Next Article in Journal
A Graphical Tool to Describe the Operating Point of Direct Reduction Shaft Processes
Next Article in Special Issue
Physical Factors Controlling Large Shape Memory Effect in FCC ↔ HCP Martensitic Transformation in CrMnFeCoNi High-Entropy-Alloy Single Crystals
Previous Article in Journal
Study on the Deformation and Fracture Mechanisms of Plastic Metals Considering Void Damage
Previous Article in Special Issue
Hydrogen’s Effect on the Shape Memory Effect of TiNi Alloy Single Crystals
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Orientation Dependence of B19’-Martensite Reorientation Stress and Yield Stress in TiNi Single Crystals

by
Elena Y. Panchenko
,
Anna S. Eftifeeva
*,
Ilya D. Fatkullin
,
Anton I. Tagiltsev
,
Nikita Y. Surikov
,
Maria V. Zherdeva
,
Ekaterina E. Timofeeva
and
Yuriy I. Chumlyakov
Laboratory for Physics of High-Strength Crystals, Siberian Physical-Technical Institute, Tomsk State University, Lenina Str. 36, 634050 Tomsk, Russia
*
Author to whom correspondence should be addressed.
Metals 2023, 13(9), 1567; https://doi.org/10.3390/met13091567
Submission received: 11 August 2023 / Revised: 29 August 2023 / Accepted: 4 September 2023 / Published: 6 September 2023
(This article belongs to the Special Issue Research Progress of Metal Smart Materials)

Abstract

:
This paper deals with the effect of crystal orientation on the B19’-martensite reorientation stress and yield stress in compression in TiNi single crystals with different Ni contents varying from 50.4 to 51.2 at.%. It was experimentally shown that the martensite yield stress appears to be higher for the [111]B2-oriented single crystals than for the [001]B2-oriented single crystals regardless of Ni content. The difference between martensite yield stress for the two investigated orientations increases with the growth of Ni content. The maximum difference between martensite yield stress σcrM for two investigated orientations is 996 MPa at Ni content of 51.2 at.% (σcrM = 1023 MPa for the [001]B2-orientation and σcrM = 2019 MPa for the [111]B2-orientation). As a result of comparison with the B2-austenite yield stress, it was found that this is not an ordinary case. The [001]B2 orientation is a high-strength in B2-austenite and a low-strength in B19’-martensite. It was experimentally shown that the B19’-martensite reorientation stresses weakly depend on the orientation and chemical composition compared with the martensite yield stress. The reasons for the orientation dependence of the martensite yield stress in compression and the deformation mechanisms of B19’-martensite are discussed.

1. Introduction

The TiNi alloy is the headliner among shape memory alloys (SMAs) in terms of knowledge. This determines the use of this alloy in medical practice, space, mining industry, robotics, the production of thermal mechanical devices, and the development of special technologies. Binary quenched TiNi alloys with a nickel content of CNi < 51.2 at.% undergoes thermoelastic martensitic transformation (MT) from the cubic B2 austenite lattice to the monoclinic B19’ martensite lattice and back in stress-free cooling/heating [1,2]. In quenched high-nickel alloys with CNi ≥ 51.2 at.% there is a strain-glass transition during stress-free cooling/heating, while B2-B19’ MT occurs only under applied stress [3,4].
Most of the research works on the TiNi alloy are related to the study of the critical stress levels for reorientation and detwinning of multivariant the B19’-martensite structure σreor (T < Ms), the critical stress levels for the B19’-martensite formation σcr in the stress-induced MT temperature range (Md > T > Af) and the levels of B2-austenite yield stress σcrA (T > Md) (Figure 1). This is understandable because these stress levels σreor, σcr, and σcrA are responsible for the temperature interval and characteristics of the stress-induced B2-B19’ MT and related phenomena, such as shape memory effect (SME) and superelasticity (SE) [1,2,5,6,7,8,9].
It is well known that at Md temperature, the critical stress for the stress-induced martensite formation and B2-austenite yield stress are equal to each other [2,5,8,9,10]. Therefore, MT cannot be stress-induced above Md temperature, and this temperature determines the criterion and temperature range for SE (Figure 1). The Md temperature strongly depends on the microstructure, chemical composition, and orientation of single crystals or texture of polycrystals. It varies from 323 to 423 K [11,12]. The higher level of B2-austenite yield stresses and the greater the ratio of σcr(Md)/σcr(Ms) > 5÷7 (the Wayman–Otsuka condition [8]), the more probable the manifestation of SE in the SMAs. One of the ways to obtain a wide temperature range of SE with a high level of σcrA is to minimize the critical stress for the stress-induced martensite formation σcr and, accordingly, to minimize the coefficient α = dσ/dT (Figure 1) [9]. The orientation dependence of the B2-phase yield stress is well studied in both single- and hetero-phase TiNi single crystals [11,12,13,14]. The Schmid factors analysis for the a <001> B2{110}B2 slip system in the B2-phase, experimental data of the temperature dependence of the yield stress and deformation mechanisms of the B2-phase for the single crystals of TiNi alloys show that the <001>B2 orientation is a high-strength and is characterized by high yield stress ~G/100 (G is the shear modulus of austenite). The main mechanism of the B2-phase plastic deformation for this orientation is twinning [12,13,14,15,16,17,18]. For example, the B2-phase yield stress of σcrA = 1300–1800 MPa for the [001]B2 orientation is 1.6 times higher and the SE interval of ~200–250 K is wider than for the [111]B2 orientation (σcrA = 800–1100 MPa and SE interval of <150 K) in quenched Ti-(50.7–51.8 at.%)Ni single crystals in compression [9,11,13].
The SMAs must have the high-strength properties of both the austenite and martensite to show a wide temperature range of SE, a high reversible strain of the MT, as well as good cyclic stability of the SE. The martensite yield stress σcrM characterizes the ability of the alloy to resist plastic deformation of martensite. The mechanism of B19’-martensite plastic deformation has been studied in the TiNi alloy. Numerous experimental studies mainly on polycrystalline equiatomic TiNi and Ti-(50.5–50.9 at.%)Ni bars and wires report the activity of the and <001>{100}B19’ dislocation slip in B19’-martensite in combination with the twinning [2,10,19,20,21,22,23,24,25]. It should be noted that the thermo-induced martensite, existing without stress below Mf temperature, consists of 24 different crystallographically equivalent internally twinned variants. The two twin-related variants constitute the correspondent variants pair (CVP). The CVPs form a self-accommodating structure of martensite and define an invariant habit plane {0.89 0.40 0.22} [7,26,27,28,29,30,31]. Type II <011>B19’ and the type I { 1 ¯ 1 ¯ 1}B19’ are the preferred twinning modes required to obtain the lattice invariant strain and the habit plane at the B2-B19’ MT in the quenched TiNi alloys [7]. Type II <011>B19’ twinning mode is described in terms of the twinning elements as follows K1 = (0.7205 1 1 ¯ ), η1 = [011], K2 = (011), η2 = [1.5727 1 1 ¯ ]. Type I { 1 ¯ 1 ¯ 1}B19’ is described as follows K1 = ( 1 ¯ 11), η1 = [0.5404 0.49597 1], K2 = (0.2469 0.561 1), η2 = [ 2 ¯ 1 ¯ 1]. Another twinning mode is the compound {001}B19’ twins (K1 = (001), η1 = [100], K2= (100), η2 = [001]), which are not a solution for the phenomenological crystallographic theory of the MT and can exist both in the single- and hetero-phase TiNi alloys in the stress concentration places [7,27]. Such twins grow during thermo-induced MT, do not lead to macrodeformation of the sample, and are not inherited to austenite when heated above Af during the reverse MT. If the B19’-martensite yield stress is reached, then the plastic deformation, in addition to the dislocation slip, can proceed via the twinning of types {100}B19’, {20 1 ¯ }B19’, or {113}B19’ in the martensite B19’-phase. After unloading and heating above Af, these twins can be inherited to the austenite phase. The {114}B2 twins of B2-austenite originates from the {20 1 ¯ }B19’ twins of the B19’-martensite, the {211}B2 twins of B2-austenite originates from the {113}B19’ twins of the B19’-martensite [2,7,9,10,23].
The B19’-martensite yield stress σcrM, as well as the stresses of σcr, σreor, and σcrA, depends on the deformation way, chemical composition, and single crystal orientation or polycrystal texture. On Ti-(50.5 at.%)Ni and Ti-(50.8–50.9 at.%)Ni wires in tension, it is shown [19,32,33] that the martensite yield stress σcrM at T < Ms depends on the Ni concentration (σcrM = 1150 MPa for Ti49.5Ni50.5 wire, σcrM = 1420 MPa for Ti49.2Ni50.8 wire, σcrM = 1000 MPa for Ti49.1Ni50.9 wire) and exceeds the B2-phase yield stress at T~MdcrA = 800 MPa for Ti49.5Ni50.5 wire, σcrA = 1230 MPa for Ti49.2Ni50.8 wire, σcrA = 830 MPa for Ti49.1Ni50.9 wire). With an increase in the test temperature to T = Md, the B19’-martensite yield stress σcrM decreases and becomes close to the B2-austenite yield stress at T~Md [19,32,33]. Šittner et al. have found that the oriented martensite plastic deformation can be either homogeneous or localized, depending on the yield stress and strain hardening rate [10,19,33]. The martensite plastic deformation proceeds via a peculiar deformation mechanism which combines {100}B19’ deformation twinning with <100>B19 dislocation slip based on kinking in nanocrystalline Ti49.1Ni50.9 and Ti49.5Ni50.5 wires. The reverse MT after the martensite plastic deformation during unloading and heating leaves behind a large irreversible strain and a high density of lattice defects in austenite. The martensite transforms into the B2-austenite and the {20 1 ¯ }B19’ deformation twins transform to {114}B2 austenite twins [10,19,33].
It was shown that the B19’-martensite yield stress in the Ti49.2Ni50.8 wire, Ti48.5Ni51.5 and Ti49.2Ni50.8 single crystals with <112>B2, <148>B2, <111>B2, and <100>B2 orientations increases due to the precipitation hardening after aging at T = 573–723 K compared with aging at T = 773–823 K [12]. For example, for single crystals oriented along the <112>B2, <148>B2, <111>B2, and <100>B2-directions after aging at 723 K the martensite yield stress in compression at T = Troom > Af is σcrM = 2000–2100 MPa, and after aging at 823 K σcrM decreases and becomes equal to 1500–1700 MPa. At the same time, the yield stress σcrM does not depend on the orientation in the aged at 723–823 K Ti48.5Ni51.5 and Ti49.2Ni50.8 single crystals with dispersed Ti3Ni4 particles.
Thus, currently, there are disorganized data on the study of the B19’-martensite yield stress on TiNi polycrystals and single crystals for the separate orientations/texture, composition, test temperatures and thermomechanical treatments. The effect of test temperature and deformation way on the B19’-martensite yield stress is considered only for polycrystalline alloys in tension, such as equiatomic TiNi, Ti49.5Ni50.5 and Ti-(50.8–50.9 at.%)Ni bars and wires [19,21,22,32,33]. The effect of orientation on the B19’-martensite yield stress in compression has been mainly considered only on heterophase Ti49.5Ni51.5 and Ti49.2Ni50.8 single-crystals and only at room temperature in SE conditions at T > Af [12,13]. The scientific and practical interest is the systematic study of the orientation dependence of the B19’-martensite yield stress in the single-phase state without the influence of the secondary phases in TiNi single crystals with various nickel contents and the refinement of the plastic deformation mechanisms of B19’-martensite. The purpose of this work is to investigate the orientation dependence of the critical stress for the stress-induced B19’-martensite reorientation and the B19’-martensite yield stress in the quenched TiNi single crystals with various Ni content.

2. Materials and Methods

The single crystals of Ti49.6Ni50.4 (at.%) (marked as 50.4Ni), Ti49.3Ni50.7 (at.%) (marked as 50.7Ni), Ti48.8Ni51.2 (at.%) (marked as 51.2Ni) alloys were grown by the Bridgman technique. The compressive test specimens had [001]B2 ([001]B2-oriented crystals) and [111]B2 ([111]B2-oriented crystals) deformation axes. Samples had a cuboid shape with a cross-sectional area of 9–25 mm2 and a length of 6–10 mm. The orientation of the samples was determined by X-ray diffraction analysis on a Dron-3 diffractometer (Burevesnik, St. Petersburg, Russia). The cut samples were annealed at 1253 K for 1.0 h in a helium atmosphere and quenched in room-temperature water.
The microstructure of single crystals was investigated on a transmission electron microscope (TEM) HT-7700 (Hitachi, Tokyo, Japan). The foils for TEM were disks with a diameter of 3 mm and a thickness of 100 μm. They were finally thinned out using double-jet electrochemical polishing by a Tenupol-5 (Struers, Ballerup, Denmark).
The MT temperatures were determined by differential scanning calorimetry (DSC) using a DSC 404F1 Pegasus (NETZSCH, Bayern, Germany). The cooling and heating rates of DSC analysis were 10 K/min. The loading/unloading cycles were performed using a VHS 5969 universal testing machine (Instron, High Wycombe, UK). Errors in strain and stress measurements during mechanical tests were 0.2% and 2 MPa, respectively. The strain rate during the study of SE response was 2.0·10−3 s−1. The strain-glass transformation was investigated by a DMA/SDTA861 dynamic mechanical analyzer (DMA) (Mettler Toledo, Columbus, OH, USA).

3. Results

DSC curves are shown in Figure 2, and B2-B19’ MT temperatures are summarized in Table 1 for the quenched single crystals with different Ni content. An increase in Ni concentration from 50.4 to 51.2 at.% leads to a decrease in the MT temperatures and an extension of the forward and reverse MT intervals Δ1 = Ms − Mf and Δ2 = Af − As. The MT intervals are Δ1 = 14 K and Δ2 = 17 K at a Ni concentration of 50.4 at.%. With an increase in Ni concentration by 0.3 at.% the MT intervals expand by 36 K for Δ1 and by 31 K for Δ2, and shift towards low temperature. With a further increase in Ni concentration, the endothermic and exothermic peaks are not observed in the DSC curves, which indicates the absence of the B2-B19’ MT during cooling/heating in 51.2Ni single crystals. Instead, a strain-glass transition occurs, as the DMA study shows (Figure 3).
The frequency-temperature dependencies of the elastic modulus E and defining the internal friction tan δ at cooling for [001]B2-oriented quenched 51.2Ni single crystals are shown in Figure 3. The dip in the temperature dependence of the elastic modulus E(T) in the region of the internal friction peak (tan δ) indicates a change in the microstructure of the material or a phase transition (Figure 3). As the frequency increases, the internal friction peak reduces, and the temperature of the internal friction peak Tg shifts towards high temperatures. The temperature Tg depends on the frequency according to the Vogel–Fulcher relation [3]:
ω = ω 0 e x p E a k B T g T 0
where T0 = 180 K is the temperature Tg at a frequency of ω→0 Hz, which is consistent with [4], Ea is the activation energy, kB is the Boltzman constant. This frequency dependence of internal friction is the main characteristic of the strain-glass transition, which is observed in 51.2Ni single crystals (Figure 3). The DMA curves show no shift of the internal friction peak on frequency during the MT [3].
The stress-strain σ(ε) curves under uniaxial compression for the quenched TiNi single crystals are demonstrated in Figure 4 and Figure 5. The σ(ε) curves for the 50.4Ni and 50.7Ni single crystals were obtained at temperatures T = 258 K and T = 203 K, respectively, i.e., the samples without stress were below temperature Ms in the martensite phase. The lowest possible temperature T = 203 K was chosen for the stress-induced MT in the high-nickel single crystals of 51.2Ni alloys not undergoing the MT upon stress-free cooling. Two yield stress σcrreor) < σcrM and four deformation stages can be distinguished on the σ(ε) curves. Stage I is an elastic deformation of the initial structure under stress. The initial structure of the 50.4Ni and 50.7Ni single crystals is the self-accommodating structure of the thermal-induced B19’-martensite. The initial structure of 51.2Ni single crystals is the austenite B2-phase. The σreor stress corresponds to the critical stresses for the B19’-martensite reorientation in 50.4Ni and 50.7Ni single crystals and is weakly dependent on the orientation and Ni content (Figure 4). The critical stress is σreor = 169–192 MPa for 50.4Ni single crystals and σreor = 138–153 MPa for 50.7Ni single crystals. In the high-nickel 51.2Ni single crystals undergoing a strain-glass transition upon cooling, the B2-B19’ MT can be induced under the action of uniaxial stresses [3,4]. In the 51.2Ni crystals, the value of σcr characterizes the beginning of the stress-induced B19’-martensite formation and significantly exceeds stress for the martensite variants reorientation in the single crystals with CNi = 50.4–50.7 at.%. In addition, a strong orientation dependence of σcr is observed in the 51.2Ni single crystals: σcr = 876 MPa for the [111]B2 orientation is 2.1 times higher than σcr = 412 MPa for the [001]B2 orientation (Figure 5).
Stage II length depends on the composition and orientation of single crystals. In the case of the [001]B2 orientation, stage II is observed up to a given strain of 6.0–7.0% in 50.4Ni and 50.7Ni single crystals and up to 9.5% in 51.2Ni single crystals. Stage II ends at a given strain of 6.0% in 50.4Ni single crystals, 3.5% in 50.7Ni single crystals, and 7.5% in 51.2Ni single crystals with the [111]B2 orientation. The plastic deformation occurs in stage II with a low strain hardening coefficient θ = dσ/dε because of the rearrangement of B19’-martensite twin structure by twin boundaries movement. As a result, the preferred oriented martensite variants by the action of the external load grow, and their detwinning occurs in the 50.4Ni and 50.7Ni single crystals. In the high-nickel 51.2Ni single crystals only the oriented B19’-martensite is formed at stage II due to the deformation process (martensite reorientation and/or detwinning) responsible for the stress-induced MT. The single crystal orientation defines the stage length and strain hardening coefficient θ. The [001]B2-oriented crystals are characterized by low coefficient θ and high length of stage II compared with the [111]B2-oriented single crystals independently of Ni content (Figure 4 and Figure 5).
In the case of the [001]B2 orientation, stage III is up to a given strain of 10.5% in 50.4Ni and 50.7Ni single crystals and 13.0% in 51.2Ni single crystals. Stage III for the [111]B2 orientation is observed up to a given strain of 12.0% in 50.4Ni single crystals, 9.0% in 50.7Ni single crystals, and 11.5% in 51.2Ni single crystals. Stage III is accompanied by a higher coefficient θ and is usually associated with the elastic deformation of the formed oriented B19’-martensite. In addition, the further oriented B19’-martensite reorientation and the detwinning, MT undergoing, as well as the formation of new twin modes, are possible at this stage. Such a combination of various deformation processes leads to the different values of θ = dσ/dε at stage III depending on the crystal orientation and the Ni content and does not explicitly characterize the elastic modulus of oriented B19’-martensite (Figure 4 and Figure 5).
The samples were unloaded at the given strain of εg < 10% when the stress was lower than σcrM. The loading/unloading curves are shown in the insets of Figure 4 and Figure 5. The SME occurs in the low-nickel 50.4Ni and 50.7Ni single crystals. The complete shape and size recovery of these samples is observed after unloading and heating to T > Af due to the reverse MT. In the high-nickel 51.2Ni single crystals, the given strain is completely reversible without heating due to SE effect. The [001]B2-oriented crystals with long stages II are characterized by higher reversible strain εrev compared with the [111]B2-oriented crystals regardless of the Ni content.
The B19’-martensite plastic deformation occurs above σcrM at stage IV. The σcrM stress is the yield stress of the oriented B19’-martensite. It has been experimentally shown that there is a strong dependence of the martensite yield stress σcrM on the Ni content and the single crystal orientation in the studied TiNi single crystals (Figure 4 and Figure 5). The martensite yield stress σcrM for the [111]B2-oriented crystals is higher than for the [001]B2-oriented crystals. The difference ΔσcrM = σcrM[111] − σcrM{001] at one Ni concentration shows how much the martensite yield stress for the [111]B2-oriented crystals is higher than in the [001]B2-oriented crystals. The value of ΔσcrM increases from 272 MPa to 996 MPa with rising the Ni concentration from 50.4 to 51.2 at.%. The martensite yield stress σcrM = 2019 MPa for 51.2Ni single crystals with the [111]B2 orientation is two times higher than the one for single crystals with the [001]B2 orientation, where it is σcrM = 1023 MPa.
The microstructure after plastic deformation of B19’-martensite up to the given strain εg = 12.0–15.5% was studied in detail in the [001]B2- and [111]B2-oriented single crystals of 50.4Ni and 50.7Ni alloys by the TEM. It should be noted that the study of the microstructure after deformation in the martensite was carried out after the unloading, heating to room temperature, and subsequent preparation of thin foils from deformed samples. In the process of heating and preparation of thin foils, the reverse MT was in the deformed structure of the samples. Then, the thermo-induced martensite may appear with a decrease in temperature of the foil. Therefore, the microstructure observed by TEM differs from the one after plastic deformation of oriented B19’-martensite up to εg = 12.0–15.5%. It is very difficult to identify the stages at which one or another defective structure of austenite or martensite was formed, and what processes preceded this.
It has been experimentally shown that the B2-phase areas with a high dislocation density are observed after deformation above the B19’-martensite yield stress in compression and subsequent heating, regardless of the crystal orientations (Figure 6). As the samples were deformed plastically in B19’-martensite, the dislocation slip occurred in B19’-martensite, and this dislocation structure was inherited by the B2-phase from B19’-martensite during reverse MT. The elongated dislocations were observed along the (110)B2, (101)B2, and (01 1 ¯ )B2 (Figure 6b). No large difference could be distinguished in the dislocation density between [001]B2 and [111]B2 crystal orientations (Figure 6).
The multiple deformation twinning of B2-phase and lamellas of the residual martensitic phase are detected by TEM in [001]B2-oriented 50.4Ni single crystals (Figure 7). The azimuthal smearing of reflections in the B2-matrix and B19’-martensite in [001]B2-oriented single crystals indicates the presence of a misoriented substructure caused by the intense action of dislocation deformation mechanisms. Misorientation of the structure is formed as a result of the forward and reverse stress-induced B2-B19’ MT with the retention of the boundaries of the martensite twin structure. The twin boundaries are retained in the B2-austenite structure after the reverse MT because they already interacted with dislocation slip in the matrix. The material of the former martensite twins has become slightly misoriented [34].
A more detailed analysis of the microstructure of the [001]B2-oriented single crystals showed that the {411}B2 twins of the B2-phase are observed after plastic deformation of B19’-martensite and subsequent heating to room temperature (Figure 8). Figure 8 shows dark field images in the matrix and twin reflections of the {411}B2 twins up to ~100 nm wide.
In [111]B2-oriented single crystals the morphology of deformation structure differs from the morphology of the [001]B2-oriented single crystals after the plastic deformation above the B19’-martensite yield stress to εg = 12.0–15.5% with followed by heating to room temperature. A V-shaped microstructure is formed (Figure 9), which was repeatedly observed in the deformed thin TiNi wire with the [111]B2 texture [28]. The microstructure of plastically deformed TiNi wire consists of a high density of the {114}B2 austenite twins that often form wedges-like twins inside strongly dislocated austenite phase [28].
The analysis of structure shows that there is a mix of B2- and B19’-phases in the [111]B2-oriented 50.4Ni single crystals deformed in martensite and heated to room temperature. The bands of B2-austenite (bright bands in the bright field image in Figure 10a) and of B19’-martensite (dark bands in the bright field image in Figure 10a) with widths of up to ~100 nm are detected by TEM. Moreover, the SADPs from martensite variants c and d are rotated by 90° around their common [100]B19’ zone axis.
A similar V-shaped microstructure is observed in [111]B2-oriented 50.7Ni single crystals (Figure 11). Only B2-phase reflections are present in the joint SADP from dark and light bands (Figure 11b). The two zones are simultaneously in the reflective position: the [011]m matrix zone and the [111]tw twin zone. The misorientation angle between these types of zones in a cubic lattice is 54.8°. The reorientation vector of the crystal lattice at twinning is the <110>B2-direction. The orientation relation {011}m||{111}tw definitely characterizes the {114}B2 twinning in the B2-phase [35]. It can be seen from the joint SADP that the structure shown in Figure 11a,b is fragmented by the twinning in the two systems with {114}B2 planes in the B2 phase. The two systems of twins are misoriented by 60° around the {011}m||{111}tw zone axis. The width of {114}B2 twins is ~100–200 nm.
Thus, the main defect structure elements after plastic deformation above the B19’-martensite yield stress up to εg = 12.0–15.5% in compression and subsequent heating are the dislocation structure and {114}B2 twins of the B2-phase regardless of the TiNi single crystal orientation.

4. Discussion

It has been experimentally shown that the critical stresses for the stress-induced B19’-martensite reorientation σreor during the martensite deformation in compression (T < Mf) are practically independent of the chemical composition and orientation as opposed to the martensite yield stress (Figure 4). The critical stress for the stress-induced martensite reorientation is σreor = 169–192 MPa in 50.4Ni single crystals and σreor = 138–153 MPa in 50.7Ni single crystals (Figure 4). This is primarily determined by the self-accommodating structure formation, which consists of 24 CVPs (Table 2). Moreover, these CVPs are not randomly distributed in the matrix. They form the combination of a triangular or hexagonal shape that minimizes the elastic strain energy during transformations [7,26,27,28,29,30,31]. There are 12 lattice correspondences between the B2- and B19’-phase (Table 2). The lattice invariant strain and the habit plane formation are achieved by the CVP formation, which consist of two different martensite domains in single-phase TiNi crystals. The relationship between the martensite domain pairs exists as type II-1 <011>B19’ twin (Table 2 and Table 3) [7], while the relationship between different CVPs appears as type I { 1 ¯ 1 ¯ 1}B19’ twin [31]. Such a twinned structure is isotropic and is characterized by low critical stresses for the twin boundaries motion.
The σcr value for high-nickel 51.2Ni single crystals significantly exceeds the σcr values for 50.4Ni and 50.7Ni single crystals. The σcr value for high-nickel 51.2Ni single crystals characterizes the stress for the start of stress-induced martensite formation and strongly depends on the crystal orientation. The σcr = 876 MPa for [111]B2-oriented crystals is two times higher than σcr = 412 MPa for [001]B2-oriented crystals (Figure 5). The strain-glass transition is observed in the cooling/heating cycles, and the self-accommodating structure of thermo-induced martensite is not formed in TiNi single crystals with a nickel content CNi ≥ 51.2 at.%. As shown in Ref. [3], there are numerous point defects in the matrix at a high nickel content that suppress the MT. The defects form local distortions that frustrate the long-range order of the crystal and, therefore, prevent the appearance of macroscopic martensite upon cooling. Instead, nano-sized martensite domains are formed with randomly distributed domains, within which there is a local order. It is known [11,13,36] that there is stress-induced MT in the strain glass crystals. This MT requires high critical stress σcr > 400–800 MPa depending on the orientation. According to the Clausius–Clapeyron relation [8,37], the theoretical critical driving stress for the MT σcr can be expressed by the following equation:
σ c r = σ 0 + S V m ε t r ( T M s σ 0 )
where T is the test temperature, Vm is the molar volume, εtr is the theoretical transformation strain at MT, Msσ0 is the temperature of the MT start under the action of the minimum uniaxial stress σ0 able to induce thermoelastic MT in these crystals. The theoretical transformation strain at B2-B19’ MT is εtr = 4.38% for the [001]B2 orientation and εtr = 3.58% for the [111]B2 orientation in the quenched TiNi single crystals in compression [11,12,14,15]. This primarily determines the orientation dependence of the critical stress for the B19’-martensite formation at the same test temperature σcr [111] > σcr [001] in quenched high-nickel 51.2Ni single crystals. Thus, in accordance with Equation (2), the lower transformation strain εtr = 3.58% corresponds to higher critical stress for martensite formation σcr = 876 MPa in [111]B2-oriented crystals compared with [001]B2-oriented crystals.
It has been experimentally found that there is an orientation dependence of the strain hardening coefficient θ at stages II and III, stress hysteresis Δσ at the SE, and the reversible SME/SE strain in all the studied TiNi single crystals (Figure 4 and Figure 5). This is due to the different transformation strain resources and the presence of the martensite detwinning contribution εdetw for the different oriented crystals.
The theoretical transformation strain at B2-B19’ MT, taking into account the martensite CVP formation and its subsequent martensite detwinning εCVP + detw in the single crystals of TiNi alloy, was calculated in the Refs. [11,12,14,15]. The [001]B2 orientation is characterized by a larger transformation strain εCVP + detw = 4.38% and the absence of the detwinning contribution εdetw→0 in the general strain of the MT. In contrast to the [111]B2 orientation, where the transformation strain is εCVP + detw = 3.58% and the detwinning contribution is εdetw = 0.6%. The CVP detwinning under stress leads to a deviation of the habit plane from the invariant position. This is accompanied by an increase in the friction force for the interfacial boundary motion and energy dissipation during the MT, which is characterized by stress hysteresis Δσ. The experimental reversible strain at the SME and SE for the [001]B2 orientation is higher than for the [111]B2 orientation. This corresponds to the theoretical calculations. The maximum reversible strain in 51.2Ni crystals is not achieved due to the selected test temperature and the low given strain in the loading/unloading cycle.
Thus, the [111]B2-oriented crystals are characterized by a higher strain hardening coefficient θ at stage II, a wide stress hysteresis Δσ at the SE in 51.2Ni single crystals, and a small reversible SME/SE strain, in contrast to the [001]B2-oriented crystals with the detwinning contributions of εdetw→0 (Figure 4 and Figure 5). A wide stress hysteresis Δσ also observed in ferromagnetic single crystals with B2-L10 MT, such as NiFeGa(Co), NiMnGa, and CoNi (Al, Ga) [38,39,40].
The strain hardening coefficient at stage III increases compared with stage II, but the orientation dependence of θ is retained θ[111] > θ[001]. The elastic deformation of oriented B19’-martensite proceeds at stage III in all studied single crystals. Moreover, the B19’-martensite reorientation and detwinning processes typical for stage II continue at stage III in 50.4Ni and 50.7Ni single crystals. Further MT undergoing and the B19’-martensite detwinning typical for stage II continue at stage III in 51.2Ni single crystals. This is evidenced by the reversible strain value, which does not reach its theoretical resource of transformation strain after unloading at stage II (in the insets of Figure 4 and Figure 5).
As a result of the completion of the oriented martensite formation, a different number of active martensite CVPs is formed at stage III, depending on the crystal orientation. These active CVPs have a maximum Schmid factor under the action of compressive stress. It is known [12] that there are eight active CVPs in the [001]B2-oriented crystals and six active CVPs in [111]B2-oriented crystals (Table 4). Thus, the combination of different deformation processes and different morphology of martensite formed under stress lead to different values of θ = dσ/dε at stage III depending on the crystal orientation and the Ni content and do not characterize the elastic modulus of B19’-martensite (Figure 4 and Figure 5).
The deformation behavior of single crystals from σcrM and above is associated with the plastic deformation of B19’-martensite. The deformation above σcrM does not recover even after heating to T > Af. As previously noted, there is a strong orientation dependence of the B2-phase yield stress σcrA in TiNi single crystals: [001]B2-oriented crystals are “hard” and have a higher B2-phase yield stress compared with the “soft” [111]B2-oriented crystals (σcrA[111] < σcrA[001]) in the quenched and aged states [9,12,13]. For example, in the quenched TiNi single crystals with nickel concentration of CNi = 50.0–51.8 at.% the austenite yield stress is σcrA = 1300–1800 MPa for the [001]B2 orientation. The austenite yield stress σcrA is lower by 500–700 MPa and is equal to 800–1100 MPa for the [111]B2 orientation [9,11,12,13]. In this work, it is experimentally shown for the first time that the quenched TiNi single crystals have an inverse orientation dependence of the B19’-martensite yield stress σcrM in contrast to the conventional orientation dependence of the B2-phase yield stress. The [111]B2-oriented crystals are characterized by a higher level of σcrM compared with the [001]B2-oriented crystals σcrM[111] > σcrM[001]. The difference of the B19’-martensite yield stress between two orientations ΔσcrM increases from 272 MPa to 996 MPa with an increase in Ni concentration from 50.4 to 51.2 at.% (Figure 4 and Figure 5). Moreover, an increase in the difference ΔσcrM occurs as a result of a large growth of σcrM in [111]B2-oriented crystals.
In studies by Sehitoglu et al. [12,13], the orientation dependence of the martensite yield stress in compression at temperature T = Troom > Af was not found in the aged Ti-51.5 at.% Ni and Ti-50.8 at.% Ni single crystals. These studies are concerned only with single crystals aged at 723–823 K containing dispersed Ti3Ni4 particles. In the precipitation-hardened single crystals after aging at 723 K, the martensite yield stress is σcrM = 1700–2100 MPa, regardless of the <112>B2, <148>B2, <111>B2, and <100>B2 orientations. Aged TiNi single crystals are characterized by a weakening of the orientation dependence of the critical stress for the martensite formation and the SME strain compared with quenched ones. This is determined, firstly, by the change in the morphology and twin structure of B19’-martensite. The dispersed Ti3Ni4 particles change the type of B19’-martensite twinning from type II <011>B19’ twin to the compound <001>{100}B19’ twin, and the density of the compound twins increases as the interparticle distance decreases [7,41]. Secondly, this is the multiple nucleation of martensite, generated both by elastic fields from several crystallographic variants of particles and by external applied stress. Such a multivariate martensite structure determines the difficulty of its reorientation and detwinning under the action of applied stress. As a result, the martensite morphology becomes independent of the crystal orientation and is controlled by microstructure parameters (particle size and interparticle distance).
On the basis of the TEM studies of the microstructure of quenched 50.4Ni and 50.7Ni single crystals after plastic deformation of B19’-martensite at T < Mf and subsequent heating, it was concluded that the crystals deformed in martensite contain a mix of B2- and B19’-phases (Figure 6, Figure 7, Figure 8, Figure 9, Figure 10 and Figure 11). Firstly, there is a high dislocation density in the B2-phase inherited from B19’-martensite (Figure 6). Secondly, there is a complex twinned structure consisting of B2- and B19’-phases. The main twinning system is {411}B2 twins in the B2-phase, regardless of the crystal orientation (Figure 7, Figure 8, Figure 9, Figure 10 and Figure 11).
Numerous previous studies on the TiNi polycrystals and thin wires have shown that the typical defects in the B2-phase after plastic deformation of martensite and subsequent heating are as follows [2,7,9,19,20,23]. The <100>{011}B2 dislocation slip acts as the <001>{100}B19′ slip in the corresponding B19’-martensite. The {411}B2 twins of B2-phase correspond to the {201}B19′ twins in the B19’-martensite. As shown in [42], the <001> {100}B19′ dislocation slip, the {100}B19′ and {001}B19′ twins are energetically favorable over the {201}B19′ twinning. Nevertheless, {201}B19’ twinning is one of the most common mechanical twins observed during plastic deformation of B19’-martensite. Kadeřávek et al. proposed a plastic straining mechanism of B19’-martensite, which explains the formation of {201}B19’ deformation-twin bands. This mechanism of plastic deformation is called kwinking and involves a combination of deformation twinning and dislocation-based kinking [19,33,42,43].
Since plastic deformation was realized in B19’-martensite in the studied crystals, the Schmid factors analysis m was carried out for dislocation slip systems and twinning in B19’-martensite. The Schmid factors are calculated as follows m = cosλcosχ, where λ is the angle between the normal to the habit plane and the compression axis, χ is the angle between the shear direction and the compression axis. The crystallographic indices of the deformation axis in B19’-martensite can be determined considering the lattice correspondences between the B2- and B19’-phases (Table 3) and active CVPs with the maximum Schmid factor under compressive stress along the [111]B2- and [001]B2-deformation axis [7,19,20]. The [001]B2 crystal orientation in the austenite will correspond to one of the four possible {011}B19’-directions in the B19’-martensite, while the [111]B2 orientation will correspond to one of the four {120}B19’ directions in the B19’-martensite (Table 4). Calculations of the Schmid factors for the <001>{100}B19’ slip systems in B19’-martensite and the {20 1 ¯ }B19’ twinning in B19’-martensite are presented in Table 4.
Calculations showed that the Schmid factor for the {20 1 ¯ }B19’ twinning in B19’-martensite is close for the two studied orientations. However, for the [111]B2 orientation m = 0.39–0.5 turns out to be slightly higher than for the [001]B2 orientation m = 0.37–0.39 (Table 4). The Schmid factor for the <001>{100}B19’ slip systems in the [111]B2 orientation strongly depends on the martensite orientation, while for the [001]B2 orientation, the m value is the same for all martensite orientations (Table 4). In contrast, the Schmid factor for a slip in one domain is higher than in another for one CVP in the [111]B2-oriented crystals (Figure 12). For example, the Schmid factors for slip are m = 0.27 in domain 2’, and m = 0.46 in domain 1 in active CVP 2’(+). As a result, the plastic deformation of B19’-martensite is difficult. The various Schmid factors of CVP domains can determine the high B19’-martensite yield stress in the [111]B2-oriented crystals and the structure morphology after plastic deformation in martensite.
The main difference between the microstructure of the [111]B2 and [001]B2-oriented single crystals after plastic deformation in B19’-martensite is the morphology of the twinned structure of B2- and B19’-phases: step microstructure in [001]B2-oriented crystals and V-shaped microstructure in [111]B2-oriented crystals (Figure 7, Figure 8, Figure 9, Figure 10 and Figure 11). The difference in morphology is primarily determined by the number of active CVPs, which have the maximum Schmid factor for B2-B19’ MT and determine the oriented B19’-martensite formation under compressive stress along the corresponding directions. In the case of [001]B2 orientation, the number of CVPs that are activated is 8. The number of active CVPs of [111]B2-oriented crystals is 6. It is assumed that there is a high dislocation density and/or residual martensite pinned by dislocations in one twin domain of CVP with the maximum Schmid factor m = 0.46 compared with another domain with a low Schmid factor m = 0.27 for the <001>{100}B19’ slip system in V-shaped microstructure of the [111]B2-oriented crystals after plastic deformation in martensite.
Thus, in the present work, it is shown that the orientation dependence of the B19’-martensite yield stress σcrM can have inverse regularities, in contrast to the orientation dependence of the B2-austenite yield stress in TiNi single crystals and textured polycrystals. The “hard” [001]B2-oriented single crystals in B2-austenite, which are characterized by a high yield stress level σcrA, on the contrary, are the “soft” in B19’-martensite σcrA[001] > σcrM[001]. For example, in the quenched [001]B2-oriented crystals of TiNi with CNi = 51.0–51.8 at.% the B2-austenite yield stress at T = Md is σcrA ≈ 1300–1800 MPa [9,11,13]. Whereas the B19’-martensite yield stress is below and equal to σcrM = 1023–1124 MPa for the [001]B2-oriented crystals of TiNi with CNi = 50.7–51.2 at.% (Table 4). As a result, the plastic deformation of martensite can be achieved faster than that of austenite in the [001]B2-oriented single crystals. When σcrM is reached, the plasticity sets in, and the reversible stress-induced MT is no longer possible because irreversible dislocation slip begins in B19’-martensite. Recoverable strain decreases after reaching σcrM.
In the case of the [111]B2-oriented crystals, an inverse dependence of σcrA[111] < σcrM[111] takes place. The B2-austenite yield stress is equal to σcrA≈800–1100 MPa at CNi = 51.0–51.8 at.% [9,11,13], and the B19’-martensite yield stress is equal to σcrM = 1714–2019 MPa at CNi = 50.7–51.2 at.% (Table 4). The degradation of reversible strain at stress-induced MT in this orientation will primarily occur due to the austenite plastic deformation, which is characterized by a lower yield stress compared with the B19’-martensite.
In addition, the strong orientation dependence of the B19’-martensite yield stress can be the cause of a nonhomogeneous deformation of TiNi polycrystals in the martensite. The plastic deformation of martensite begins earlier in the grains with the <001>B2 orientation than in the grains with the <111>B2 orientation with a high B19’-martensite yield stress. This is especially pronounced in high-nickel TiNi single crystals with CNi ≥ 51.2 at.% where the B19’-martensite yield stress is σcrM = 2019 MPa for the [111]B2 orientation, and it is almost 1000 MPa lower and equal to σcrM = 1023 MPa for [001]B2 orientation.
Therefore, as the dependence of the B19’-martensite yield stress σcrM on the crystal orientation established in this work, the orientation dependence of the B2-austenite yield stress σcrA must be carefully taken into account when estimating the SE temperature range, the cyclic stability of the SME and SE.

5. Conclusions

The orientation dependence of the plastic deformation of B19’-martensite in compression has been investigated on quenched TiNi single crystals with various Ni content from 50.4 to 51.2 at.%: Stress-strain curves up to plastic deformation of martensite were obtained, Schmid factors for the existing dislocation slip and twinning systems in martensite were calculated, and the mechanism of plastic deformation of martensite was studied. Based on the obtained results, it can be concluded that:
  • The weak orientation dependence of the critical stresses for the reorientation of the thermo-induced B19’-martensite at T < Mf is observed in 50.4Ni and 50.7Ni single crystals (σreor = 169–192 MPa for 50.4Ni and σreor = 138–153 MPa for 50.7Ni) due to the same initial self-accommodating structure consisting of 24 B19’-martensite variants.
  • In the high-nickel strain-glass 51.2Ni single crystals not undergoing thermoelastic MT during cooling/heating, the critical stress for the stress-induced B19’-martensite formation σcr depends on the crystal orientations. The level σcr is much higher than σreor for the reorientation of martensitic variants in single crystals with Ni content 50.4 and 50.7 at.%. In accordance with the Clausius–Clapeyron relation, the [111]B2-oriented 51.2Ni single crystals are characterized by higher values of the critical stress for the stress-induced martensite formation σcr = 876 MPa because of a smaller transformation strain εtr = 3.58% compared with the [001]B2-oriented single crystals (εtr = 4.38% and σcr = 412 MPa).
  • It is shown that an inverse orientation dependence of the B19’-martensite yield stress σcrM[111] > σcrM[001] is observed in comparison with the B2-phase yield stress σcrA[111] < σcrA[001] in TiNi single crystals with different Ni content CNi = 50.4–51.2 at.% in compression. With the Ni content, the difference between the stresses σcrM[111] and σcrM[001] increases from 272 MPa at CNi = 50.4 at.% to 996 MPa at CNi = 51.2 at.%. The main mechanisms of plastic deformation of B19’-martensite in compression are <001>{100}B19’ dislocation slip and {20 1 ¯ }B19’ mechanical twinning.
  • It is necessary to take into account the orientation dependence of martensite yield stress to predict the SE temperature range and cyclic stability of SME and SE in TiNi alloys. A significant effect of stress level σcrM on the functional properties should be expected in high-nickel [001]B2-oriented single crystals of TiNi alloys with CNi ≥ 51.2 at.%. In this case, martensite yield stress σcrM can be lower than austenite yield stress σcrA. This will contribute to the relaxation of elastic energy due to the plastic deformation of B19’-martensite during the stress-induced MT and degradation of functional properties.

Author Contributions

Conceptualization, E.Y.P. and E.E.T.; data curation, A.S.E. and E.Y.P.; investigation, A.S.E., I.D.F., N.Y.S. and M.V.Z.; methodology, I.D.F.; project administration, A.S.E.; resources, Y.I.C.; supervision, E.Y.P. and Y.I.C.; validation, M.V.Z., A.I.T. and N.Y.S.; visualization, A.S.E., I.D.F. and E.E.T.; writing—original draft, A.S.E.; writing—review and editing, E.Y.P., A.S.E., A.I.T. and Y.I.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by grant under the Decree of the Government of the Russian Federation No. 220 of 9 April 2010 (Agreement No. 075-15-2021-612 of 4 June 2021) and by the Tomsk State University Development Programme Priority-2030 (2.0.10.22).

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

Electron microscopy studies were carried out on the equipment of the Krasnoyarsk Regional Center for Collective Use of the FRC KSC SB RAS.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ma, J.; Karaman, I.; Noebe, R.D. High temperature shape memory alloys. Int. Mater. Rev. 2010, 55, 257–315. [Google Scholar] [CrossRef]
  2. Chowdhury, P.; Sehitoglu, H. Deformation physics of shape memory alloys. Prog. Mater. Sci. 2017, 88, 49–88. [Google Scholar] [CrossRef]
  3. Ren, X. Strain glass and ferroic glass—Unusual properties from glassy nano-domains. Phys. Status Solidi B 2014, 251, 1982–1992. [Google Scholar] [CrossRef]
  4. Wang, Y.; Ren, X.; Otsuka, K. Strain glass: Glassy martensite. Mater. Sci. Forum 2008, 583, 67–84. [Google Scholar] [CrossRef]
  5. Chowdhury, P.; Sehitoglu, H. A revisit to atomistic rationale for slip in shape memory alloys. Prog. Mater. Sci. 2017, 85, 1–42. [Google Scholar] [CrossRef]
  6. Yinong, L.; Galvin, S.P. Criteria for pseudoelasticity in near-equiatomic NiTi shape memory alloys. Acta Mater. 1997, 45, 4431–4439. [Google Scholar] [CrossRef]
  7. Otsuka, K.; Ren, X. Physical metallurgy of Ti–Ni-based shape memory alloys. Prog. Mater. Sci. 2005, 50, 511–678. [Google Scholar] [CrossRef]
  8. Otsuka, K.; Wayman, C.M. Shape Memory Materials; Cambridge University Press: Cambridge, UK, 1998; p. 284. [Google Scholar]
  9. Chumlyakov, Y.I.; Kireeva, I.V.; Panchenko, E.Y.; Timofeeva, E.E.; Kretinina, I.V.; Kuts, O.A. Physics of thermoelastic martensitic transformation in high-strength single crystals. Mater. Sci. Found 2015, 81, 107–173. [Google Scholar] [CrossRef]
  10. Chaugule, P.S.; le Graverend, J.-B. Crystal-plasticity modeling of phase transformation-viscoplasticity coupling in high-temperature shape memory alloys. Int. J. Plast. 2022, 153, 103243. [Google Scholar] [CrossRef]
  11. Timofeeva, E.E.; Panchenko, E.Y.; Tagiltsev, A.I.; Larchenkova, N.G.; Tokhmetova, A.B.; Chumlyakov, Y.I.; Volochaev, M.N. Orientation dependence of superelasticity in quenched high-nickel Ti-51.8Ni single crystals. Mater. Lett. 2021, 282, 128677. [Google Scholar] [CrossRef]
  12. Sehitoglu, H.; Karaman, I.; Anderson, R.; Zhang, X.; Gall, K.; Maier, H.J.; Chumlyakov, Y. Compressive response of NiTi single crystals. Acta Mater. 2000, 48, 3311–3326. [Google Scholar] [CrossRef]
  13. Sehitoglu, H.; Jun, J.; Zhang, X.; Karaman, I.; Chumlyakov, Y.; Maier, H.J.; Gall, K. Shape memory and pseudoelastic behavior of 51.5%Ni-Ti single crystals in solutionized and overaged state. Acta Mater. 2001, 49, 3609–3620. [Google Scholar] [CrossRef]
  14. Gall, K.; Sehitoglu, H.; Chumlyakov, Y.I.; Kireeva, I.V. Tension-compression asymmetry of the stress-strain response in aged single crystal and polycrystalline NiTi. Acta Mater. 1999, 47, 1203–1217. [Google Scholar] [CrossRef]
  15. Sehitoglu, H.; Hamilton, R.; Canadinc, D.; Zhang, X.Y.; Gall, K.; Karaman, I.; Chumlyakov, Y.; Maier, H.J. Detwinning in NiTi alloys. Metall. Mater. Trans. A Phys. 2003, 34, 5–13. [Google Scholar] [CrossRef]
  16. Kaya, I.; Karaca, H.E.; Souri, M.; Chumlyakov, Y.I.; Kurkcu, H. Effects of orientation on the shape memory behavior of Ni51Ti49 single crystals. Mater. Sci. Eng. A 2017, 686, 73–81. [Google Scholar] [CrossRef]
  17. Surikova, N.S.; Klopotov, A.A.; Korznikova, E.A. Mechanisms of plastic deformation in microcrystalline and nanocrystalline TiNi-based alloys. Phys. Met. Metallogr. 2010, 110, 269–278. [Google Scholar] [CrossRef]
  18. Zhang, J.; Xu, Y.; Otsuka, K.; Ren, X.; Chumlyakov, Y.I.; Asai, J. Orientation dependence of stress-induced martensitic transformation in quenched Ti–50.8at.%Ni single crystals. J. Phys. IV Fr. 2003, 112, 669–672. [Google Scholar] [CrossRef]
  19. Šittner, P.; Molnárová, O.; Bian, X.; Heller, L.; Seiner, H. Tensile deformation of B19’ martensite in nanocrystalline NiTi wires. Shap. Mem. Superelasticity 2023, 9, 11–34. [Google Scholar] [CrossRef]
  20. Molnárová, O.; Klinger, M.; Duchoň, J.; Seiner, H.; Šittner, P. Plastic Deformation of B19’ monoclinic martensite in NiTi shape memory alloys: HRTEM analysis of interfaces in martensite variant microstructures. Acta Mater. 2023, 258, 119242. [Google Scholar] [CrossRef]
  21. Liu, Y.; Xie, Z. Progress in Smart Materials and Structures; Chapter 3; Reece, P.L., Ed.; Nova Science Publisher, Inc.: New York, NY, USA,, 2007; pp. 29–65. [Google Scholar]
  22. Liu, Y.; Xie, Z.; Van Humbeeck, J.; Delaey, L. Asymmetry of stress-strain curves under tension and compression for NiTi shape memory alloys. Acta Mater. 1998, 46, 4325–4338. [Google Scholar] [CrossRef]
  23. Alarcon, E.; Heller, L.; de Prado, E.; Kopeček, J. Temperature and microstructure dependent tensile behavior of coarse grained superelastic NiTi. Mater. Des. 2023, 226, 111617. [Google Scholar] [CrossRef]
  24. Chen, Y.; Li, A.; Ma, Z.; Kang, G.; Jiang, D.; Zhao, K.; Ren, Y.; Cui, L.; Yu, K. Revealing the mode and strain of reversible twinning in B19’ martensite by in situ synchrotron X-ray diffraction. Acta Mater. 2022, 236, 118131. [Google Scholar] [CrossRef]
  25. Zhang, J.X.; Sato, M.; Ishida, A. Deformation mechanism of martensite in Ti-rich Ti-Ni shape memory alloy thin films. Acta Mater. 2006, 54, 1185–1198. [Google Scholar] [CrossRef]
  26. Sridhar, S.K.; Stebner, A.P.; Rollett, A.D. Statistical variations in predicted martensite variant volume fractions in superelastically deformed NiTi modeled using habit plane variants versus correspondence variants. Int. J. Solids Struct. 2021, 221, 60–76. [Google Scholar] [CrossRef]
  27. Krishan, M.; Singh, J.B. A novel B19’ martensite in nickel titanium shape memory alloys. Acta Mater. 2000, 48, 1325–1344. [Google Scholar] [CrossRef]
  28. Akamine, H.; Heima, A.; Soejima, Y.; Mitsuhara, M.; Inamura, T.; Nishida, M. Where and when are dislocations induced by thermal cycling in Ti-Ni shape memory alloys? Acta Mater. 2023, 244, 118588. [Google Scholar] [CrossRef]
  29. Polatidis, E.; Šmíd, M.; Kuběna, I.; Hsu, W.-N.; Laplanche, G.; Van Swygenhoven, H. Deformation mechanisms in a superelastic NiTi alloy: An in-situ high resolution digital image correlation study. Mater. Des. 2020, 191, 108622. [Google Scholar] [CrossRef]
  30. Nishida, M.; Nishiura, T.; Kawano, H.; Inamura, T. Self-accommodation of B19’ martensite in Ti-Ni shape memory alloys-Part I. Morphological and crystallographic studies of the variant selection rule. Philos. Mag. 2012, 92, 2215–2233. [Google Scholar] [CrossRef]
  31. Nishida, M.; Yamauchi, K.; Itai, I.; Ohgi, H.; Chiba, A. High resolution electron microscopy studies of twin boundary structures in B19’ martensite in the Ti-Ni shape memory alloy. Acta Metall. Mater. 1995, 43, 1229–1234. [Google Scholar] [CrossRef]
  32. Pelton, A.R.; Dicello, J.; Miyazaki, S. Optimisation of processing and properties of medical grade Nitinol wire. Minim. Invasive Ther. Allied Technol. 2000, 9, 107–118. [Google Scholar] [CrossRef]
  33. Kadeřávek, L.; Šittner, P.; Molnárová, O.; Klimša, L.; Heller, L. Localized plastic deformation of superelastic NiTi wires in tension. Shap. Mem. Superelast. 2023, 9, 261–279. [Google Scholar] [CrossRef]
  34. Surikova, N.S.; Korznikova, E.A. Structural fragmentation in TiNi(Fe, Mo) single crystals during intense plastic deformation. Steel Transl. 2013, 43, 630–634. [Google Scholar] [CrossRef]
  35. Surikova, N.S.; Evtushenko, O.V.; Pavlyuk, V.A. Strain localization and temperature dependence of yield stress in TiNi-based single crystals. Phys. Mesomech. 2010, 13, 96–102. [Google Scholar] [CrossRef]
  36. Niitsu, K.; Kainuma, R. Effect of annealing on stress-induced transformation behaviors at low temperatures in a Ti-51.8 at.% Ni shape memory alloy. Phys. Status Solidi B 2014, 251, 2041–2047. [Google Scholar] [CrossRef]
  37. Wang, H.; Huang, H.-Y.; Su, Y.-J. Tuning the operation temperature window of the elastocaloric effect in Cu-Al-Mn shape memory alloys by composition design. J. Alloys Compd. 2020, 828, 154265. [Google Scholar] [CrossRef]
  38. Masdeu, F.; Pons, J.; Torrens-Serra, J.; Chumlyakov, Y.; Cesari, E. Superelastic behavior and elastocaloric effect in a Ni51.5Fe21.5Ga27.0 ferromagnetic shape memory single crystal under compression. Mater. Sci. Eng. A 2022, 833, 142362. [Google Scholar] [CrossRef]
  39. Surikov, N.Y.; Panchenko, E.Y.; Timofeeva, E.E.; Tagiltsev, A.I.; Chumlyakov, Y.I. Orientation dependence of elastocaloric effect in Ni50Mn30Ga20 single crystals. J. Alloys Compd. 2021, 880, 160553. [Google Scholar] [CrossRef]
  40. Li, P.; Karaca, H.E.; Chumlyakov, Y.I. Orientation dependent compression behavior of Co35Ni35Al30 single crystals. J. Alloys Compd. 2017, 718, 326–334. [Google Scholar] [CrossRef]
  41. Nishida, M.; Wayman, C.W.; Chiba, A. Electron microscopy studies of the martensitic transformation in an aged Ti-51at.%Ni shape memory alloy. Metallography 1988, 21, 275–291. [Google Scholar] [CrossRef]
  42. Ezaz, T.; Sehitoglu, H.; Abuzaid, W.; Maier, H.J. Higher order twin modes in martensitic NiTi-The (20 1 ¯ ) case. Mater. Sci. Eng. A 2012, 558, 422–430. [Google Scholar] [CrossRef]
  43. Seiner, H.; Sedlák, P.; Frost, M.; Šittner, P. Kwinking as the plastic forming mechanism of B19’ NiTi martensite. Int. J. Plast. 2023, 168, 103697. [Google Scholar] [CrossRef]
Figure 1. Model of stress versus temperature for the SMAs, where Ms, Mf are the forward MT start and finish temperatures; As, Af are the reverse MT start and finish temperatures, Md is the temperature above which the stress-induced MT is no longer possible and the plastic deformation of high-temperature B2-phase sets in [1,2,5,8,9,10]. The temperature intervals of SME and SE are shown.
Figure 1. Model of stress versus temperature for the SMAs, where Ms, Mf are the forward MT start and finish temperatures; As, Af are the reverse MT start and finish temperatures, Md is the temperature above which the stress-induced MT is no longer possible and the plastic deformation of high-temperature B2-phase sets in [1,2,5,8,9,10]. The temperature intervals of SME and SE are shown.
Metals 13 01567 g001
Figure 2. DSC curves during cooling/heating for quenched TiNi single crystals.
Figure 2. DSC curves during cooling/heating for quenched TiNi single crystals.
Metals 13 01567 g002
Figure 3. Frequency-temperature dependencies of the elastic modulus E and internal friction tan δ at cooling for [001]B2-oriented quenched 51.2Ni single crystals.
Figure 3. Frequency-temperature dependencies of the elastic modulus E and internal friction tan δ at cooling for [001]B2-oriented quenched 51.2Ni single crystals.
Metals 13 01567 g003
Figure 4. The compressive stress-strain curves at loading in the martensite phase at T < Ms for quenched TiNi single crystals. Loading/unloading curves are shown in inset: (a) 50.4Ni at T = 258 K; (b) 50.7Ni at T = 203 K.
Figure 4. The compressive stress-strain curves at loading in the martensite phase at T < Ms for quenched TiNi single crystals. Loading/unloading curves are shown in inset: (a) 50.4Ni at T = 258 K; (b) 50.7Ni at T = 203 K.
Metals 13 01567 g004
Figure 5. The compressive stress-strain curves at loading in the austenite phase at T = 203 K for quenched 51.2Ni single crystals. Loading/unloading curves are shown in inset.
Figure 5. The compressive stress-strain curves at loading in the austenite phase at T = 203 K for quenched 51.2Ni single crystals. Loading/unloading curves are shown in inset.
Metals 13 01567 g005
Figure 6. Bright field image of the dislocation structure in TiNi single crystals after the plastic deformation of the B19’-martensite up to εg = 12.0–15.5% in compression, followed by heating to room temperature: (a) [001]B2-oriented 50.4Ni single crystal; (b) [111]B2-oriented 50.7Ni single crystal.
Figure 6. Bright field image of the dislocation structure in TiNi single crystals after the plastic deformation of the B19’-martensite up to εg = 12.0–15.5% in compression, followed by heating to room temperature: (a) [001]B2-oriented 50.4Ni single crystal; (b) [111]B2-oriented 50.7Ni single crystal.
Metals 13 01567 g006
Figure 7. TEM image of the microstructure of the [001]B2-oriented 50.4Ni single crystals after the plastic deformation of the B19’-martensite up to εg = 12.0–15.5% in compression, followed by heating to room temperature: (a) joint bright field image of multiple twinning of B2-phase and B19’-martensite with corresponding selected area diffraction pattern (SADP); (b) bright field image of the residual B19’-martensite band (M) in the B2-austenite matrix (A) and the corresponding SADP from areas A ([001]B2 zone axis) and M ([100]B19’ zone axis).
Figure 7. TEM image of the microstructure of the [001]B2-oriented 50.4Ni single crystals after the plastic deformation of the B19’-martensite up to εg = 12.0–15.5% in compression, followed by heating to room temperature: (a) joint bright field image of multiple twinning of B2-phase and B19’-martensite with corresponding selected area diffraction pattern (SADP); (b) bright field image of the residual B19’-martensite band (M) in the B2-austenite matrix (A) and the corresponding SADP from areas A ([001]B2 zone axis) and M ([100]B19’ zone axis).
Metals 13 01567 g007
Figure 8. TEM image of the {411}B2 twins of the B2-phase in the [001]B2-oriented 50.4Ni single crystals after the plastic deformation of the B19’-martensite up to εg = 12.0–15.5% in compression, followed by heating to room temperature: (a) joint SADP from B2-matrix and twin, the [ 1 ¯ 13]m matrix zone is parallel to the [1 1 ¯ 3 ¯ ]tw twin zone; (b,c) dark field image from the matrix (m) and twin (tw) reflections marked by yellow circles, respectively.
Figure 8. TEM image of the {411}B2 twins of the B2-phase in the [001]B2-oriented 50.4Ni single crystals after the plastic deformation of the B19’-martensite up to εg = 12.0–15.5% in compression, followed by heating to room temperature: (a) joint SADP from B2-matrix and twin, the [ 1 ¯ 13]m matrix zone is parallel to the [1 1 ¯ 3 ¯ ]tw twin zone; (b,c) dark field image from the matrix (m) and twin (tw) reflections marked by yellow circles, respectively.
Metals 13 01567 g008
Figure 9. TEM image of the [111]B2-oriented 50.4Ni single crystals after the plastic deformation of the B19’-martensite up to εg = 12.0–15.5% in compression, followed by heating to room temperature.
Figure 9. TEM image of the [111]B2-oriented 50.4Ni single crystals after the plastic deformation of the B19’-martensite up to εg = 12.0–15.5% in compression, followed by heating to room temperature.
Metals 13 01567 g009
Figure 10. TEM image of the [111]B2-oriented 50.4Ni single crystals after the plastic deformation of the B19’-martensite up to εg = 12.0–15.5% in compression, followed by heating to room temperature: (a) joint bright field image of B2- and B19’-phases mix; (bd) SADPs from areas 1, 2, 3 marked by circles on (a).
Figure 10. TEM image of the [111]B2-oriented 50.4Ni single crystals after the plastic deformation of the B19’-martensite up to εg = 12.0–15.5% in compression, followed by heating to room temperature: (a) joint bright field image of B2- and B19’-phases mix; (bd) SADPs from areas 1, 2, 3 marked by circles on (a).
Metals 13 01567 g010
Figure 11. TEM image of the [111]B2-oriented 50.7Ni single crystals after the plastic deformation of the B19’-martensite up to εg = 12.0–15.5% in compression, followed by heating to room temperature: (a) bright field image; (b) joint SADP from B2-matrix and twin, marked areas 1 on (a), [011]m||[111]tw zone axis; (c,d) SADPs from B2-matrix of two neighboring domains, marked areas 2 and 2’ on (a), ([011]m zone axis); (e,f) SADPs from twin of two neighboring domains, marked areas 3 and 3’ on (a), ([ 1 ¯ 11]tw zone axis).
Figure 11. TEM image of the [111]B2-oriented 50.7Ni single crystals after the plastic deformation of the B19’-martensite up to εg = 12.0–15.5% in compression, followed by heating to room temperature: (a) bright field image; (b) joint SADP from B2-matrix and twin, marked areas 1 on (a), [011]m||[111]tw zone axis; (c,d) SADPs from B2-matrix of two neighboring domains, marked areas 2 and 2’ on (a), ([011]m zone axis); (e,f) SADPs from twin of two neighboring domains, marked areas 3 and 3’ on (a), ([ 1 ¯ 11]tw zone axis).
Metals 13 01567 g011
Figure 12. The scheme of the CVP in the [111]B2-oriented single crystals. The Schmid factors m for a slip in active CVP 2’(+) consisting of domain 2’ (m = 0.27) and domain 1 (m = 0.46).
Figure 12. The scheme of the CVP in the [111]B2-oriented single crystals. The Schmid factors m for a slip in active CVP 2’(+) consisting of domain 2’ (m = 0.27) and domain 1 (m = 0.46).
Metals 13 01567 g012
Table 1. B2-B19’ MT temperature during cooling/heating for quenched TiNi single crystals.
Table 1. B2-B19’ MT temperature during cooling/heating for quenched TiNi single crystals.
Ni, at.%Ms(±2), KMf(±2), KAs(±2), KAf(±2), KΔ1(±2), KΔ2(±2), K
50.42862722993161417
50.72401902182665048
51.2Undefined
Table 2. Lattice correspondences between B2-austenite and B19’-martensite. Data from [28,29,30].
Table 2. Lattice correspondences between B2-austenite and B19’-martensite. Data from [28,29,30].
Martensite Domain[100]B19’[010]B19’[001]B19’
1[100]B2[011]B2[0 1 ¯ 1]B2
1’[ 1 ¯ 00]B2[0 1 ¯ 1 ¯ ]B2[0 1 ¯ 1]B2
2[100]B2[0 1 ¯ 1]B2[0 1 ¯ 1 ¯ ]B2
2’[ 1 ¯ 00]B2[01 1 ¯ ]B2[0 1 ¯ 1 ¯ ]B2
3[010]B2[101]B2[10 1 ¯ ]B2
3’[0 1 ¯ 0]B2[ 1 ¯ 0 1 ¯ ]B2[10 1 ¯ ]B2
4[010]B2[10 1 ¯ ]B2[ 1 ¯ 0 1 ¯ ]B2
4’[0 1 ¯ 0]B2[ 1 ¯ 01]B2[ 1 ¯ 0 1 ¯ ]B2
5[001]B2[110]B2[ 1 ¯ 10]B2
5’[00 1 ¯ ]B2[ 1 ¯ 1 ¯ 0]B2[ 1 ¯ 10]B2
6[001]B2[ 1 ¯ 10]B2[ 1 ¯ 1 ¯ 0]B2
6’[00 1 ¯ ]B2[1 1 ¯ 0]B2[ 1 ¯ 1 ¯ 0]B2
Table 3. The 24 CVPs and system for the habit plane indices relative to the B2-phase. Data from [28,29,30].
Table 3. The 24 CVPs and system for the habit plane indices relative to the B2-phase. Data from [28,29,30].
CVPMartensite Domain PairsHabit Plane
1(+)1–2(−0.89 0.22 −0.40)
1(−)1–2’(−0.89 −0.40 0.22)
1’(+)1’–2(0.89 −0.40 0.22)
1’(−)1’–2’(0.89 −0.22 0.40)
2(+)2–1(0.89 0.22 0.40)
2(−)2–1’(0.89 0.40 0.22)
2’(+)2’–1(−0.89 0.40 0.22)
2’(−)2’–1’(−0.89 0.22 0.40)
3(+)3–4(−0.40 0.89 0.22)
3(−)3–4’(−0.22 0.89 0.40)
3’(+)3’–4(−0.22 −0.89 0.40)
3’(−)3’–4’(−0.40 −0.89 0.22)
4(+)4–3(0.40 0.89 0.22)
4(−)4–3’(0.22 0.89 0.40)
4’(+)4’–3(0.22 −0.89 0.40)
4’(−)4’–3’(0.40 −0.89 0.22)
5(+)5–6(0.22 −0.40 0.89)
5(−)5–6’(0.40 −0.22 0.89)
5’(+)5’–6(−0.40 0.22 0.89)
5’(−)5’–6’(−0.22 0.40 0.89)
6(+)6–5(0.22 0.40 0.89)
6(−)6–5’(0.40 0.22 0.89)
6’(+)6’–5(−0.40 −0.22 0.89)
6’(−)6’–5’(−0.22 −0.40 0.89)
Table 4. Active CVPs, calculated Schmid factors of martensite, and experimental yield stress of martensite (this study) depending on crystal orientation in TiNi single crystal.
Table 4. Active CVPs, calculated Schmid factors of martensite, and experimental yield stress of martensite (this study) depending on crystal orientation in TiNi single crystal.
Compression Axis in B2-PhaseCompression Axis in B19’-MartensiteActive CVPs
[12,15]
Schmid Factors of MartensiteσcrM, MPa
<001>B19’
{100}B19’ Dislocation Slip
< 1 ¯ 0 2 ¯ >B19’
{ 20 1 ¯ }B19’
Twinning
50.4
Ni
50.7
Ni
51.2
Ni
[001]B2[011]B19’ (1)(1,3) (3,1) (1,4) (4,1) (2,3) (3,2) (2,4) (4,2)0.490.37–0.3969211241023
[0 1 ¯ 1]B19’ (1’)
[01 1 ¯ ]B19’ (2, 3, 4’)
[0 1 ¯ 1 ¯ ]B19’ (2’, 3’, 4)
[100]B19’ (5, 6)
[ 1 ¯ 00]B19’ (5’, 6’)
[111]B2[120]B19’ (1,3, 5)(3,1) (4,1) (7,5) (8,5) (11,9) (12,9)0.27–0.460.39–0.596417142019
[ 1 ¯ 2 ¯ 0]B19’ (1’, 3’, 5’)
[ 1 ¯ 0 2 ¯ ]B19’ (2’, 4’, 6’)
[10 2 ¯ ]B19’ (2, 4, 6)
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Panchenko, E.Y.; Eftifeeva, A.S.; Fatkullin, I.D.; Tagiltsev, A.I.; Surikov, N.Y.; Zherdeva, M.V.; Timofeeva, E.E.; Chumlyakov, Y.I. Orientation Dependence of B19’-Martensite Reorientation Stress and Yield Stress in TiNi Single Crystals. Metals 2023, 13, 1567. https://doi.org/10.3390/met13091567

AMA Style

Panchenko EY, Eftifeeva AS, Fatkullin ID, Tagiltsev AI, Surikov NY, Zherdeva MV, Timofeeva EE, Chumlyakov YI. Orientation Dependence of B19’-Martensite Reorientation Stress and Yield Stress in TiNi Single Crystals. Metals. 2023; 13(9):1567. https://doi.org/10.3390/met13091567

Chicago/Turabian Style

Panchenko, Elena Y., Anna S. Eftifeeva, Ilya D. Fatkullin, Anton I. Tagiltsev, Nikita Y. Surikov, Maria V. Zherdeva, Ekaterina E. Timofeeva, and Yuriy I. Chumlyakov. 2023. "Orientation Dependence of B19’-Martensite Reorientation Stress and Yield Stress in TiNi Single Crystals" Metals 13, no. 9: 1567. https://doi.org/10.3390/met13091567

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop