You are on page 1of 8

Acta Materialia 227 (2022) 117705

Contents lists available at ScienceDirect

Acta Materialia
journal homepage: www.elsevier.com/locate/actamat

Full length article

New dislocation dissociation accompanied by anti-phase shuffling in


the α  martensite phase of a Ti alloy
Masaki Tahara a,b,∗, Nao Otaki a,b,1, Daichi Minami c,2, Tokuteru Uesugi c, Yorinobu Takigawa c,
Kenji Higashi c,3, Tomonari Inamura a,b, Hideki Hosoda a,b
a
Laboratory for Materials and Structures, Institute of Innovative Research, Tokyo Institute of Technology, 4259 Nagatsutacho, Midori-ku, Yokohama
226-8503, Japan
b
Laboratory for Future Interdisciplinary Research of Science and Technology, Institute of Innovative Research, Tokyo Institute of Technology, 4259
Nagatsutacho, Midori-ku, Yokohama 226-8503, Japan
c
Department of Materials Science, Graduate School of Engineering, Osaka Prefecture University, 1-1, Gakuen-cho, Naka-ku, Sakai, Osaka 599-8531, Japan

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

Article history: Martensite phase plays a crucial role in the shape memory effect and superelasticity. In this study, a new
Received 7 June 2021 dislocation dissociation was found in the α  martensite phase of a Ti-based biomedical shape memory
Revised 1 February 2022
alloy by systematic transmission electron microscopy analysis and first-principles calculations. The slip
Accepted 1 February 2022
direction of dislocation was along 101, which was recently found in a single-crystalline martensite fab-
Available online 5 February 2022
ricated using stress-induced martensitic transformation. The 101 dislocation dissociated into two partial
Keywords: dislocations, and the Burgers vector for both partial dislocations was equal to 1/2<101>. A planar fault
martensitic transformation with an anti-phase shuffling of basal plane was observed between the dissociated partial dislocations. The
dislocation dissociation reaction of the 101 dislocation was summarized as 101 → 1/2<101> + anti-phase shuf-
first principle calculation fling boundary + 1/2<101>. Most of the observed dislocations were in the form of screw dislocations,
titanium alloy and the dissociated partial dislocations were twisted around the dislocation line.
transmission electron microscopy
© 2022 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

1. Introduction martensite phase is also crucial for the fundamental understand-


ing of martensitic transformation and the development of shape
The martensitic transformation in the metastable β -Ti alloy memory effect and superelasticity of β -Ti alloys, it has not been
from the bcc β phase (parent phase) to the c-centered orthorhom- investigated in detail [21–24]. This is mainly because of the dif-
bic α  phase (martensite phase) has attracted considerable atten- ficulty involved in preparing single crystals of the α  martensite
tion, because the metastable β -Ti alloy is a promising candidate phase. We recently synthesized an α  martensite single crystal
for Ni-free biomedical shape memory alloys [1,2]. Significant ef- of Ti-27mol%Nb alloy using a stress-induced martensitic transfor-
forts have been dedicated over the past decade to clarify the fun- mation of a single crystal of the parent β phase, and reported
damental aspect of the martensitic transformation of β -Ti alloys, the plastic deformation behavior of single-crystalline α  marten-
such as the effect of alloying element on the martensitic transfor- site [25]. Two slip systems (with a Burgers vector (b) parallel to
mation [3–14] and the crystallography of martensitic transforma- the <110>o or <101>o directions) and two deformation twinning
tion [15–20]. Although the plastic deformation behavior of the α  systems ({130}o <3̄10>o and {103}o <3̄01>o ) were operated using
compressive deformation at room temperature. These operated slip
systems and twinning systems in the α  martensite phase corre-

Corresponding author at: Laboratory for Materials and Structures, Institute of
sponded to the <111>b slip (i.e., b // <111>b ) and {332}b <113̄>b
Innovative Research, Tokyo Institute of Technology, 4259 Nagatsutacho, Midori-ku, twinning in the parent β phase, respectively; the subscripts ‘o’
Yokohama 226-8503, Japan. and ‘b’ indicate the α  and β phases, respectively. The <111>b
E-mail address: tahara.m.aa@m.titech.ac.jp (M. Tahara). slip and {332}b <113̄>b twinning in the β phase have been exten-
sively observed in metastable β -Ti alloys [26–30]. The operation
1
Graduate student, Tokyo Institute of Technology / Current affiliation: Nippon
Steel Corporation, 1–8 Fusoucho, Amagasaki, Hyogo 660–0891, Japan.
2
of a <110>o dislocation in the α  martensite phase has been re-
Graduate student, Osaka Prefecture University / Current affiliation: Kanagawa
Institute of Industrial Science and Technology, 705–1, Shimoimaizumi, Ebina, Kana- ported by Tobe [21] in a tensile-deformed Ti-20mol%Nb polycrys-
gawa 243–0435, Japan. tal alloy, and the Burgers vector of the [110]o dislocation (b = 1/2
3
Osaka Prefecture University College of Technology, 26-12, Saiwai-cho, Neyagawa, [110]o ) is shown in Fig. 1. However, the operation of dislocation
Osaka, 572-8572, Japan.

https://doi.org/10.1016/j.actamat.2022.117705
1359-6454/© 2022 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
M. Tahara, N. Otaki, D. Minami et al. Acta Materialia 227 (2022) 117705

The first-principles calculations were performed using the Cam-


bridge Serial Total Energy Package (CASTEP) based on density func-
tional theory (DFT) [31]. The electronic exchange-correlation en-
ergy used was given by the generalized gradient approximation
(GGA) [32] with the plane-wave ultrasoft pseudopotential [33].
A supercell containing 18 Ti atoms and 6 Nb atoms (i.e., Ti-
25mol%Nb) with a vacuum region of ∼1 nm between the periodi-
cally repeated slabs was used for the first-principles calculations in
this study, as shown in the Supplementary Materials (Fig. S1). After
adequate tests of convergence for the total energy, a k-points mesh
of 6 × 8 × 1 and a cut-off energy of 350 eV for the plane-wave ba-
Fig. 1. Burgers vectors of the [110]o and [101]o dislocations. Basal plane shuffling sis were used. The convergence tests showed that the error bar for
along [010]o is indicated by the thick black arrow, and δ represents the magnitude
the total energy was less than 1 meV/atom with the above param-
of shuffling.
eters. The details regarding γ -surface calculations are presented in
the Supplementary Materials.
slip along <101>o , which was recently discovered in a compressed
single-crystalline sample by our group [25], in the α  martensite
3. Results
phase is not clear. The Burgers vector of the [101]o dislocation
in α  martensite is shown in Fig. 1. Despite both the [110]o and
3.1. Prediction of dislocation dissociation
[101]o dislocations in the α  phase corresponding to the <111>b
dislocation in the β phase, the magnitude of [101]o dislocation is
3.1.1. Dissociation models
approximately twice as large as that of [110]o . This is because of
The expected dissociation models of perfect dislocation with
the shuffling of the (001)o basal plane along the adjacent [010]o
a Burgers vector, bT = [101]o , are shown in Fig. 2. The Burgers
direction (indicated by the thick black arrow in Fig. 1), which was
vectors of the leading and trailing dislocations are indicated by
introduced by the β -α  martensitic transformation. Owing to this
b1 = [b1x , b1y , b1z ]o and b2 = [b2x , b2y , b2z ]o , respectively. As
basal plane shuffling, the translation vector of the lattice along the
shown in Fig. 1, the adjacent (001)o planes shift along the <010>o
[101]o direction (i.e., b of perfect dislocation) is approximately two
direction owing to the atomic shuffles introduced by the marten-
times larger than that of the [110]o direction; with respect to the
sitic transformation. The magnitude of this shift, which is indicated
Ti-27mol%Nb alloy [3,25], the values of 0.563 and 0.228 nm cor-
by δ changes from 0 (bcc) to 1/6 (hcp); the shift in binary Ti-(8–
respond to the [101]o and [110]o directions, respectively. Although
22)mol%Nb alloys is ∼0.07–0.1 [34–36]. The following dissociation
the <101>o dislocation can be expected to be dissociated into two
partial dislocations and a planar fault [25], there is no conclusive
evidence for dislocation dissociation. Therefore, the proposed study
aims to clarify the dissociation process of <101>o dislocation in α 
martensite via first-principles calculations and systematic trans-
mission electron microscopy (TEM) investigations.

2. Materials and methods

The mother alloy rod with a composition of Ti-27mol%Nb was


fabricated via the Ar arc melting method using high-purity Ti
(99.99%) and Nb (99.9%). The single crystal of Ti-27mol%Nb alloy
was prepared using an optical floating zone method with a growth
rate of 5 mm/h under high-purity Ar flow. The crystallographic ori-
entation of the single crystal was determined by X-ray back Laue
diffraction. The rectangular specimen (3 × 3 × 6 mm3 ) for the
compression test was prepared using a precision diamond wheel
cutter with a compression axis of [1̄28]b . When compressive stress
is applied in this direction, the sample has the largest Schmid fac-
tor for [101]o slip, and only [101]o slip is activated. The specimen
was solution treated at 1173 K for 3.6 ks in an Ar atmosphere, fol-
lowed by water quenching. To remove the oxidized surface layer
formed by water quenching, chemical etching was performed at
∼333 K with a solution of HF:HNO3 :H2 O = 7:8:10 in volume. The
cyclic loading–unloading compression test was performed at 293 K
with a strain rate of 3.3 × 10−4 s − 1 . The sample surface during
compression was observed by optical microscopy. Details regard-
ing sample preparation and compression test settings have been
described in our previous report [25]. The specimen for TEM anal-
ysis was prepared using the focused ion beam (FIB) technique. A
cryo-FIB milling system (FEI, Scios) was used to preclude reverse
martensitic transformation and the formation of an isothermal ω
phase, which can occur during FIB milling owing to localized sam-
ple heating. TEM observations were carried out at room tempera- Fig. 2. Expected dissociation models of the [101]o dislocation. (a) and (c) Model
ture using a JEOL JEM2100 instrument operated at 200 kV. representing b1y = b2y = 0; (b) and (d) model representing b1y = −b2y = δ .

2
M. Tahara, N. Otaki, D. Minami et al. Acta Materialia 227 (2022) 117705

models can be expected by focusing on the y components of the energy exists at a point indicated by “X”(0.45[101]o ). This im-
two partial dislocations (b1y and b2y ): plies that the [101]o dislocation dissociates into two partial dis-
locations, and the stable planar fault exists at 0.45[101]o , i.e., the
(i) b1y = b2y = 0, (Fig. 2a) Burgers vectors of the two partial dislocations (b1 and b2 ) con-
(ii) b1y = −b2y = δ (Fig. 2b) nect at 0.45[101]o , as shown in Fig. 3b. To investigate the effect of
disorder in the atomic configuration on the GSF energy, we per-
The (1̄01)o plane was used as a slip plane to investigate the
formed additional calculations at 0.45[101]o and 1/2[101]o using
two dissociation models of the [101]o dislocation in this study.
four supercells with different configurations of Nb atoms from that
Although the macroscopic slip plane could not be determined in
shown in Fig. S1. These four supercells included the case that Nb
our previous study owing to wavy slip traces [25], the maxi-
atoms across the fault plane were positioned in the nearest neigh-
mum resolved shear-stress plane in the sample compressed along
bor sites to each other. The standard errors of the GSF energies at
[0.98, 0.06, 0.18]o in α  martensite (= [1̄28]b in the β parent
0.45[101]o and 1/2[101]o were 4.6 mJ/m2 and 5.1 mJ/m2 , respec-
phase) was determined to be (100, 24, 99)o , which was close to
tively, which are much smaller than the magnitude of GSF energy.
(1̄01)o . Therefore, the (1̄01)o plane was employed as the slip plane
The differences in GSF energies between 0.45[101]o and 1/2[101]o
in the dissociation models of [101]o dislocation. A six-layer period-
ranged from −8.0 mJ/m2 to +3.3 mJ/m2 depending on the super-
icity exists for the (1̄01)o planes (A, B, C, D, E, and F) in α  marten-
cells. This implies more stable planar fault exists at 1/2[101]o than
site. The slip layer (A) and two layers underneath it (B and C) are
0.45[101]o in different supercells. Although the magnitudes of the
shown in Figs. 2c and 2d. The shuffling along <010>o direction in
Burgers vectors of partial dislocations are slightly different, the γ -
both the models in the A’ layer, which slips owing to the lead-
surface calculation supports the dissociation model (i) (Fig. 2a), be-
ing partial dislocation (b1 ), becomes “anti-phase” compared to the
cause the y components of the partial dislocations are zero.
original A layer. After passing the trailing partial dislocation (b2 ),
the shuffling phase of the slip layer returns to the original (A) in 3.2. Fabrication of specimen for TEM analysis
both the models. Therefore, the planar fault with an anti-phase
shuffling can be expected to exist between the partial dislocations. The TEM specimen of the α  martensite single crystal, in which
the [101]o dislocation slip was operated, was fabricated to ex-
3.1.2. γ -surface analysis perimentally elucidate the dissociation of [101]o dislocation. After
The generalized stacking fault (GSF) energy, which is denoted as the solution treatment, the single-crystalline sample existed in the
a γ -surface, is a critical parameter that is considered during analy- parent β phase at room temperature, and stress-induced marten-
sis of slip systems. In this study, the γ -surface of the (1̄01)o [101]o sitic transformation was induced via compression. Fig. 4a shows
slip was calculated using the first-principles plane-wave pseudopo- the stress–strain curve obtained in the compression test. Loading–
tential method. As shown in Fig. 3, the local minimum of GSF unloading was repeated for the same sample, and the first (black
arrow) and second yielding (white arrow) correspond to the crit-
ical stress required for inducing α  martensite and slip deforma-
tion, respectively. The sudden drop in stress after the second yield-
ing occurs because of the introduction of deformation twinning
[25]. As shown in our previous study [25], only the single lat-
tice corresponding variant (CV) of α  martensite is induced from
the parent β phase, and the entire sample transforms into the
single CV upon compression. This suggests that the sample be-
comes a martensite single crystal upon compression. Internal twin-
ning, which is typically observed in self-accommodated marten-
site (i.e., thermally induced martensite), is not noticed in the sam-
ple compressed in this study. The compression axes of single-
crystalline sample in the parent β phase (before compression) and
α  martensite phase (after stress-induced martensitic transforma-
tion) are shown in Figs. 4b and 4c, respectively. After the stress-
induced martensitic transformation (indicated by the black arrow),
the compression strain does not recover upon unloading. This in-
dicates that the stress-induced α  martensite does not reverse
transform into the parent β phase upon unloading and remains as

Fig. 3. (a) (1̄01)o γ -surface obtained from first-principles calculations for the Ti-
25Nb alloy by changing the color bar range from 0 to 30 0 0 mJ/m2 . The easiest
slip path is noted to be along ψ [010]o = 0. “X” indicates the stable position of
the planar fault introduced by the dislocation dissociation. Here, ξ and ψ represent
the magnitudes of the [101]o and [010]o lattice vectors on the (1̄01)o γ -surface,
respectively. (b) GSF energy curve of the (1̄01)o [101]o slip. The minimum of GSF
energy occurs at ξ = 0.45, which corresponds to the position of stable planar fault, Fig. 4. (a) Stress–strain curves obtained in the compression test for the Ti-27Nb
indicating that the [101]o dislocation can dissociate into two partial dislocations single-crystal alloy [25]. Compression axes in the (b) parent β and (c) stress-
connected by the stable planar fault. induced α  martensite phases.

3
M. Tahara, N. Otaki, D. Minami et al. Acta Materialia 227 (2022) 117705

such after the compression test. Plastic deformation of the single- Table 1
Summary of the g∗ •b analyses. ◦: visible and ×: invisible.
crystalline α  martensite, which was formed via stress-induced
martensitic transformation, occurred upon further compression. Zone g∗ Visibility
Two-face trace analyses of surface traces of the unloaded sample [0 1 0] 2 0 0 ◦
were subsequently performed, and the following plastic deforma- [0 1 0] 4 0 0 ◦
tion modes were operated [25]: (1) dislocation slip along [101]o , [0 1 0] 0 0 2̄ ◦
(2) {130}o <3̄10>o twinning, and (3) {103}o <3̄01>o twinning. The [0 1 0] 0 0 4̄ ◦
[0 1 0] 0 0 6̄ ◦
TEM specimen was prepared using an unloaded single-crystalline
[0 1 0] 2 0 2 ◦
sample using cryo-focused ion beam (FIB) milling. FIB milling typi- [0 1 0] 4 0 4 ◦
cally causes localized heat damage during milling, which results in [0 1 0] 2 0 2̄ ×
the reverse martensitic transformation from the α  to the parent β [0 1 0] 4 0 4̄ ×
phase and the formation of an isothermal ω phase. Therefore, the [1 1 0] 1 1̄ 1̄ ×
[1 1 0] 2 2̄ 2̄ ×
TEM samples in this study were prepared by FIB milling at a low [1 1 0] 3 3̄ 3̄ ×
temperature to obtain samples without reverse martensitic trans- [1 1 0] 4 4̄ 4̄ ×
formation. [1 1 0] 1 1̄ 0 ◦
[1 1 0] 2 2̄ 0 ◦
[1 1 0] 3 3̄ 0 ◦
3.3. TEM analysis of dissociated dislocations [1 1 0] 4 4̄ 0 ◦
[1 0 0] 0 2̄ 0 ×
Fig. 5a shows the low-magnification bright field (BF) TEM im- [1 0 0] 0 4̄ 0 ×
age of the sample. The entire sample represents the α  marten- [1 0 0] 0 6̄ 0 ×
[1 0 0] 0 8̄ 0 ×
site phase and no internal twinning is observed. Although the
(103)o [3̄01]o twinning is noticed, this twinning is introduced be-
cause of plastic deformation in the single-crystalline α  phase. The
straight dislocations and dislocation loops are homogeneously dis-
similar (or identical), i.e., the observed dislocation pair does not
tributed in the sample. The high-magnification BF image featuring
correspond to the dislocation dipole but to the dissociated partial
the enlarged small rectangular region in Fig. 5a (indicated by “b”)
dislocations.
is in Fig. 5b. A contrast in two parallel dislocation lines is observed,
To experimentally determine the Burgers vectors of the partial
and Fig. 5c shows the contrast profile of the A-B line in Fig. 5b.
dislocations (b1 and b2 ), g∗ •b analyses with an axial dark field (DF)
In both the dislocations, bright and dark contrasts (represented
mode were performed, where g∗ represents the reciprocal lattice
by white and black arrows in Fig. 5c, respectively) are observed
vector of a diffracting plane. Three types of zones were selected for
on the A and B sides, respectively. This suggests that the Burg-
the g∗ •b analyses: [010]o (Fig. 6), [110]o (Fig. 7), and [100]o zones
ers vectors of the two adjacent dislocations are not opposite but
(Fig. 8). The g∗ •b analyses for the [010]o and [110]o zones were
carried out in similar regions of the same sample, whereas that for
the [100]o zone was performed using a different sample. The BF
images of dislocations in the [010]o , [110]o , and [100]o zones are
shown in Figs. 6a, 7a, and 8a, respectively. The dislocation pairs
are clearly observed in each BF image as indicated by the red ar-
rows. DF images corresponding to the BF images were obtained us-
ing several g∗ values in each zone; the positions of the dislocation
pairs observed in the BF mode are indicated by the red arrows in
all the DF images. There are no significant differences in the visi-
bility of two partial dislocations in the DF images. The contrasts in
dislocation visibilities in the DF images for each g∗ are summarized
in Table 1.
The Burgers vector of perfect dislocation (bT ) corresponds to
[101]o which is parallel to the slip direction, as determined by sur-
face trace analysis [25]. The Burgers vectors of the two partial dis-
locations in the present study, b1 = [b1x , b1y , b1z ]o and b2 = [b2x ,
b2y , b2z ]o , were determined using the g∗ •b analyses as described
henceforth. The contrasts in dislocation lines were observed in the
DF images for g∗ = 200, 400, 002̄, 004̄, and 006̄ in the [010]o zone
(Figs. 6b–f). This suggested that the x and z components of b1 and
b2 (b1x , b1z , b2x , and b2z ) were not zero. In contrast, the y compo-
nents of b1 and b2 (b1y and b2y ) were zero, because the contrast
of dislocations disappeared in the DF images for g∗ = 02̄0, 04̄0,
06̄0, and 08̄0 in the [100]o zone (Figs. 8b–e). The dislocation con-
trast disappears in g∗ •b analysis when the dot product of g∗ and b
equals zero. In these experiments, b1 and b2 were perpendicular to
g∗ (02̄0, 04̄0, 06̄0, and 08̄0); therefore, the y components of partial
dislocations (b1y and b2y ) were zero. The disappearance of disloca-
tion contrast is also observed in the DF images for g∗ = 202̄, 404̄,
11̄1̄, 22̄2̄, 33̄3̄, and 44̄4̄ in the [010]o (Figs. 6i and 6j) and [110]o
Fig. 5. (a) (b) Bright field transmission electron microscopy (BF-TEM) images of dis- zones (Figs. 7b–e). Considering b1y = b2y = 0, these results indi-
locations. The small rectangular region in (a) is enlarged in (b). (c) Contrast profile cate that the x and z components of the partial dislocations were
corresponding to the A-B line shown in (b). equal (i.e., b1x = b1z and b2x = b2z ). Therefore, the Burgers vec-

4
M. Tahara, N. Otaki, D. Minami et al. Acta Materialia 227 (2022) 117705

Fig. 6. TEM images obtained in the [010]o zone. Arrows indicate the positions of dislocations. (a) BF image. Dark field (DF) images obtained for (b) g∗ = 200, (c) g∗ = 400,
(d) g∗ = 002̄, (e) g∗ = 004̄, (f) g∗ = 006̄, (g) g∗ = 202, (h) g∗ = 404, (i) g∗ = 202̄, and (j) g∗ = 404̄.

tors of the two partial dislocations (b1 and b2 ) were parallel to pair in Fig. 9a are also observed in Figs. 9b and 9c. At the position
that of the perfect dislocation, bT ; this corresponds to the disso- indicated by the black arrow in the [010]o zone (Fig. 9a), the two
ciation described in model (i), which was expected as mentioned partial dislocations coalesce and are observed as a single disloca-
above. When b1 equals b2 , they correspond to 1/2[101]o , because tion line, whereas they clearly separate into two dislocation lines
bT = b1 + b2 = [101]o . in the [110]o zone (Fig. 9c), and vice versa for the position indi-
Several BF images with different zone axes were obtained at the cated by the white arrow. This indicates that the [101]o dislocation
same position by tilting the TEM sample to analyze the direction does not dissociate on a specific plane but on planes that lie in
of the dislocation line (u). The sample was tilted along [001]o , and the [101]o zone parallel to u. As shown in the schematic image
the same reciprocal lattice vector (g∗ = 002) was used to obtain (Fig. 9e), the two partial dislocations (P1 and P2 ) twist with the
the BF images. The representative BF images are shown in Figs. 9a– axis parallel to u.
c. The direction of the incident beam (B) in each BF image and
the projected direction of the dislocation line on the BF image (u’) 4. Discussion
were determined using Kikuchi line analysis. The B-u’ planes for
each zone axis were subsequently determined and plotted as large As expected, the [101]o dislocation dissociates into two partial
circles in a stereogram of α  martensite centered at 001 (Fig. 9d). dislocations (P1 and P2 ). However, the contrasts of two kinds of
The intersection point of the large circles corresponds to u because planar faults are also observed in the DF image for g∗ = 11̄0 in
the B-u’ planes are always parallel to u. Therefore, the dislocation the [110]o zone (Fig. 7f): (1) a curved “ribbon-like” contrast ho-
line (u) of the observed dislocations was determined to be [101]o , mogeneously distributed in the sample and (2) a straight “belt-
which was parallel to b1 and b2 (and bT ). This indicated that the like” contrast observed between the partial dislocations. Although
straight partial dislocation observed by TEM was a screw disloca- the contrast is blurred, similar planar faults are visible in the DF
tion. image for g∗ = 33̄0 in the [110]o zone, as shown in Fig. 7h. An
The detailed configuration of the dissociated partial dislocations enlarged DF image of the planar faults is shown in the inset of
was also investigated by tilting the sample. In the BF image for Fig. 7f. The positions of partial dislocations are indicated by the
the [010]o zone (Fig. 9a), the contrast in the two partial dislo- double-headed arrows, and the contrast in belt-like planar fault
cations coalesces and separates at the positions indicated by the is observed between the partial dislocations. This belt-like con-
black and white arrows, respectively. After tilting the sample along trast corresponds to the planar fault that exhibits anti-phase shuf-
the [001]o axis, similar positions corresponding to the dislocation fling of the basal plane owing to the introduction of partial dis-

5
M. Tahara, N. Otaki, D. Minami et al. Acta Materialia 227 (2022) 117705

Fig. 7. TEM images obtained in the [110]o zone. Arrows indicate the positions of dislocations. (a) BF image; DF images obtained for (b) g∗ = 11̄1̄, (c) g∗ = 22̄2̄, (d) g∗ = 33̄3̄,
(e) g∗ = 44̄4̄, (f) g∗ = 11̄0, (g) g∗ = 22̄0, (h) g∗ = 33̄0, and (i) g∗ = 44̄0. The rectangular region is enlarged in the inset in (f).

Fig. 8. TEM images obtained in the [100]o zone. Arrows indicate the position of dislocations. (a) BF image; DF images obtained for (b) g∗ = 02̄0, (c) g∗ = 04̄0, (d) g∗ = 06̄0,
and (e) g∗ = 08̄0.

location, which was expected as mentioned earlier (Fig. 2c). This alloys, such as Ti–Ni [41] and Ti–Pd [42]. Notably, however, the dis-
planar fault observed herein is denoted as an anti-phase shuffling location dissociation-type APSB was first observed in this study. In
boundary (APSB). The ribbon-like planar fault has been observed both types of APSB, the shuffling phase of the basal plane was in-
in quenched α  and α  (hcp) martensite phases of Ti alloys [18,37- verted, whereas the displacement vectors (R) of the planar faults
40], and is formed by the phase shift of atom shuffling of the basal were slightly different: (1) R = [1/2, δ , 1/2]o [40] or [22/50, 2/50,
plane during martensitic transformation. There are two alterna- 1/2]o [18] in grown-in type APSB, and (2) R = b1 = [1/2, 0, 1/2]o in
tives for the direction of basal plane shuffling in the martensitic dislocation dissociation-type APSB. The APSBs in martensite feature
transformation (i.e., [010]o and [01̄0]o ), and the ribbon-like pla- several similarities to the anti-phase boundaries of ordered alloys,
nar fault is formed at the boundary of two domains with differ- such as formation mechanisms and morphologies. However, a dis-
ent shuffling directions. This suggests that the ribbon-like planar location dissociation-type APSB can be formed when the Burgers
fault is a grown-in type APSB introduced by nucleation and growth vector is not parallel to the shuffling plane of (001)o , because the
of martensite domains. Similar types of grown-in type APSBs have regularity of the basal plane shuffling is responsible for the APSB
been reported in the martensite phase of several shape memory in martensite.

6
M. Tahara, N. Otaki, D. Minami et al. Acta Materialia 227 (2022) 117705

Funding

JSPS KAKENHI JP17H04959 (MT)


JSPS KAKENHI JP19H02417 (MT)
JSPS KAKENHI JP20K20544 (HH)
Titanium Research Grant from the Japan Titanium Society (MT)

Declaration of Competing Interest

The authors declare that they have no known competing finan-


cial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Acknowledgments

The authors would like to thank Prof. Nishida of Kyushu Univer-


sity for TEM sample preparation by cryo-FIB.

Supplementary materials

Supplementary material associated with this article can be


found, in the online version, at doi:10.1016/j.actamat.2022.117705.

References
Fig. 9. BF-TEM images of the [101]o dislocation obtained by tilting the sample for
the (a) [010]o , (b) [130]o , and (c) [110]o zones. (d) Stereogram of α  martensite cen- [1] S. Miyazaki, H.Y. Kim, H. Hosoda, Development and characterization of Ni-free
tered at 001. B-u’ planes are plotted as large circles, and the point of intersection Ti-base shape memory and superelastic alloys, Mater. Sci. Eng. A 438-440
(2006) 18–24.
corresponds to u. (e) Schematic of the dissociated partial dislocations.
[2] H.Y. Kim, S. Miyazaki, Martensitic transformation and superelastic properties
of Ti-Nb base alloys, Mater. Trans. 56 (5) (2015) 625–634.
[3] H.Y. Kim, Y. Ikehara, J.I. Kim, H. Hosoda, S. Miyazaki, Martensitic transforma-
Using the results obtained in this study, the dissociation reac- tion, shape memory effect and superelasticity of Ti–Nb binary alloys, Acta
tion of [101]o dislocation can be written as follows: Mater. 54 (9) (2006) 2419–2429.
[4] T. Furuhara, S. Annaka, Y. Tomio, T. Maki, Superelasticity in Ti–10V–2Fe–3Al
[101]o → 1/2[101]o + APSB + 1/2[101]o (2) alloys with nitrogen addition, Mater. Sci. Eng. A 438-440 (2006) 825–829.
[5] H.Y. Kim, S. Hashimoto, J.I. Kim, T. Inamura, H. Hosoda, S. Miyazaki, Effect of
This dislocation dissociation expectedly corresponds to model Ta addition on shape memory behavior of Ti–22Nb alloy, Mater. Sci. Eng. A 417
(1–2) (2006) 120–128.
(i) (Figs. 2a and 2c). Using first-principles calculations, |b1 | was [6] Y.L. Hao, S.J. Li, S.Y. Sun, C.Y. Zheng, R. Yang, Elastic deformation behaviour
found to be slightly smaller than |b2 |; however, the difference in of Ti–24Nb–4Zr–7.9Sn for biomedical applications, Acta Biomater. 3 (2) (2007)
magnitudes of the Burgers vectors of partial dislocations could not 277–286.
[7] M. Niinomi, Metallic biomaterials, J. Artif. Organs 11 (3) (2008) 105.
be obtained via TEM. The observed partial dislocations were noted [8] R.J. Talling, R.J. Dashwood, M. Jackson, D. Dye, On the mechanism of supere-
to be screw dislocations, and freely dissociated into the plane be- lasticity in Gum metal, Acta Mater. 57 (4) (2009) 1188–1198.
longing to the [101]o zone, i.e., the partial dislocations are not dis- [9] M. Tahara, H.Y. Kim, H. Hosoda, T.H. Nam, S. Miyazaki, Effect of nitrogen ad-
dition and annealing temperature on superelastic properties of Ti–Nb–Zr–Ta
sociated on a fixed plane (Fig. 9e). However, these twisted par-
alloys, Mater. Sci. Eng. A 527 (26) (2010) 6844–6852.
tial dislocations and APSB could not be altered by external stress. [10] M. Tahara, H.Y. Kim, T. Inamura, H. Hosoda, S. Miyazaki, Lattice modulation
When the [101]o dislocation operated as a dislocation slip, the two and superelasticity in oxygen-added beta-Ti alloys, Acta Mater. 59 (16) (2011)
partial dislocations lied on the slip plane. We assume that the lo- 6208–6218.
[11] A. Biesiekierski, J. Wang, M. Abdel-Hady Gepreel, C. Wen, A new look
cal twist of partial dislocations is caused by the trapping of the at biomedical Ti-based shape memory alloys, Acta Biomater. 8 (5) (2012)
partial dislocations in the pre-existing grown-in type APSB (ribbon- 1661–1669.
like planar fault). The interaction between the grown-in type APSB [12] M. Tahara, H.Y. Kim, T. Inamura, H. Hosoda, S. Miyazaki, Role of interstitial
atoms in the microstructure and non-linear elastic deformation behavior of
and dislocation dissociation-type APSB plays an important role in Ti–Nb alloy, J. Alloys Compd. 577 (0) (2013) S404–S407.
further understanding of [101]o dislocations. Therefore, studies re- [13] J.G. Niu, W.T. Geng, Oxygen-induced lattice distortion in β -Ti3Nb and its sup-
lated to this are currently in progress. pression effect on β to α  transformation, Acta Mater. 81 (0) (2014) 194–203.
[14] M. Tahara, T. Inamura, H.Y. Kim, S. Miyazaki, H. Hosoda, Role of oxygen atoms
in α  martensite of Ti-20 at.% Nb alloy, Scr. Mater. 112 (2016) 15–18.
5. Conclusion [15] T. Inamura, J.I. Kim, H.Y. Kim, H. Hosoda, K. Wakashima, S. Miyazaki, Compo-
sition dependent crystallography of α  -martensite in Ti–Nb-based β -titanium
The dissociation process of <101>o dislocation in the α 
alloy, Philos. Mag. 87 (23) (2007) 3325.
[16] Y.W. Chai, H.Y. Kim, H. Hosoda, S. Miyazaki, Interfacial defects in Ti–Nb shape
martensite phase of a Ti–Nb shape memory alloy was systemat- memory alloys, Acta Mater. 56 (13) (2008) 3088–3097.
ically investigated by first-principles calculations and TEM obser- [17] Y.W. Chai, H.Y. Kim, H. Hosoda, S. Miyazaki, Self-accommodation in Ti–Nb
shape memory alloys, Acta Mater. 57 (14) (2009) 4054–4064.
vation. The [101]o dislocation dissociated into two partial disloca-
[18] T. Inamura, H. Hosoda, H.Young Kim, S. Miyazaki, Antiphase boundary-like
tions. The Burgers vectors of both partial dislocations were equal to stacking fault in α  -martensite of disordered crystal structure in β -titanium
1/2[101]o . A planar fault with an anti-phase shuffling of the basal shape memory alloy, Philos. Mag. 90 (25) (2010) 3475–3498.
[19] T. Inamura, H. Hosoda, S. Miyazaki, Incompatibility and preferred morphology
plane was observed between the dissociated partial dislocations.
in the self-accommodation microstructure of β -titanium shape memory alloy,
The dissociation reaction of [101]o dislocation was summarized as Philos. Mag. 93 (6) (2013) 618–634.
[101]o → 1/2[101]o + anti-phase shuffling boundary + 1/2[101]o . [20] T. Inamura, M. Ii, M. Tahara, H. Hosoda, Formation process of the incompatible
Most of the observed dislocations were noted to be screw disloca- martensite microstructure in a beta-titanium shape memory alloy, Acta Mater.
124 (2017) 351–359.
tions, and the dissociated partial dislocations were twisted around [21] H. Tobe, Deformation Mechanism and Cold Rolling Textures in β -Ti alloys, Uni-
the dislocation line. versity of Tsukuba, Tsukuba, Japan, 2012.

7
M. Tahara, N. Otaki, D. Minami et al. Acta Materialia 227 (2022) 117705

[22] H. Tobe, H.Y. Kim, T. Inamura, H. Hosoda, S. Miyazaki, Origin of {3 [32] J.P. Perdew, J.A. Chevary, S.H. Vosko, K.A. Jackson, M.R. Pederson, D.J. Singh,
3 2} twinning in metastable β -Ti alloys, Acta Mater. 64 (2014) 345– C. Fiolhais, Atoms, molecules, solids, and surfaces: Applications of the general-
355. ized gradient approximation for exchange and correlation, Phys. Rev. B 46 (11)
[23] E. Bertrand, P. Castany, Y. Yang, E. Menou, T. Gloriant, Deformation twinning (1992) 6671–6687.
in the full-α  martensitic Ti–25Ta–20Nb shape memory alloy, Acta Mater. 105 [33] D. Vanderbilt, Soft self-consistent pseudopotentials in a generalized eigenvalue
(2016) 94–103. formalism, Phys. Rev. B 41 (11) (1990) 7892–7895.
[24] P. Castany, Y. Yang, E. Bertrand, T. Gloriant, Reversion of a parent {130}{310}α  [34] A.R.G. Brown, D. Clark, J. Eastabrook, K.S. Jepson, The Titanium–Niobium sys-
martensitic twinning system at the origin of {332}{113}β twins observed in tem, Nature 201 (1964) 914–915.
metastable β titanium alloys, Phys. Rev. Lett. 117 (24) (2016) 245501. [35] S. Banumathy, R.K. Mandal, A.K. Singh, Structure of orthorhombic martensitic
[25] M. Tahara, N. Okano, T. Inamura, H. Hosoda, Plastic deformation behaviour phase in binary Ti–Nb alloys, J. Appl. Phys. 106 (9) (2009) 093518–093518-6.
of single-crystalline martensite of Ti-Nb shape memory alloy, Sci. Rep. 7 (1) [36] M. Bonisch, M. Calin, L. Giebeler, A. Helth, A. Gebert, W. Skrotzki, J. Eckert,
(2017) 15715. Composition-dependent magnitude of atomic shuffles in Ti-Nb martensites, J.
[26] S. Hanada, M. Ozeki, O. Izumi, Deformation characteristics in β phase Ti-Nb Appl. Crystallogr. 47 (4) (2014) 1374–1379.
alloys, Metall. Trans. A 16 (5) (1985) 789–795. [37] R. Davis, H.M. Flower, D.R.F. West, Martensitic transformations in Ti-Mo alloys,
[27] K. Hagihara, T. Nakano, Experimental clarification of the cyclic deformation J. Mater. Sci. 14 (3) (1979) 712–722.
mechanisms of β -type Ti–Nb–Ta–Zr-alloy single crystals developed for the sin- [38] J.S. Lee Pak, C.Y. Lei, C.M. Wayman, Atomic ordering in Ti-V-Al shape memory
gle-crystalline implant, Int. J. Plast. 98 (Supplement C) (2017) 27–44. alloys, Mater. Sci. Eng. A 132 (Supplement C) (1991) 237–244.
[28] K. Hagihara, T. Nakano, H. Maki, Y. Umakoshi, M. Niinomi, Isotropic plasticity of [39] T. Ahmed, H.J. Rack, Martensitic transformations in Ti-(16–26 at%) Nb alloys, J.
β -type Ti-29Nb-13Ta-4.6Zr alloy single crystals for the development of single Mater. Sci. 31 (16) (1996) 4267–4276.
crystalline β -Ti implants, Sci. Rep. 6 (2016) 29779. [40] D. Banerjee, K. Muraleedharan, J.L. Strudel, Substructure in titanium alloy
[29] Y. Kamimura, S. Katakura, K. Edagawa, S. Takeuchi, S. Kuramoto, T. Furuta, Ba- martensite, Philos. Mag. A 77 (2) (1998) 299–323.
sic deformation mechanism of Bcc titanium-based alloy of gum metal, Mater. [41] M. Matsuda, K. Kuramoto, Y. Morizono, S. Tsurekawa, E. Okunishi, T. Hara,
Trans. 57 (9) (2016) 1526–1534. M. Nishida, Transmission electron microscopy of antiphase boundary-like
[30] S.-.H. Lee, K. Hagihara, T. Nakano, Microstructural and orientation dependence structure of B19 martensite in Ti–Ni shape memory alloy, Acta Mater. 59 (1)
of the plastic deformation behavior in β -type Ti-15Mo-5Zr-3Al alloy single (2011) 133–140.
crystals, Metall. Mater. Trans. A 43 (5) (2012) 1588–1597. [42] M. Matsuda, T. Hara, M. Nishida, Crystallography and morphology of antiphase
[31] W. Kohn, L.J. Sham, Self-consistent equations including exchange and correla- boundary-like structure induced by martensitic transformation in Ti-Pd shape
tion effects, Phys. Rev. 140 (4A) (1965) A1133–A1138. memory alloy, Mater. Trans. 49 (3) (2008) 461–465.

You might also like