalline NiTi
Supplementary data to this article can be found online at http://dx. shape memory alloy around the transformation regime, Scr. Mater. 68 (2013)
doi.org/10.1016/j.matdes.2017.08.037. 571–574.
[32] O. Benafan, R.D. Noebe, S.A. Padula, D.W. Brown, S. Vogel, R. Vaidyanathan,
Thermomechanical cycling of a NiTi shape memory alloy-macroscopic response
References and microstructural evolution, Int. J. Plast. 56 (2014) 99–118.
[33] P. Šittner, Y. Liu, V. Novak, On the origin of Lüders-like deformation of NiTi shape
[1] J. Mohd Jani, M. Leary, A. Subic, M.A. Gibson, A review of shape memory alloy re- memory alloys, J. Mech. Phys. Solids 53 (2005) 1719–1746.
search, applications and opportunities, Mater. Des. 56 (2014) 1078–1113. [34] J.A. Shaw, Simulations of localized thermo-mechanical behavior in a NiTi shape
[2] K. Otsuka, X. Ren, Physical metallurgy of Ti-Ni-based shape memory alloys, Prog. memory alloy, Int. J. Plast. 16 (2000) 541–562.
Mater. Sci. 50 (2005) 511–678. [35] J.A. Shaw, A thermochemical model for a 1-D shape memory alloy wire with prop-
[3] C. Cisse, W. Zaki, T.B. Zineb, A review of constitutive models and modeling tech- agating instabilities, Int. J. Solids Struct. 39 (2002) 1275–1305.
niques for shape memory alloys, Int. J. Plast. 76 (2016) 244–284. [36] A. Duval, M. Haboussi, T.B. Zineb, Modelling of localization and propagation of phase
[4] P. Chowdhury, H. Sehitoglu, A revisit to atomistic rationale for slip in shape memory transformation in superelastic SMA by a gradient nonlocal approach, Int. J. Solids
alloys, Prog. Mater. Sci. 85 (2017) 1–42. Struct. 48 (2011) 1879–1893.
[5] J.A. Shaw, S. Kyriakides, On the nucleation and propagation of phase transformation [37] H. Badnava, M. Kadkhodaei, M. Mashayekhi, A non-local implicit gradient-enhanced
fronts in a NiTi alloy, Acta Mater. 45 (1997) 683–700. model for unstable behaviors of pseudoelastic shape memory alloys in tensile load-
[6] Z.Q. Li, Q.P. Sun, Phase transformation in superelastic NiTi polycrystalline micro- ing, Int. J. Solids Struct. 51 (2014) 4015–4025.
tubes under tension and torsion from localization to homogeneous deformation, [38] H. Badnava, M. Kadkhodaei, M. Mashayekhi, Modeling of unstable behaviors of shape
Int. J. Solids Struct. 39 (2002) 3797–3809. memory alloys during localization and propagation of phase transformation using a
[7] M.A. Iadicola, J.A. Shaw, Rate and thermal sensitivities of unstable transformation gradient-enhanced model, J. Intell. Mater. Syst. Struct. 26 (2015) 2531–2546.
behavior in a shape memory alloy, Int. J. Plast. 20 (2004) 577–605. [39] B.C. Chang, J.A. Shaw, M.A. Iadicola, Thermodynamics of shape memory alloy wire:
[8] S. Daly, G. Ravichandran, K. Bhattacharya, Stress-induced martensitic phase trans- modeling, experiments, and application, Contin. Mech. Thermodyn. 18 (2006)
formation in thin sheets of Nitinol, Acta Mater. 55 (2007) 3593–3600. 83–118.
[9] J.F. Hallai, S. Kyriakides, Underlying material response for Lüders-like instabilities, [40] Y. Liu, A. Mahmud, F. Kursawe, T.H. Nam, Effect of pseudoelastic cycling on the
Int. J. Plast. 47 (2013) 1–12. Clausius-Clapeyron relation for stress-induced martensitic transformation in NiTi,
[10] B. Reedlunn, C.B. Churchill, E.E. Nelson, J.A. Shaw, S.H. Daly, Tension, compression J. Alloys Compd. 449 (2008) 82–87.
and bending of superelastic shape memory tubes, J. Mech. Phys. Solids 63 (2014) [41] Y. Wu, E. Ertekin, H. Sehitoglu, Elastocaloric cooling capacity of shape memory alloys
506–537. – role of deformation temperatures, mechanical cycling, stress hysteresis and inho-
[11] N.J. Bechle, S. Kyriakides, Evolution of localization in pseudoelastic NiTi tubes under mogeneity of transformation, Acta Mater. 135 (2017) 158–176.
biaxial stress states, Int. J. Plast. 82 (2016) 1–31. [42] H. Yin, Y. He, Q. Sun, Effect of deformation frequency on temperature and stress os-
[12] Y. Xiao, P. Zeng, L. Lei, H. Du, Local mechanical response of superelastic NiTi shape- cillations in cyclic phase transition of NiTi shape memory alloy, J. Mech. Phys. Solids
memory alloy under uniaxial loading, Shape Mem. Superelasticity 1 (2015) 67 (2014) 100–128.
468–478. [43] Q. Kan, C. Yu, G. Kang, J. Li, W. Yan, Experimental observations on rate-dependent
[13] Y. Xiao, P. Zeng, L. Lei, Experimental investigation on local mechanical response of cyclic deformation of super-elastic NiTi shape memory alloy, Mech. Mater. 97
superelastic NiTi shape memory alloy, Smart Mater. Struct. 25 (2016), 017002. (2016) 48–58.
[14] Y. Xiao, P. Zeng, L. Lei, Experimental observations on mechanical response of three- [44] Y. Kudoh, M. Tokonami, S. Miyazaki, K. Otsuka, Crystal structure of the martensite in
phase NiTi shape memory alloy under uniaxial tension, Mater. Res. Express 3 (2016) Ti–49.2 at% Ni alloy analyzed by the single crystal diffraction method, Acta Metall.
105701. 33 (1985) 2049–2056.
120 Y. Xiao et al. / Materials and Design 134 (2017) 111–120

[45] A. Ahadi, Q. Sun, Stress-induced nanoscale phase transition in superelastic NiTi by in [48] G. Dini, R. Ueji, A. Najafizadeh, S.M. Monir-Vaghefi, Flow stress analysis of TWIP steel
situ X-ray diffraction, Acta Mater. 90 (2015) 272–281. via the XRD measurement of dislocation density, Mater. Sci. Eng. A 527 (2010)
[46] G.K. Williamson, W.H. Hall, X-ray line broadening from filed aluminium and wol- 2759–2763.
fram, Acta Metall. 1 (1953) 22–31. [49] K. Kim, S. Daly, Martensite strain memory in the shape memory alloy nickel-
[47] G.K. Williamson, R.E. Smallman, Dislocation densities in some annealed and titanium under mechanical cycling, Exp. Mech. 51 (2011) 641–652.
coldworked metals from measurements on the X-ray Debye-Scherrer spectrum, [50] P. Sedmák, P. Šittner, J. Pilch, C. Curfs, Instability of cyclic superelastic deformation of
Philos. Mag. 1 (1956) 34–46. NiTi investigated by synchrotron X-ray diffraction, Acta Mater. 94 (2015) 257–270.

View publication stats

You might also like