Next Article in Journal
Development of Ternary Ti-Ag-Cu Alloys with Excellent Mechanical Properties and Antibiofilm Activity
Previous Article in Journal
Study on Bond Performance between Corroded Deformed Steel Bar and DS-ECC
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Thermodynamic Analysis of Anomalous Shape of Stress–Strain Curves for Shape Memory Alloys

1
Department of Solid State Physics, University of Debrecen, P.O. Box 400, H-4002 Debrecen, Hungary
2
Physics Department, Faculty of Science, Ain Shams University, Cairo 11566, Egypt
*
Author to whom correspondence should be addressed.
Materials 2022, 15(24), 9010; https://doi.org/10.3390/ma15249010
Submission received: 6 November 2022 / Revised: 5 December 2022 / Accepted: 9 December 2022 / Published: 16 December 2022
(This article belongs to the Section Materials Physics)

Abstract

:
In some shape-memory single crystals the stress–strain (σ~ε) curves, belonging to stress induced martensitic transformations from austenite to martensite at fixed temperature, instead of being the usual slightly increasing function or horizontal, have an overall negative slope with sudden stress drops in it. We discuss this phenomenon by using a local equilibrium thermodynamic approach and analysing the sign of the second derivative of the difference of the Gibbs free energy. We show that, considering also the possible nucleation and growth of two martensite structural modifications/variants, the stress–strain loops can be unstable. This means that the overall slope of the uploading branch of the stress–strain curve can be negative for smooth transformation if the second martensite, which is more stable with larger transformation strain, is the final product. We also show that local stress-drops on the stress–strain curve can appear if the nucleation of the second martensite is difficult, and the presence of such local stress-drops alone can also result in an overall negative slope of the stress–strain curves. It is illustrated that the increase of the temperature of the thermal recovery during burst-like transition is a measure of the change of the nucleation energy: the more stable martensite has larger nucleation energy.

1. Introduction

It is known that in some shape memory single crystals the stress–strain (σ~ε) curves, belonging to stress induced martensitic transformations at fixed temperature, can show anomalous shape: e.g., the uploading branch of the stress–strain curve, instead being the usual smooth, slightly increasing function, can have an overall negative slope with sudden stress drops on it [1,2,3,4,5,6,7,8,9,10] (Figure 1). This behaviour shows quite a wide variety. It can be strongly anisotropic and can differ considerably along different crystallographic orientations (see e.g., Figure 1 in [2], where the upper branch of the σ~ε curves, along [100]A as well as [110]A directions in a Ni49Fe18Ga27Co6 single crystal, was normal as well as anomalous, respectively). It is often observed that the transformation is not complete [1,7,8,9], but there are also examples when the whole sample has been transformed [4,10]. Furthermore, the strain recovery of the stress induced martensite to austenite during heating can be very fast (burst-like recovery). For instance, it was observed that the DSC peak of this recovery was only about 10−3–10−5 degree wide (instead of the usual 1–50 K wide transitions at typical 1–10 K/min rates) and was accompanied with jumping of the sample as well as with audible click [1,2,3,4,5,6,7,8,9,10]. In addition, the DSC peak appeared at higher temperature (by 10–60 K higher) than the corresponding peak of the reverse transformation measured after thermally induced cycling. This indicates that the martensite structure formed during stress induced changes is more stable than the thermally induced one. Both the martensite stabilization, manifested in the shift of the DSC peak to higher temperatures during heating, and the anomalous stress–strain curves, are interpreted by the presence/competition of two different martensitic structural modifications [2,3,9]. These can be denoted as M1 and M2 structures, respectively. For instance, in Ni49Fe18Ga27Co6 single crystal M1 and M2 were identified as twinned 14 M martensite, as well as L1o detwinned tetragonal martensite, respectively [3], or in Cu–Al–Ni alloys as β′ (18 R monoclinic) and/or twinned γ′ (2 H orthorhombic) as well as detwinned γ′ phases, respectively [7,8,11]. It can be added that the appearance of stress drops/jumps or the macroscopic jump of the sample itself can cause difficulties in experiments. For instance, in ref. [2] it was observed that the DSC peak of the thermal recovery showed an anomalous shift to lower temperatures with increasing heating rates. In [7] it was mentioned that the slow motion of the crosshead of the machine did not go down so fast and a stress drop could be registered or that the testing machine can be slightly deformed itself. However, in these papers the above effects were carefully handled, and it can be concluded that the observations summarized above on the anomalous stress–strain curves and on the burst-like recovery are real effects.
In fact, the above martensite stabilization is similar to the effect of the so-called SIM aging, where the stabilization is achieved by aging under uniaxial stress after the formation of stress induced martensite [11,12,13,14,15,16,17,18]. While the SIM aging effect is well interpreted by the symmetry confirming short range ordering [14,19,20], the interpretation of the anomalous shape of the stress–strain curves and the burst-like strain recovery still is in infancy [1,3,4,9] and mostly qualitative arguments were offered. There are only few articles in which more quantitative treatment can be found from two groups [1,3,21,22,23]. In [1,3,21] the so-called theory of diffuse martensitic transformation is used, which involves a second order phase transformation. The models used in [22,23] are based on first order transformation and explain the anomalous σ~ε curves only for a transformation between the austenite and one martensite modification. It was obtained in [22] that the slope of the upper branch of the σ~ε curve can be negative only if S A S M > 0 , where SA and SM are the stiffness of the austenite and martensite, respectively. However, this conclusion contradicts to a set of experimental results where anomalous stress–strain curves were observed for S A S M < 0 [3,7,9].
Thus, the main subject of our paper is to provide a simple thermodynamic analysis, based on our local equilibrium formalism [24,25]. In contrast to [21], we investigate the stability conditions always for a first order transformation, which is in line with [22,23]. It will be shown that the uploading stress–strain curve always has positive slope if only one martensite forms and grows smoothly (i.e., the martensite volume fraction approximately continuously increases). Negative slope can be obtained for smooth transformations if there is a growth/competition of two martensite structural modifications/variants and if the second (more stable) one is the final product. Furthermore, local stress drops on the σ~ε curve can appear if the nucleation of the second martensite is difficult. We hope that our results provide an important, new contribution to the understanding of the above frequently observed experimental behaviour, which can also be manifested in burst-like temperature induced strain recovery of the stabilized martensite structure. Such martensite stabilizations are very useful in producing two-way shape memory behaviour and rubber-like behaviour with extended recoverable strains, important in many applications.

2. Model Calculations

2.1. The Local Equilibrium Formalism

Let us start from our local equilibrium formalism summarized in [25] and used in [24]. In local equilibrium the derivative of the total change of the Gibbs free energy per unit volume, Δ G , versus the transformed martensite volume fraction, ξ, should be equal to zero. For the forward, (from austenite to martensite) transformation,
Δ G ξ = Δ G c + E + D ξ = Δ G c ξ + e ξ + d ξ = 0 .
Here, ξ = V M V M + V A with V = V A + V M (V is the volume of the sample), E as well as D denote the elastic and dissipative energies per unit volume, belonging to the AM (AM) transformation and e ξ = d E ξ d ξ as well as d ξ = d D ξ d ξ , respectively. The total dissipative energy, D t = 0 1 d ξ d ξ , is obviously positive in both directions ( D t A M = D t D t M A > 0 ) and, if the thermoelastic balance is also assumed, then E t A M = E t = 0 1 e ξ d ξ = E t M A > 0 . Usually, there is one additional term in the expression of ΔG, which is the nucleation energy related to the formation of new interfaces during the transformation. Since this is always positive in both directions, like the dissipative energy, it can be considered as it would be included in the dissipative term [24,25].
In Equation (1) ∆Gc is the change in the chemical free energy per unit volume and its derivative can be written as
Δ G c ξ = ξ G M + 1 ξ G A G A ξ = ξ G M G A ξ = G M G A + ξ G M G A   ξ
where
( G M G A ) = u s ε tr
Here ∆s = sMsA (∆s < 0) as well as Δu = uMuA = Δu (<0) are the entropy and internal energy changes per unit volume, respectively, and they are independent of ξ. T and σ are the temperature and stress as well as εtr is the transformation strain, which is positive for tension.
In the literature it is frequently assumed that the transformation strain, εtr, is constant and independent of ξ (i.e., the derivative of Equation (3) is zero). However, in general εtr can depend on the T and/or σ values since in Equation (3) the stress-term has a tensor character [24,25]. Thus, even for the application of uniaxial stress (which leads to a scalar term, as in Equation (3)), εtr in principle can depend on T and σ too. In case of formation and growth of two martensite structural modifications, it can also depend on the volume fraction of one of these, η(ξ) = η(T,σ) = VM2/VM (VM = VM1 + VM2, V = VM + VA), [24,25,26]. The η dependence of εtr can be given by the relation [24,25,26]:
εtr(η) = ε1 + (ε2ε1)η
where ε1 and ε2 are the transformation strains when fully one certain martensite structure forms (at η ≅ 0 and η ≅ 1, respectively). Of course, the details can be very complex and the value of ε can also be different for twinned or detwinned martensite variants [27,28,29]. As a consequence, η can have a ξ-dependence, η(ξ), which explains the ξ-dependence of εtr: εtr(η(ξ)). Furthermore, εi, due to the different temperature dependence of the elastic moduli of the austenite and martensite phases, can have a direct temperature dependence too [22,30,31]). Thus, considerations on the ξ-dependence of the actual strain, ε ξ (which should be distinguished from the transformation strain, ε t r can be important for estimation of the σ(ε) curve (see also below). It is worth mentioning that our description, in order to concentrate on the interpretation of the main important features, in its form is a simplified one, and is applicable indeed for single crystals (with two martensite modifications), where the effects summarized in the introduction were observed. Of course, for more complex systems (e.g., for polycrystalline materials with numerous martensite variants) one would need a more sophisticated treatment, such as the phase field method in the form of the well-known Ginzburg–Landau theory (see e.g., [32]).

2.2. Expressions for the σ ξ σ ε Functions

In contrast to [24,25], we will assume that εtr depends on ξ, and we have from (2) and (3)
Δ G c ξ = Δ u T Δ s σ ε t r ξ σ ε t r ξ = Δ u T Δ s σ ε t r 1 + ξ ε t r ε t r ξ
Since a similar relation holds for Δ G c M A 1 ξ ( 1 ξ is the austenite volume fraction) belonging to the reverse (martensite to austenite, MA) transformation [24,25]. In the following we give detailed considerations only for the forward transformation, and expressions with upper indexes MA as well as AM will be used only if a comparison of the forward and reverse transformations is made.
Taking Equation (5) equal to zero for T = const.
σ 0 T , ξ = 1 ε t r T 1 + ξ ε t r ε t r ξ Δ u T Δ s = σ o 0 , ξ T Δ s ε t r T 1 + ξ ε t r ε t r ξ
Here σ o 0 , ξ = Δ u ε t r T = 0 1 + ξ ε t r ε t r ξ   ( Δ u < 0 , Δ s     0 ) is the equilibrium transformation stress (at T = 0 ). If the transformation strain is independent of ξ, the usual form of the well-known Clausius–Clapeyron equation is obtained [24,25], i.e., σ o T = σ o 0 Δ s ε t r T .
Furthermore, from the condition (1), using Equations (5) and (6) as well, we obtain the expression for the forward branch of the σ (ξ) function (at a fixed value of T) as
σ T , ξ = σ o T , ξ + e ξ + d ξ ε t r T 1 + ξ ε t r ε t r ξ
(see also [24,25] for constant εtr).
The start and finish stresses, for both the forward and reverse transformations, can be given as
σ M s T , ξ = 0 = σ o T , 0 + e o + d o ε t r
σ M f T , ξ = 1 = σ o T , 1 + e 1 + d 1 ε t r 1 + 1 ε t r ε t r ξ ξ = 1
σ A s T ,   1 ξ = 0 = σ o T , 1 ξ = 0 + e 1 M A + d 1 M A ε t r M A
σ A f T ,   1 ξ = 1 = σ o T , 1 ξ = 1 + e o M A + d o M A ε t r M A   1 + 1 ε t r M A ε t r M A ε t r M A 1 ξ 1 ξ = 1

3. Discussion

3.1. Expressions for the Widths of Transformations—Investigation of the Stability during Phase Transformation

Let us consider the sign of the second derivative of the difference of the Gibbs free energy: this gives information about the stability of the system against fluctuations in the volume fraction during growth. If it is positive the two-phase system is stable during the transformation. From Equations (1), (4) and (5) we have, for the austenite to martensite, AM, transformation
2 Δ G ξ 2 T = 2 Δ G c ξ 2 + e ξ + d ξ = σ 2 ε t r ξ + ξ 2 ε t r ξ 2 + e ξ + d ξ .
It can be seen that for a detailed stability analysis we have to discuss the meaning of the terms on the right had side of Equation (9).

3.1.1. Meaning of the Elastic and Dissipative Terms

During thermoelastic AM transformations the elastic energy, E(ξ) > 0, is due to the local strain fields around the martensite nuclei formed and due to the overlap of the elastic fields: the latter contribution can be proportional to ξ 2 [24,25]. Thus, one can assume that
E ξ ξ = e ξ = e o + e 1 e o ξ > 0
and the elastic contribution to Equation (9) is
e ξ = e 1 e o     0
For the reverse transformation the stored elastic energy is released, i.e., e M A ξ = e M A 1 ξ , and we can similarly write
e M A 1 ξ = e o M A + e 1 M A e o M A 1 ξ < 0
and
e M A 1 ξ = e 1 M A e o M A = e 1 e o     0
The dissipative energy, D(ξ), is also positive for the forward transformation and can be considered as the sum of two terms, D = D f + D n . Df (>0) originates from the frictional-type motion of the interfaces and can be supposed that it is proportional to ξ, while D n (>0) is due to the nucleation energy. In the simple case when a large number of martensite nuclei form (smooth transformation [22])   D n can also be approximately a monotonic linear function of ξ and thus
d ξ = d o = d 1 = c o n s t .   > 0
Refs. [24,25] and its contributions to (9) is zero. In a more general case, e.g., if the nucleation is difficult and happens suddenly at certain temperatures, d ξ = 2 D ξ 2     0 , and Dn(ξ) can be a complicated (step-wise) function of ξ (see also below).

3.1.2. Stress–Strain Loops

First consider smooth thermoelastic transformations, when only one type of martensite structural modification grows, i.e., ε t r is independent of ξ ( ε t r = c o n s t . ) and Δ s   < 0 = Δ s M A [24,25]. Then, one gets from Equation (9) with Equation (7) that
2 Δ G ξ 2 T = σ ξ ε t r = e ξ + d ξ     e 1 e o = ( σ M f σ M s ) ε t r
where Equations (11) and (13) were also used. Note, that in this case the second derivative of ΔG is proportional to the slope of the σ(ξ) function and ( σ M f σ M s ) is the width of the upper branch of the schematic hysteresis loop shown in Figure 2. Since, in accordance with Equation (11) e 1 e o     0 , it is usually positive, or close to zero (for horizontal branches), the stability condition fulfils for this simple case. Similarly, we can deduce for the down branch (using also that ε t r = ε t r M A and Equation (12))
2 Δ G M A 1     ξ 2 T = σ M A 1     ξ ε t r M A     e M A 1     ξ = e M A ξ = e 1     e o       σ A s     σ A f ε t r
for which the stability condition also fulfils, since σ A s     σ A f     0 (Figure 2).

3.1.3. Thermal Hysteresis Loops

Under similar conditions as (14) was obtained we can write for the cooling as well as for the heating branches of the thermal hysteresis loop (i.e., for the ξ T function) [24] that
2 Δ G ξ 2 σ = 0 = 2 E T + D T ξ 2     e T ξ       e 1     e o     M s     M f Δ S
as well as
2 Δ G M A 1     ξ 2 σ = 0 e T M A 1     ξ e 1 T M A     e o T M A = e 1 T     e o T = A f     A S     Δ s > 0
The lower indexes T indicate that the derivatives of the elastic (and dissipative) energies can be different for the ξ(T) and σ(ξ) functions (see also below). It is clear, that since both M s     M f and A f A S are positive the thermal hysteresis loops are always stable. It is worth noting that Equations (16) and (17) are also valid if ε t r is not constant, since at σ = 0 the term σ 2 ε t r ξ + ξ 2 ε t r ξ 2 , which is present in Equation (9), cancels out.

3.2. Growth of Two Martensite Modifications

Let us now consider the more general case when ε t r has ξ-dependence as given by Equation (4) with η(ξ). It is clear from Equation (9) that, if the first term is negative and its magnitude is larger than the last two terms, the system can be unstable. Using Equation (4) we can write
ε t r ξ = ε 2     ε 1 η ξ
and the first term of Equation (9) will be
σ ε 2     ε 1 2 η ξ + 2 η ξ 2
This term can be negative if ε 2 > ε 1 and the second martensite grows monotonically with ξ (smooth transition). Of course, in general, the situation can be more complex. For example, during nucleation and growth of two martensite modifications, when the first one of them grows and suddenly the second one starts to nucleate and grow, there can be sudden changes in the stored elastic and nucleation energies (the nucleation energy is included in D). Thus, in the second derivatives of E and D there can also appear sudden sign changes (jumps), which can be quite local at around a certain value of ξ. Another complication is involved in the fact that, in this general case, the martensite volume fraction is not proportional to the actual strain.
Nevertheless, the slope 2 Δ G ξ 2 | T (see Equation (9)) can be negative if the magnitude of the term given by Equation (19) is larger than the value e ξ + d ξ e 1     e o , and then the overall slope of the uploading branch of the σ(ε) curve can be negative, i.e., the two-phase system can be unstable during the transition. For burst-like recovery, when A s     A f     0 , e ξ + d ξ     e 1     e o is close to zero, and in this case the condition of instability can be given as σ ε 2 ε 1 2 η ξ + 2 η ξ 2 < 0 .

3.3. Nucleation Difficulties

Let us consider the following simple case: assume that with increasing stress the M1 martensite modification nucleates first and grows and at a certain transformed volume fraction, ξc (~εc) the second martensite M2 nucleates. Suppose that M2 is more stable than M1, in accordance with the higher temperature of burst-like MA transformation and, as above, ε 2 = ε t r 2 > ε 1 = ε t r 1 . Under the same assumptions as Equation (14) was obtained, we can write for the slopes of the linearized σ 1 ξ and σ 2 ξ functions
( σ M f 1 σ M s 1 ) = e 11 e o 1 ε t r 1     0
and
( σ M f 2 σ M s 2 ) = e 12 e o 2 ε t r 2     0
where relations (7) and (8) were also used and, writing Equation (21), we assumed that for a burst-like recovery of M2 the slope is approximately zero. Thus, the slope of σ 1 ξ is larger. In addition, consider the case when σ M s 1 < σ M s 2 i.e., if
σ o 1 T + e 01 + d o 1 ε 1 < σ o 2 T + e 02 + d o 2 ε 2
which can be fulfilled if do2 is large enough. Indeed, it can be the case if we assume that the nucleation of M2 is more difficult than the nucleation of M1. For instance, it is well-known in the literature [33,34] that there can be a competition between habit-plane variants, which can easily nucleate from austenite (and accompanied with smaller transformation strain), and oriented martensites. The nucleation of the latter one is more difficult because of crystallographic compatibility problems, depending on the orientation relationships, and the formation and growth of which can be accompanied with smaller accumulated stresses (see e.g., [9,34,35,36]). The formation of M2 at a certain critical stress/strain can happen either from the not transformed yet austenite or from the growing M1 (see e.g., [34]).
Figure 3 shows schematically the σ 1 ξ and σ 2 ξ functions (see also Equations (7) and (8)). The intersection of the two straight lines gives the value of ξc(~εc), at which σ ξ c = σ M s 2 , and thus M2 can nucleate which leads to a stress drop. For the estimation of the stress drop we can use the results of [22]. It was obtained, for the stress induced austenite to one-martensite-modification transformation, that (due to the transformation strain) there should exists an overall decrease in stress. This decrease, as compared to the pure elastic contribution to the stress–strain curve, is proportional to the product of the martensite volume fraction, ξ , the transformation strain, ε t r , and the effective stiffness in the two-phase region, S(ξ), as compared to the pure elastic contribution to the stress–strain curve: Δ σ   ξ ε t r S ξ ( ξ corresponds to the parameter ξ     n z ε N y used in [22] and S ξ = S A S M S M + ξ S A S M ). Thus, the stress drop due to the nucleation of M2 martensite at ξc is proportional to ε 2 t r ε 1 t r :
Δ σ   ξ c S ξ c ε 2 t r ε 1 t r .
Now, if the magnitude of Δ σ is large enough to decrease the stress below the M1 start stress then after the stress drop, the austenite + M1 two-phase system will be elastically deformed until the nucleation of M1 martensite happens again and the process is repeated. In this (i) case the slope after the stress drop will be the similar as the initial slope of the stress–strain curve, with the effective stiffness S(ξc). In the opposite case, (ii), after the stress drop there is a further growth of M1 (with a similar course of the stress–strain curve as observed during the first stage of the process before the stress-drop) until a new M2 nucleates and the process is repeated. Both can be observed in experiments: (i) see Figure 1b of [2] where the segment of the σ~ε curve after the stress-drop is similar to the initial part, or (ii) in Figure 1a,b in [1] as examples. Figure 1 in [1] also illustrates that with decreasing test temperature there is a transition from (i) to (ii). Regarding the unloading process under compression, with decreasing stress the M2 is stable well below the austenite start stress of M1, σ a s 1 , (even below a certain temperature M2 remains stable at zero stress). Thus, retwinning of M2 (i.e., the formation of M1 from M2) can start at a certain low stress level and this leads to the expansion of the sample ( ε 2 > ε 1 ) and a stress jump can appear (see Figure 1b). Further details, like the course of the curve during the stress-drops before reaching the minimum, or where and how M2 nucleates can only be explained on the basis of microscopic investigations (see e.g., [35,36,37]) and out of the scope of thermodynamic considerations used here.
Thus, we demonstrated that if the nucleation of the second martensite is difficult, then the interplay of the elastic and dissipative contributions to the σ(ξ) curve, can be such which can explain the formation of several steps on the σ(ε) plots. Of course, during the local stress drops the system is also unstable. It can also happen, if e.g., there are only two stress drops, which are frequently observed [2,3], that the presence of these alone can result in an overall negative slope of the stress–strain curves, i.e., the negative slope can be produced by pure nucleation difficulties alone.
It can be mentioned that experimental results on superelastic behaviour of Fe based alloys sometimes can also show stress drops on the loading branch of the stress–strain curves (see e.g., Figure 6a in [38] or Figure 2d in [39]) and these show quite strong differences in compression and tension. The interpretation of these needs a deeper analysis than the one presented above, where the elastic energy accumulation during martensite formation is handled mainly by its difference when twinned or detwinned martensite modifications grow. In the presence of precipitates and/or retaining martentsites, and when the stress level can be high enough, dislocations can also play a role in the elastic energy accumulation/relaxation as well as in the nucleation barriers and the corresponding microstructure evolution can be quite complex. The better understanding of these phenomena is still in infancy and call for further efforts.

3.4. About the Burst-like Recovery

The shift of the forward transformation peak of the burst-like thermal recovery, as suggested in [12], can happen through re-twinning of M2: once the twinning is achieved an “easy way” is opened for transformation to the M1 or austenite phase(s). Thus, the observed overheating is the consequence of the (nucleation) barrier against re-twinning. We can obtain an approximate expression for the shift of the forward transformation peak of the burst-like recovery as compared to the same thermally induced DSC peak, ΔT, using the results described in the preceding sections. ΔT, can be given as the difference of the peak temperatures, Tp, during heating:
Δ T = T p 1 T p 2 = A f 2 + A s 2 2 Δ s 2 A f 1 + A s 1 2 Δ s 1 = Δ T o + d 12 + d 02 + e 12 + e 02 2 Δ s 2 d 11 + d 01 + e 11 + e 01 2 Δ s 1
where we used that the start and finfish temperatures can be given from (5) at σ = 0 (see also [24]. Since the burst-like peak belongs to M2/A transformation and for this e 12 e o 2     0 (see (21)), as well as for two martensite variants we can also assume that the difference of the equilibrium transformation temperatures ( Δ T o = T o 2 T o 1 ) and the transformation entropy is approximately zero ( Δ s 1     Δ s 2 = Δ s ):
Δ T     2 d o 2 d 01 + 2 e o 2 e 11 + e 01 2 Δ s
where it was also used (as in Equations (20) and (21)), that d 1 i     d 0 i (i = 1, 2). In accordance with the previous considerations, we can assume that the difference of the elastic terms is small (and even can be negative, if e 11 + e o 1 > 2 e o 2 since e 11 + e o 1 > 0 ) as compared to the first term. Thus, since ΔT > 0, the first term dominates in Equation (25). Furthermore, it can also be used that the main difference in the first term is due to the difference of the nucleation energy, i.e., d o 2 d 01 = Δ d n > 0 , and thus
Δ T     Δ d n Δ s = D 2 D 1 Δ s
where the relation D i = 0 1 d o i d ξ = d o i (see Equation (13) too) was used and Di is the nucleation energy for the heating process. Thus, the shift of the transformation peak is a measure of the change in the nucleation energy: Δ D = D 2 D 1 . As a numerical example we can use the results of [40], where it was obtained that ΔT was about 36 K and using the transformation entropy in this alloy, 0.75 J/mol K [40], we find that D 2 D 1     27   J / mol . This can be compared to the half of the dissipative energy per one thermal cycle, 2.8 J/mol [41], which shows that the nucleation energy for M2 can be about an order of magnitude larger than for M1.

4. Conclusions

The second derivatives of the total Gibbs-free energy, in the framework of a local equilibrium description, were investigated for shape memory alloys showing a burst-like shape recovery after anomalous stress–strain load. We have shown that the thermal hysteresis loops are usually stable and an approximately vertical up branch (i.e., AsAf) can be obtained during burst-like thermal recovery, indicating that the second derivative of the elastic energy is approximately zero in this case. It is also shown that the stress–strain loops for smooth transformations are also stable if only one type of martensite growths. Instability can appear, i.e., the overall slope of the A M branch can be negative, if nucleation and growth of two martensite modifications (variants) takes place and if the second (more stable) one is the final product with ε 2 t r > ε 1 t r . It is found that local stress-drops, Δσ, on the stress–strain curve can appear if the nucleation of the second martensite is difficult and the presence of few local stress-drops alone can also result in an overall negative slope of the σ~ε curve. Depending on the magnitude of Δσ, the course of the stress–strain curve after the stress drop is similar to the initial (elastic) part or to the part belonging to A/M1 transformation. It is illustrated that shift of the temperature of the thermal recovery of M2 is a measure of the change in the nucleation energy Δ D = D 2 D 1 .

Author Contributions

Conceptualization, D.L.B. and L.D.; methodology, L.Z.T.; validation, D.L.B., L.D. and L.Z.T., writing—original draft preparation, S.M.K., writing—review and editing, D.L.B. and S.M.K. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Research, Development and Innovation Office: NKFIH PD131784 project.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Nikolaev, V.; Malygin, G.; Averkin, A.; Stepanov, S.; Zograf, G. Anomalous stress-strain behaviour in Ni49Fe18Ga27Co6 crystals compressed along [110]. Mater. Today Proc. 2017, 4, 4807–4813. [Google Scholar] [CrossRef]
  2. Nikolaev, V.; Stepanov, S.; Yakushev, P.; Krymov, V.; Kustov, S. Burst-like shape recovery and caloric effects in Ni–Fe–Ga–Co single crystalline shape memory alloys. Intermetallics 2020, 119, 106709. [Google Scholar] [CrossRef]
  3. Malygin, G.A.; Nikolaev, V.I.; Krymov, V.M.; Pul’Nev, S.A.; Stepanov, S.I. Interfacial Stresses and Anomalous Shape of Pseudoelastic Deformation Curves in Ni49Fe18Ga27Co6 Alloy Crystals Compressed along the [011]A Axis. Tech. Phys. 2019, 64, 819–827. [Google Scholar] [CrossRef]
  4. Nikolaev, V.I.; Yakushev, P.N.; Malygin, G.A.; Averkin, A.I.; Pulnev, S.A.; Zograf, G.; Kustov, S.B.; Chumlyakov, Y.I. Influence of partial shape memory deformation on the burst character of its recovery in heated Ni–Fe–Ga–Co alloy crystals. Tech. Phys. Lett. 2016, 42, 399–402. [Google Scholar] [CrossRef]
  5. Yang, S.; Omori, T.; Wang, C.; Liu, Y.; Nagasako, M.; Ruan, J.; Kainuma, H.; Ishida, K.; Liu, X. A jumping shape memory alloy under heat. Sci. Rep. 2016, 6, 21754. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Zhao, D.; Xiao, F.; Nie, Z.; Cong, D.; Sun, W.; Liu, J. Burst-like superelasticity and elastocaloric effect in [011] oriented Ni50Fe19Ga27Co4 single crystals. Scr. Mater. 2018, 149, 6–10. [Google Scholar] [CrossRef]
  7. Picornell, C.; Pons, J.; Cesari, E. Stress–temperature relationship in Cu–Al–Ni single crystals in compression mode. Mater. Sci. Eng. A 2004, 378, 222–226. [Google Scholar] [CrossRef]
  8. Picornell, C.; Pons, J.; Cesari, E. Stress-Temperature Relationship in Compression Mode in Cu-Al-Ni Shape Memory Alloys. Mater. Trans. 2004, 45, 1679–1683. [Google Scholar] [CrossRef] [Green Version]
  9. Lazpita, P.; Villa, E.; Villa, F.; Chernenko, V. Temperature Dependent Stress–Strain Behavior and Martensite Stabilization in Magnetic Shape Memory Ni51.1Fe16.4Ga26.3Co6.2 Single Crystal. Metals 2021, 11, 920. [Google Scholar] [CrossRef]
  10. Nikolaev, V.I.; Yakushev, P.N.; Malygin, G.A.; Pul’nev, S.A. Burst character of thermoelastic shape memory deformation in ferromagnetic Ni-Fe-Ga-Co alloy. Tech. Phys. Lett. 2010, 36, 914–917. [Google Scholar] [CrossRef]
  11. Picornell, C.; Pons, J.; Cesari, E. Stabilisation of martensite by applying compressive stress in Cu-Al-Ni single crystals. Acta Mater. 2001, 49, 4221–4230. [Google Scholar] [CrossRef]
  12. Niendorf, T.; Krooß, P.; Somsen, C.; Eggeler, G.; Chumlyakov, Y.I.; Maier, H.J. Martensite aging-Avenue to new high temperature shape memory alloys. Acta Mater. 2015, 89, 298–304. [Google Scholar] [CrossRef]
  13. Chernenko, V.A.; Pons, J.; Cesari, E.; Zasimchuk, I.K. Transformation behaviour and martensite stabilization in the ferromagnetic Co-Ni-Ga Heusler alloy. Scr. Mater. 2004, 50, 225–229. [Google Scholar] [CrossRef]
  14. Kadletz, P.M.; Krooß, P.; Chumlyakov, Y.I.; Gutmann, N.J.; Schmahl, W.W.; Maier, H.J.; Niendorf, T. Martensite stabilization in shape memory alloys-Experimental evidence for short-range ordering. Mater. Lett. 2015, 159, 16–19. [Google Scholar] [CrossRef]
  15. Panchenko, E.; Timofeeva, E.; Larchenkova, N.; Chumlyakov, Y.; Tagiltsev, A.; Maier, H.; Gerstein, G. Two-way shape memory effect under multi-cycles in [001]-oriented Ni49Fe18Ga27Co6 single crystal. Mater. Sci. Eng. A 2017, 706, 95–103. [Google Scholar] [CrossRef]
  16. Gerstein, G.; L’vov, V.A.; Kosogor, A.; Maier, H.J. Internal pressure as a key thermodynamic factor to obtain high-temperature superelasticity of shape memory alloys. Mater. Lett. 2018, 210, 252–254. [Google Scholar] [CrossRef]
  17. Panchenko, E.; Timofeeva, E.; Eftifeeva, A.; Osipovich, K.; Surikov, N.; Chumljakov, Y.; Gerstein, G.; Maier, H.J. Giant rubber-like behavior induced by martensite aging in Ni51Fe18Ga27Co4 single crystals. Scr. Mater. 2019, 162, 387–390. [Google Scholar] [CrossRef]
  18. Ahlers, M.; Pelegrina, J.L. Ageing of martensite: Stabilisation and ferroelasticity in Cu-based shape memory alloys. Mater. Sci. Eng. A 2003, 356, 298–315. [Google Scholar] [CrossRef]
  19. Ren, X.; Otsuka, K. Origin of rubber-like behaviour in metal alloys. Nature 1997, 389, 579–582. [Google Scholar] [CrossRef]
  20. Otsuka, K.; Ren, X. Mechanism of martensite aging effects and new aspects. Mater. Sci. Eng. A 2001, 312, 207–218. [Google Scholar] [CrossRef]
  21. Malygin, G.A. Diffuse martensitic transitions and the plasticity of crystals with a shape memory effect. Phys.-Uspekhi 2001, 44, 173–197. [Google Scholar] [CrossRef]
  22. L’Vov, V.A.; Rudenko, A.A.; Chernenko, V.A.; Cesari, E.; Pons, J.; Kanomata, T. Stress-induced Martensitic Transformation and Superelasticity of Alloys: Experiment and Theory. Mater. Trans. 2005, 46, 790–797. [Google Scholar] [CrossRef] [Green Version]
  23. L’Vov, V.; Matsishin, N.; Glavatska, N.; Matsishin, M. Thermoelastic behaviour of martensitic alloy in the vicinity of critical point in the stress–temperature phase diagram. Phase Transit. 2010, 83, 293–301. [Google Scholar] [CrossRef]
  24. Beke, D.L.; Daroczi, L.; Samy, N.M.; Toth, L.Z.; Bolgar, M.K. On the thermodynamic analysis of martensite stabilization treatments. Acta Mater. 2020, 200, 490–501. [Google Scholar] [CrossRef]
  25. Beke, D.L.; Daróczi, L.; Elrasasi, T.Y. Determination of Elastic and Dissipative Energy Contributions to Martensitic Phase Transformation in Shape Memory Alloys. In Shape Memory Alloys-Processing, Characterization and Applications; Fernandes, F.M.B., Ed.; InTech: Rijeka, Croatia, 2013; Chapter 7; pp. 167–196. ISBN 978-953-51-1084-2. [Google Scholar]
  26. Beke, D.L.; Daroczi, L.; Palanki, Z.; Lexcellent, C. Proceedings of the International Conference on Shape Memory and Superleastic Technologies, Tsukuba, Japan, 3–5 December 2007; Miyazaki, S., Ed.; ASM International: Materials Park, OH, USA, 2008; pp. 607–614. [Google Scholar]
  27. Hamilton, R.; Efstathiou, C.; Sehitoglu, H.; Chumljakov, Y. Thermal and stress-induced martensitic transformations in NiFeGa single crystals under tension and compression. Scr. Mater. 2006, 54, 465–469. [Google Scholar] [CrossRef]
  28. Efstathiou, C.; Sehitoglu, H.; Carroll, J.; Lambros, J.; Maier, H.J. Full-field strain evolution during intermartensitic transformations in single-crystal NiFeGa. Acta Mater. 2008, 56, 3791–3799. [Google Scholar] [CrossRef]
  29. Panchenko, E.Y.; Timofeeva, E.; Chumlyakov, Y.I.; Osipovich, K.; Tagiltsev, A.; Gerstein, G.; Maier, H. Compressive shape memory actuation response of stress-induced martensite aged Ni51Fe18Ga27Co4 single crystals. Mater. Sci. Eng. A 2019, 746, 448–455. [Google Scholar] [CrossRef]
  30. Liu, Y.; Yang, H. The concern of elasticity in stress-induced martensitic transformation in NiTi. Mater. Sci. Eng. A 1999, 260, 240–245. [Google Scholar] [CrossRef]
  31. Panchenko, E.; Chumlyakov, Y.; Eftifeeva, E.; Maier, H.J. Two-way shape memory effect in ferromagnetic Co35Ni35Al30 single crystals aged under stress. Scipta Mater. 2014, 90–91, 10–13. [Google Scholar] [CrossRef]
  32. Xi, S.; Su, Y. A phase field study ot the grain size effect on the thermomechanical behavior of polycrystalline NiTi thin film. Acta Mech. 2021, 232, 4545–4566. [Google Scholar] [CrossRef]
  33. Christian, J.W. Deformation of moving interfaces. Met. Mater. Trans. A 1982, 13, 509–538. [Google Scholar] [CrossRef]
  34. Basinski, Z.S.; Christian, J.W. Experiments on the martensitic transformation in single crystals of indium-thallium alloys. Acta Met. 1954, 2, 148–166. [Google Scholar] [CrossRef]
  35. Karaca, H.H.; Karaman, I.; Basaran, B.; Lagoudas, D.C.; Chumljakov, Y.I.; Maier, H.J. On the stress-assisted magnetic-field-induced phase transformation in Ni2MnGa ferromagnetic shape memory alloys. Acta Mater. 2007, 55, 4253–4269. [Google Scholar] [CrossRef]
  36. Delpueyo, D.; Crediac, M.; Balardraud, X.; Badulescu, C. Investigation of martensite structures in a monocrystalline Cu-Al-Be shape memory alloy with the grid method and infrared thermography. Mech. Mater. 2012, 45, 34–51. [Google Scholar] [CrossRef]
  37. Ball, J.M.; Koumatos, K.; Seiner, H. Nucleation of austenite in mechanically stabilized martensite by localized heating. J. Alloys Compd. 2013, 577, S37–S42. [Google Scholar] [CrossRef] [Green Version]
  38. Tseng, L.W.; Ma, J.; Wang, S.J.; Karaman, I.; Kaya, M.; Luo, Z.P.; Chumljakov, Y.I. Superelastic response of a single crystalline FeMnAlNi shape memory alloy under tension and compression. Acta Mater. 2015, 89, 374–383. [Google Scholar] [CrossRef]
  39. Chumlyakov, Y.; Kireeva, I.; Kuksgauzen, I.; Kuksgauzen, D.; Niendorf, T.; Krooβ, P. Tension-compression assymmetryof the superelastic behavior of high-strength [001]-oriented FeNiCoAlNi crystals. Mater. Lett. 2021, 289, 129395. [Google Scholar] [CrossRef]
  40. Kamel, S.M.; Daróczi, L.; Tóth, L.Z.; Panchenko, E.; Chumljakov, Y.I.; Beke, D.L. Burst like recovery of Ni49Fe18Ga27Co6 shape memory single crystal after compressive stress-strain loading. In Proceedings of the 12th European Symposium on Martensitic Transformations, Ankara, Turkey, 5–9 September 2022; Kockar, B., Gortan, M.O., Eds.; Hacettepe University: Ankara, Turkey; pp. 88–89. [Google Scholar]
  41. Bolgár, M.K.; Tóth, L.Z.; Szabó, S.; Gyöngyösi, S.; Daróczi, L.; Panchenko, E.Y.; Chumlyakov, Y.I.; Beke, D.L. Thermal and acoustic noises generated by austenite/martensite transformation in NiFeGaCo single crystals. J. Alloys Compd. 2016, 658, 29–35. [Google Scholar] [CrossRef]
Figure 1. Stress–strain curves schematically: (a) typical loop with positive slope of the uploading branch, (b) anomalous loop with negative overall slope and stress drops on it.
Figure 1. Stress–strain curves schematically: (a) typical loop with positive slope of the uploading branch, (b) anomalous loop with negative overall slope and stress drops on it.
Materials 15 09010 g001
Figure 2. Schematic σ(ξ) hysteresis loop for εtr = const. > 0. Since ε = ξεtr, with the actual value of ε, this plot corresponds to the σ versus ε plots.
Figure 2. Schematic σ(ξ) hysteresis loop for εtr = const. > 0. Since ε = ξεtr, with the actual value of ε, this plot corresponds to the σ versus ε plots.
Materials 15 09010 g002
Figure 3. Schematic σ 1 ξ and σ 1 ξ functions, and the stress drop, Δ σ at the critical volume fraction where the second martensite nucleates (see also the text).
Figure 3. Schematic σ 1 ξ and σ 1 ξ functions, and the stress drop, Δ σ at the critical volume fraction where the second martensite nucleates (see also the text).
Materials 15 09010 g003
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Beke, D.L.; Kamel, S.M.; Daróczi, L.; Tóth, L.Z. Thermodynamic Analysis of Anomalous Shape of Stress–Strain Curves for Shape Memory Alloys. Materials 2022, 15, 9010. https://doi.org/10.3390/ma15249010

AMA Style

Beke DL, Kamel SM, Daróczi L, Tóth LZ. Thermodynamic Analysis of Anomalous Shape of Stress–Strain Curves for Shape Memory Alloys. Materials. 2022; 15(24):9010. https://doi.org/10.3390/ma15249010

Chicago/Turabian Style

Beke, Dezső L., Sarah M. Kamel, Lajos Daróczi, and László Z. Tóth. 2022. "Thermodynamic Analysis of Anomalous Shape of Stress–Strain Curves for Shape Memory Alloys" Materials 15, no. 24: 9010. https://doi.org/10.3390/ma15249010

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