1. Introduction
Industrial applications of near-equiatomic NiTi alloys benefit from the large amount of recoverable strain, on the order of percentages, that can be repeatedly induced and recovered by thermal and/or mechanical actuation. It stems from the reversible martensitic transformation (MT) between the parent austenite and product martensite phases that gives rise to related functional properties called superelasticity and shape memory. MT is a diffusionless solid-solid phase transition that proceeds via thermomechanically driven nucleation and propagation of interfaces between the two phases. One of the conditions for MT to be reversible is that the crystal structures of the two phases must be coherent at some crystallographic planes. Such habit planes ensure the propagation of austenite-martensite interfaces without creating lattice defects that would otherwise lead to unrecoverable strains and unstable thermomechanical response.
The reversibility of MT was related to no or negligible specific volume changes and compatibility at austenite-martensite interfaces [
1]. The latter is related to the existence of compatible habit planes. These are invariant lattice planes (ILPs) shared by the austenite and martensite lattices. It has been proposed that such low-interface-energy habit planes propagate and transform the austenite structure into the martensite structure, or vice versa, without the generation of crystal defects. The ILP exists when the symmetric distortion matrix of the MT is such that its second eigenvalue, denoted
, is equal to 1 and the associated transformation strain, denoted
, is equal to 0. Furthermore, invariant planes also exist between austenite and martensite microstructures formed from martensite variants twinned at arbitrary volume fraction ratios [
1] if the distorsion matrix satisfies a set of so-called co-factor conditions, including
. Such supercompatibility [
2] ensures invariant lattice planes for a range of transformation strains, which can thus help accommodate applied macroscopic deformation. Finally, four- [
3] and three-fold [
4] conditions of compatibility between domains of twinned martensite have been formulated. Such compatibility is thought to ensure that no crystal defects are generated in the process of detwinning of martensite when it is subjected to external loading.
Whether or not the crystallographic conditions are satisfied depends on the type of lattice and the lattice parameters. Unfortunately, NiTi alloys do not satisfy the
condition or the set of cofactor conditions. Interestingly, however, NiTi alloys are the most widely used SMAs due to their resistance to plastic deformation and relatively stable thermo-mechanical response. Theoretically, the only way for monoclinic B19’ martensite in NiTi alloys to have ILPs with cubic B2 austenite is on a larger scale of Type I or Type II twins of pairs of lattice correspondence variants (CVs) of martensite. The Phenomenological Theory of Martensite Crystallography (PTMC) extended to shape memory alloys [
5,
6] provides several solutions for so-called Habit Plane Variants (HPVs) specified in terms of pairs of CVs and their volume fraction ratios for which ILPs with austenite exist [
3]. Some of these HPVs were found to be consistent with experimental observations of specimens subjected to stress-free MT. Experimentally observed habit planes were to some consistent with PTMC solutions for
Type II HPVs [
7,
8,
9] and for
Type I HPVs, although less frequently [
9]. Nevertheless, Cayron [
10] critically evaluated these experimental results and concluded that their scatter from PTMC solutions is beyond the experimental error. More recently, specific clusters of multiple
Type II HPVs [
11] or
compound twins [
12] have been experimentally observed and justified using PTMC [
4]. These self-accommodating martensite microstructures observed during stress-free MT provide both a plausible habit plane and a minimized average transformation strain.
Although martensite microstructures provide coherent habit planes as described above, plastic deformation processes occur in these alloys whenever MT proceeds under external stress. We have systematically investigated plastic deformation processes in NiTi wires with defined initial microstructures of nano- to micrometer-sized grains. Surprisingly, plasticity in NiTi alloys was found to be related not only to loading above the yield stress [
13,
14,
15,
16,
17], but also to forward and reverse MT [
18,
19,
20,
21,
22], even when proceeding under stresses well below the yield stresses of austenite and martensite [
19,
21]. In contrast, reorientation of martensite by loading to its yield stress followed by unloading and stress-free reverse MT into austenite did not produce any appreciable irreversible strain [
19]. Thus, it is clear that strained MT proceeds differently than stress-free MT. Thus, for theoretical predictions of strain planes under stress, PTMC should consider the effects of deformation constraints and external stresses.
Published experimental results on martensite microstructures produced by stress-induced MT have led to conflicting conclusions. Partly this is certainly due to the different types of NiTi samples that were under investigation. Transmission electron microscopy (TEM) studies suggest that stress-induced martensite microstructures in single crystals or coarse-grained NiTi are formed from
type II twins [
23], while nano-grained NiTi are preferentially formed from
compound twins [
12,
24]. The preferred orientations of stress-induced martensite in
textured nano-grained NiTi suggest that stress-induced MT favors
twins, which provide the largest transformation strain in the loading direction, as measured by in-situ synchrotron X-ray diffraction during stress-induced MT [
25]. Martensite textures analyzed in a deformed NiTi sheet using EBSD [
26] also met the criterion of the largest transformation strain and interaction work. In addition, the appearance of single variants of martensite rather than HPVs was suggested by the texture predictions. On the other hand, close agreement with simulated virtual diffraction of HPVs was demonstrated by high energy diffraction experiment during in-situ tension of NiTi single crystals [
27]. In addition, the identified HPVs generally did not provide the largest transformation strain and maximum deformation work. This may be due to the presence of Ni
4Ti
3 precipitates, subgrains or intermediate R-phase in the samples studied. To sum up, there is no general agreement whether stress-induced MT tends to form martensite microstructures from HPVs, single variants of martensite or
compound twins, and whether the criterion of maximum transformation strain holds for selecting martensite microstructures formed under the characteristic stress at which MT is induced.
Experimental identification of the habit planes between austenite and stress-induced martensite has been carried out by means of TEM and scanning electron microscopy (SEM). Surprisingly, these published results suggest that stress-induced MT in polycrystalline NiTi proceeds, at least to some extent, via the propagation of interfaces between austenite and a single variant of martensite. In-situ experiments on NiTi micro-samples showed the propagation of single variants of martensite during straining in TEM [
28,
29]. Based on statistical analysis of transformation strains measured in-situ in SEM, a non-negligible fraction of single variants of martensite together with HPVs was estimated in strained NiTi [
30]. Post-mortem TEM analysis of deformed superelastic NiTi [
16] showed interfaces between austenite and a single variant of residual martensite in the vicinity of the austenite
plane. Similarly, the austenite
plane was suggested as the habit plane with single variant of martensite by post-mortem statistical electron backscatter diffraction (EBSD) analysis on the deformed NiTi sample [
10]. Note that Miyazaki et al. [
7] in 1984 identified a similar habit plane at
using optical trace analysis of deformed solution annealed NiTi single crystal. Most recently, a new habit plane
between austenite and a single variant of martensite was experimentally measured in [
31] and explained by the PTMC using dislocation slip
as lattice invariant shear. However, these published results are not conclusive with respect to the reversibility of interfaces with individual variants of the martensite. On the one hand, the statistical analysis of the transformation strains correlated the appearance of single variants with higher residual strains as measured after unloading. Moreover, post-mortem observations of residual martensite [
16,
32] at temperatures above the austenite finish temperature (
) suggest their irreversibility. This is also supported by the curved nature of the interfaces. On the other hand, in-situ TEM observations [
28,
29] have shown their reversibility in polycrystalline samples.
This work contributes to the understanding of stress-induced MT in NiTi by investigating the mechanical conditions under which a coherent interface between cubic B2 austenite and a single variant of monoclinic B19’ becomes possible. To do this, we extend PTMC to include the effect of elastic deformation of the austenite and martensite lattices due to external stress. To the best of our knowledge, such an attempt has only been made in [
27], where the authors evaluated the effect of hydrostatic strain on the volume fractions of the CVs in HBVs.
It is crucial to note that we will address the issue of stress-induced compatibility by relying on the widely accepted lattice correspondence between cubic B2 austenite and monoclinic B19’ martensite. Despite this assumption, we will determine the deformation gradients between the two lattices by employing lattice vectors elastically distorted through uniaxial tensile or compression loading, in consideration of all loading directions.
In the following sections, we introduce the problem of austenite-martensite incompatibility in NiTi resulting from the orientation-dependent transformation strain, specifically
. Additionally, we provide an overview of the elastic anisotropy of austenite and martensite, considering their orientation relationship in NiTi. The following section provides a description of the PTMC method, which incorporates external stress, along with reference to
Appendix A for detailed descriptions. Subsequently, this method is applied to gain insight into the sensitivity of the second eigenvalue
and the associated transformation strain
to uniaxial stress and lattice softening of austenite in the
shear mode. The potential existence of the latter is based on C’ modulus softening, reported as a premartensitic phenomenon in NiTi during thermally induced stress-free MT [
33]. Although C’ modulus softening has not been reported as a precursor to stress-induced MT in NiTi, it has been observed in a single crystal of CuAlNi [
34]. To exemplify the effect of C’ softening, we first present the orientation dependencies of critical stresses necessary for
to become zero that allows for the habit plane formation. Subsequently, this paper demonstrates that the softening of C’ decreases the uniaxial stress, required to reach
, to realistic values comparable to those typically utilized. Alongside, the normals of habit planes between austenite and a single variant of martensite under external stress are reported. They are compared with experimentally identified habit planes observed under tension, namely, with
,
[
7],
[
8],
[
35].
3. Preliminary Insight into Elastic Deformation Effects on Austenite-Martensite Incompatibility
The elastic properties are essential to this work, that aims to understand the role of elastic strains in the compatibility problems at the austenite-martensite interfaces. Specifically, whether the elastic deformation of both lattice might make the second principal transformation strain vanish thus enabling the habit plane formation between austenite and a single variant of martensite. The example of such situation is shown in
Figure 3 by the geometrical model of the two lattices in the initial nondeformed and elastically deformed states. The latter is result of our computations presented hereinafter. It represents the two lattices deformed due to a uniaxial tension along the indicated direction and a softening of C’ elastic constant of austenite. It will be shown that such circumstances can indeed enable the habit plane formation as illustrated in
Figure 3.
The formation of the habit plane presented in
Figure 4a is effectively due to relative differences in the directional Young’s moduli and Poisson’s ratios of the two crystals. The former is presented in
Figure 4 for austenite (
Figure 4a) and martensite (
Figure 4b). For comparison purposes, the are plotted in the common fundamental zone of martensite, taking into account the orientation relationship given by the lattice correspondence. The color mapss in
Figure 4 also divide the orientation space into zones of tensile and compressive transformation strains by the black isoline. Note that the actual values of the elastic constants used in this work are described in detail
Section 4. Furthermore, the directional Young’s modulus of austenite in the equal-area triangle is shown in
Figure A2 in
Appendix B.1.
As pointed out in [
36,
37], the stiffest
austenite direction is near the softest
martensite in the zone of tensile transformation strains. Consequently, one can estimate the effect of tensile loading along this direction on the negative value of
along the perpendicular
. Since martensite is softer than austenite along
, contractions along
due to Poisson effect will be larger for martensite, thus making the magnitude of
even larger. One can thus estimate a negative effect of this type of loading on austenite-martensite compatibility. Note that in the zone of tensile transformation strain we do not assume compressive loading and vice versa, since the negative deformation energy does not make physical sense. In the case of compressive loading, the effect of elastic deformation can be estimated for the loading direction
, i.e., along the direction of
. In this case, it is the deformation along the loading direction that affects
, as opposed to the Poisson effect in the previous case. Since martensite is stiffer than austenite along the loading direction, a reduction in the magnitude of
can be expected, i.e., an improvement in austenite-martensite compatibility. In summary, the effect of elastic deformation on
depends on the difference in elastic properties, the direction of loading and its deviation from the direction of
, and the loading mode (tension/compression). In this paper we analyze this dependence and identify the critical uniaxial loading stresses that make
vanish. It is shown that such conditions can be satisfied only for a subset of loading directions and magnitudes on the order of GPa. We also simulate the effect of austenite instability on the critical stresses. We assume an elastic constant softening C’ and search for its critical value at which
vanishes for a fixed reasonable value of the loading stress.
Figure 4.
Directional Young’s modulus of austenite
(b) and martensite
(a), both plotted in the coordinate system of the martensite lattice.
(a) is based on the set of elastic constants of martensite considered in this paper from the ab-initio calculation [
38].
(b) is based on experimental results [
33]. The black isoline divides the orientation space into tensile and compressive transformation strain zones. Superscript
M and black solid circles denote crystallographic directions of the martensite lattice. Superscript
A and white solid circles denote crystallographic directions of the austenite lattice, taking into account the orientation relationship of martensitic transformation in NiTi.
Figure 4.
Directional Young’s modulus of austenite
(b) and martensite
(a), both plotted in the coordinate system of the martensite lattice.
(a) is based on the set of elastic constants of martensite considered in this paper from the ab-initio calculation [
38].
(b) is based on experimental results [
33]. The black isoline divides the orientation space into tensile and compressive transformation strain zones. Superscript
M and black solid circles denote crystallographic directions of the martensite lattice. Superscript
A and white solid circles denote crystallographic directions of the austenite lattice, taking into account the orientation relationship of martensitic transformation in NiTi.
6. Discussion
The martensitic transformation (MT) in NiTi typically occurs by two habit plane variants (HPVs) of martensite propagating. This is because they create an invariant plane strain deformation (IPS) on a larger scale than a lattice unit cell. It is impossible on a single martensite lattice level since they are not compatible. The second principal transformation strain is -3.3%, while the IPS requires to be zero. This is a commonly accepted description of MT in NiTi, although it does not account for lattice changes caused by external stress fields. Our calculations demonstrate that elastic strains from typically applied stress of 500 MPa modify the value of by approximately 0.3 % in magnitude and alter its directions from by multiple degrees. Since the volume ratio of martensite variants and the orientation of habit planes formed by HPVs depend on , the volume fractions of HPVs and associated habit planes have to be considered as stress-dependent in tha case of stress-induced MT.
Larger stress can be considered in polycrystals where grain misorientations cause stress inhomogeneities within and among grains, with stress peaks exceeding the nominal values. Our calculations indicate that, as stresses increase, the anisotropy of elasticity effects changes due to significant deformation and considerable deviation of the direction from . Under large stresses, the regions of loading directions that positively effects the compatibility between austenite and martensite greatly increase. The changes are rather counter-intuitive as the loading orientations having negative effects of compatibility at moderate stresses become to have positive effects at large stresses. We attribute this to a change in direction that can no longer by associated with direction at high stresses and elastic distortions.
As a limit case, it is worth considering the critical loading stresses necessary to achieve compatibility between austenite and martensite, specifically the elimination of
. Our calculation showed that there is orientation selectivity and tension-compression asymmetry. Firstly, only a subset of loading orientations result in compatibility, with a larger subset for compression. For tension, only orientations around
are suitable for stress-induced compatibility, but at the cost of high stresses of 7-10 GPa. In contrast, compression provides a larger orientation space for loading direction, particularly
, where the critical stress is around 3-4 GPa. The critical loading stresses in tension were most frequently associated with habit planes in the vicinity of the pole
, including the loading directions providing the highest transformation strains of above 8 %. In contrast, under the critical compression stresses, the habit planes associated with the largest transformation strain magnitudes above 5 % do not fall in into the family of planes near the pole
most frequently enabled by compresssion. Note that, the two experimentally identified habit planes observed under tension
[
7]
[
35] fall in the proximity of the most frequent habit planes calculated for critical tensile loading.
In a similar way to thermally induced MT, the austenite lattice may be considered to be unstable when it converts into martensite under the action of an external stress. This instability can be replicated through softening of C’ elastic constant softening, resulting in a significant drop in the critical stress making austenite and martensite compatible. Furthermore, the orientation dependence of critical loading differs when considering the instability. Our simulation showed that significant C’ softening below 2 GPa reduces the critical loading stress for a habit plane with a single martensite variant to 500 MPa. It should be noted that C’ softening favors the formation of habit planes within larger loading orientation space compared to the case without softening considered. In contrast, the orientations space of compression loading was larger in the case of high stress magnitude and absence of C’ softening. Under the selected stress of 500 and C’ softening allowed down to 0.5 GPa, compression only along loading directions concentrated around
and
allowed for habit planes with a single martensite variant. In terms of habit planes activated by the combined effect of stress and C softening, they are similar to those activated in the absence of C softening for tension only. The most frequent habit planes calculated for tension and C’ softening were located near the pole
and high-index
. These poles coincide with habit planes related to the largest tensile transformation strain, which exceeds 8%. Note that, experimentally identified habit planes observed under tension
,
[
7],
[
8],
[
35] fall in the proximity of the most frequent habit planes calculated for tension and C’ softening.
In comparison with pure compression effects, the softening effect of C’ allows for the emergence of new habit planes under compression. These habit planes are situated in close proximity to the pole , similar to tension, and also consist of high-index normals situated between poles and . A larger transformation strain magnitude is generally associated with only the most common habit planes neighboring the pole . Otherwise, the significant strains resulting from transformation during compression are linked to habit planes located near and that are not among the most frequent ones.