Next Article in Journal
Effect of Build Orientation on the Microstructure, Mechanical and Corrosion Properties of a Biodegradable High Manganese Steel Processed by Laser Powder Bed Fusion
Next Article in Special Issue
Molecular Dynamics Study of Bulk Properties of Polycrystalline NiTi
Previous Article in Journal
Fabricating Homogeneous FeCoCrNi High-Entropy Alloys via SLM In Situ Alloying
Previous Article in Special Issue
Role of Nickel Content in One-Way and Two-Way Shape Recovery in Binary Ti-Ni Alloys
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Shape Memory Effect and Superelasticity of [001]-Oriented FeNiCoAlNb Single Crystals Aged under and without Stress

by
Yuriy I. Chumlyakov
1,
Irina V. Kireeva
1,*,
Zinaida V. Pobedennaya
1,
Philipp Krooß
2 and
Thomas Niendorf
2
1
Siberian Physical Technical Institute, National Research Tomsk State University, 634050 Tomsk, Russia
2
Institute für Werkstofftechnik, Universität Kassel, 34125 Kassel, Germany
*
Author to whom correspondence should be addressed.
Metals 2021, 11(6), 943; https://doi.org/10.3390/met11060943
Submission received: 14 May 2021 / Revised: 2 June 2021 / Accepted: 7 June 2021 / Published: 10 June 2021
(This article belongs to the Special Issue Structure, Texture and Functional Properties of Shape Memory Alloys)

Abstract

:
The two-step ageing of Fe-28Ni-17Co-11.5Al-2.5Nb (at.%) single crystals under and without stress, leads to the precipitation of the γ′- and β-phase particles. Research has shown that γ–α′ thermoelastic martensitic transformation (MT), with shape memory effect (SME) and superelasticity (SE), develops in the [001]-oriented crystals under tension. SE was observed within the range from the temperature of the start of MT upon cooling Ms, to the temperature of the end of the reverse MT upon heating Af, and at temperatures from Af to 323–373 K. It was found that at γ–α′ MT in the [001]-oriented crystals, with γ′- and β-phase particles, a high level of elastic energy, ΔGel, is generated, which significantly exceeds the energy dissipation, ΔGdis. As a result, the temperature of the start of the reverse MT, while heating As, became lower than the temperature Ms. The development of γ–α′ MT under stress occurs with high values of the transformation hardening coefficient, Θ = dσ/dε from 2 to 8 GPa and low values of mechanical Δσ and thermal ΔTh hysteresis. The reasons for an increase in ΔGel during the development of γ–α′ MT under stress are discussed.

1. Introduction

FeNiCoAlX alloys (X = Ta, Nb, Ti), undergoing thermoelastic martensitic transformation (MT) from the fcc (γ)-phase to bct (α′)-martensite upon cooling/heating and under stress, are characterized by high values of superelasticity (SE) from 5 to 13.5% and values of shape memory effect (SME) of up to 8%; these alloys can compete with TiNi-based alloys [1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34]. The necessary conditions for the development of thermoelastic γ–α′ MT are achieved: (i) due to the precipitation of nanosized γ′-phase particles, with L12-type ordering of between 3 and 8 nm in size, which have coherent lattice conjugation with high-temperature γ-phase and retain coherence during MT in α′-martensite [1,2,3,4,28]; (ii) particles of the γ′-phase strengthen the high-temperature phase, reduce the transformation strain by comparison with the single-phase state [1,5] and contribute to an increase in the tetragonality of martensite c/a to 1.1–1.23 [1,28]. This, in turn, decreases the value of the shear strain, g, at γ–α′ MT and the Burgers vector, b, of twinning dislocations a/6<211>(111), arising during MT [1,2,3,4,5,18].
In the FeNiCoAlX polycrystals (X = Ta, Nb, Ti) during ageing at high temperatures, T ~ 873–973 K, the precipitation of β-phase particles with B2-type ordering at the grain boundaries occurs simultaneously with the precipitation of γ′-phase particles in the grain body, leading to a catastrophic decrease in plasticity [5,31,32]. Alloying with boron to 0.05 at.% suppresses the precipitation of β-phase particles along the grain boundaries, which contributes to the manifestation of SME and SE in the polycrystals of these alloys [5,31,32]. In addition, an increase in the grain size to 100–200 µm and the creation of a sharp <001>{035} texture in polycrystals, as well as the production of oligocrystals, leads to the appearance of SE from 5 to 13.5% in these alloys [5,6,31,32].
Studies of γ–α′ MT on the FeNiCoAlX single crystals (X = Ta, Nb, Ti) have a number of advantages over polycrystals. Firstly, single crystals make it possible to study the orientation dependence and asymmetry under the tension and compression of SME and SE [11,12,13,14,15,17,27]. These results are necessary for the development of micromechanical models in the development of a γ–α′ MT under stress in polycrystals with different textures. Secondly, earlier studies of the FeNiCoAlX single crystals (X = Ta, Nb, Ti) showed that the precipitation of β-phase particles in single crystals, in contrast with the polycrystals, does not lead to a loss of plasticity [5,25,31,32,33,34]. Thirdly, nanosized β-phase particles, which are precipitated in single crystals at 973 K, with an ageing time of t ≥ 7 h, in relation to one- and two-step ageing, lead to an increase in MT temperatures, the retention of SME and SE from 3 to 8%, the appearance of rubber like elasticity and the development of SE in the temperature range from Ms to Af [13,19,33,34].
In the present paper, on the Fe-28Ni-17Co-11.5Al-2.5Nb (FeNiCoAlNb) (at.%) single crystals oriented along the [001] direction, the task was to investigate the temperature dependence of the stresses, σcr(T), required for the onset of stress-induced γ–α′ MT, SE and its temperature range, SME and the value of thermal ΔTh and mechanical Δσ hysteresis, using ageing methods under and without stress under tension. The choice of ageing under and without stress is due to the following. In alloys with coherent non-equiaxial particles, ageing under stress leads to the predominant growth of one variant of particles and the appearance of the internal oriented stresses, which are absent during ageing without stress [35,36,37]. In the case of precipitation of non-coherent and non-equiaxed particles after ageing without stress, internal stresses do not arise due to the formation of dislocations at the “particle-matrix” boundaries [29,30]. The effect of ageing under stress on the precipitation of non-coherent β-phase particles in the FeNiCoAlX alloys (X = Ta, Nb, Ti) has not yet been studied. In the present paper, such experiments using FeNiCoAlNb single crystals oriented along the [001] direction are presented for the first time.

2. Materials and Methods

Ingots of the Fe-28Ni-17Co-11.5Al-2.5Nb (FeNiCoAlNb) (at.%) alloy were melted from pure elements in a resistance furnace (InterSELT, St. Petersburg, Russia) in a helium atmosphere. To achieve a homogeneous distribution of the elements in the bulk of the ingots, they were remelted three times. Single crystals of the FeNiCoAlNb alloy were grown by the Bridgman method in a helium atmosphere, on a Russian-made Redmet installation (Firm “Kristallooptika”, Tomsk, Russia). To determine the orientation of the crystals, the diffractometric method was used by means of a DRON-3M X-ray diffractometer (Bourevestnik, St. Petersburg, Russia) with monochromatic Fe Kα radiation. Dog-bone-shaped tension samples had a gauge length of 12 mm and a cross section of 2 × 1.5 mm2. Samples were cut using wire electrical discharge machining ARTA-5.9 (DELTA-TEST, Fryazino, Moscow region, Russia). The damaged surface layer was ground off mechanically, and then electrically polished in 200 mL of an H3PO4 + 50 g CrO3 (phosphoric acid with chromium trioxide) electrolyte at room temperature. All samples were initially homogenized at 1573 K for 24 h in a quartz tube, in a helium atmosphere, followed by water quenching. Two-step ageing was used to precipitate γ′- and β-phase particles. The first step of ageing was carried out at a temperature of 973 K for 5 h without stress, in an MRU-STF15180-01_301 resistance furnace (Carbolite Gero, Hope Valley, UK) within a quartz tube, in a helium atmosphere with water quenching. The second step of ageing was performed in a vacuum at 873 K, for 2 and 4 h in a special device (Firm “Kristallooptika”, Tomsk, Russia) with a vacuum chamber, able to apply tensile stresses of 120 MPa during ageing, in the following way. Two samples were simultaneously used during each ageing treatment in the second step, with their axes in vertical and horizontal positions. The temperature was monitored by a thermocouple, attached to the vertical sample. The device was pre-heated to 473 K, subsequently, a constant tensile stress of 120 MPa was applied to the vertical sample, whereas the horizontal sample remained in a stress-free condition. Next, the device was heated to 873 K at a rate of 20 K/min and kept for 2 h and 4 h at 873 K. A controlled stress of 120 MPa, applied to the vertical sample, was maintained during all the ageing treatment. The simultaneous ageing of two samples (under and without applied stress) allows for a better comparison of these, to ascertain the effect of the applied external stress, as the samples adhere to exactly the same temperature evolutions during ageing. After completion of the ageing process under stress, the device was slowly cooled to 373 K over 1.5 h and unloaded. Start Ms and finish Mf temperatures of the forward γ–α′ MT during cooling and start As and finish Af temperatures of the reverse γ–α′ MT during heating, were determined by the intersection of tangents on the temperature dependence of the electrical resistance ρ(T) on a Russian-manufactured installation (Firm “Kristallooptika”, Tomsk, Russia). The MT temperatures under stress were determined in terms of deviation from the linear dependence in the “transformation strain-temperature” (εtr(T)) curves. The value of thermal ΔTh and mechanical Δσ hysteresis was determined in the middle of the εtr(T) and “stress-strain” (σ(ε)) hysteresis loops, respectively. Mechanical properties and SE were investigated using an Instron 5969 universal testing machine (Instron, Norwood, MA, USA) at the deformation rate of 4∙10−4 s−1. Tests at room temperature were carried out in air. Tests within the temperature range of 185–350 K were carried out in a special cooling/heating chamber, which is included in the equipment of the Instron 5969 universal testing machine (Instron, Nowood, MA, USA). The heating/cooling rate of the chamber was 2 K/min. The sample was inserted into grips, heated/cooled together with the chamber, kept at each temperature for 30 min before testing, and then deformed. SME at a constant tensile stress in the cycle, was studied using a Russian-made dilatometer (Firm “Kristallooptika”, Tomsk, Russia) during cooling and heating within a temperature range of 77–400 K, with a heating/cooling rate of 10 K/min. Temperature variations at ~10 K/min were achieved by liquid nitrogen, flowing in copper tubes, coiled around the grips and by induction heating of the grips. The sample temperature was measured by a thermocouple, fixed to the sample gauge section. The lowest sample temperature, attainable by the dilatometer, was ~77 K. A miniature extensometer, attached to the sample, was used for the strain measurements. The crystal structure after ageing was investigated using a JEOL-2010 transmission electron microscope (TEM) (JEOL, Tokyo, Japan) at an accelerating voltage of 200 kV. The thin foils were prepared using double-jet electropolishing (TenuPol-5, “Struers”, Ballerup, Denmark), with an electrolyte containing 20% sulfuric acid in methyl alcohol at room temperature, with 12.5 V applied voltage. In this paper, the following denotations are adopted for heat treatments: Crystal I—ageing without stress at 973 K for 5 h and then at 873 K for 2 h without stress and with slow cooling; Crystal II—ageing without stress at 973 K for 5 h and then ageing under stress 120 MPa at 873 K for 2 h with slow cooling; Crystal III—ageing at 973 K for 5 h, then at 873 K for 4 h without stress and with slow cooling; and Crystal IV—ageing without stress at 973 K for 5 h and then ageing under stress 120 MPa at 873 K for 4 h with slow cooling. For each ageing type, namely, Crystals I–IV, SE and SME were investigated on separate samples after determining the MT temperatures.

3. Results and Discussion

3.1. Effect of Ageing under and without Stress on Temperatures of γ–α′ MT and the Precipitation of γ′- and β-Phase Particles

The temperatures of γ–α′ MT, depending on the ageing condition under and without stress, and determined from the temperature dependence of the electrical resistivity ρ(T), are presented in Figure 1 and Table 1. It may be observed that in Crystals III with regard to two-step ageing, an increase in the ageing time at the second step at 873 K from 2 to 4 h without stress leads to an increase in the Ms temperature by 29 K compared to Crystals I. In Crystals II and IV, ageing under stress at the second step leads to an increase in the Ms temperature by 15 and 11 K, respectively, relative to Crystals I and III.
TEM studies (Figure 2) showed that in Crystals I–IV, the γ′- and β-phase particles are precipitated. The γ′-phase particles of the equiaxed shape, with L12-type ordering, which is confirmed by the presence of superstructural reflections (Figure 2c), have the chemical composition, (FeNiCo)3AlNb [5,13,19,31,32]. Equiaxial γ′-phase particles have a size d~8–12 nm and coherent conjugation with the high-temperature γ-phase. These data are consistent with previously published results in [5,13,19,31,32]. TEM studies have shown that the particle size of the γ′-phase depends weakly on the ageing time at the second step and the external, applied stresses during ageing (Figure 2a,b).
The β-phase particles of the non-equiaxial shape with B2-type ordering have the chemical composition, (FeNiCo)(AlNb) [31,32] and are non-coherent with the high-temperature γ-phase (Figure 2d,e). Due to the significant difference in the lattice parameters (aγ = 0.36 nm, aβ = 0.28 nm), the β-phase particles lose coherence at small sizes, d < 5 nm [35]. Similar lattice parameters of the γ- and β-phases were obtained in the FeNiCoAlTaB polycrystals (aγ = 0.3604 nm, aβ = 0.2880 nm) [5,31,32]. In the FeNiCoAlTaB polycrystals, β-phase particles are precipitated at the grain boundaries with the formation of regions free from γ′-phase particles. In these polycrystals, the nucleation of β-phase particles was also found on the Ta particles, with a lattice parameter close to the β-phase. The Ta particles decrease the energy barrier, associated mainly with the generation of elastic energy, for the β-phase nucleation [31,38]. In the FeNiCoAlNb single crystals, β-phase particles are precipitated in the volume of the crystal, and zones free from γ′-phase particles are not observed. During ageing without and under stress at the second step, four crystallographically equivalent variants of the β-phase particles, with an orientation relationship (111)γ||(110)β, [ 1 1 ¯ 0 ] γ|| [ 1 1 ¯ 1 ] β, are formed (Figure 2d–f) [29,38]. With an increase in the ageing time at the second step, the volume fraction and particle size of the β-phase increase from f = 0.1, d = 10–15 nm and l = 35–40 nm at 2 h, to f = 0.15, d = 22–35 nm and l = 100–135 nm at 4 h (Figure 2d,e). Ageing under a tensile stress of 120 MPa does not lead to a notable difference in the morphology of the particles of the γ′-and β-phases compared to ageing without stress. In crystals aged under stress, an increased dislocation density was observed at the “matrix–β-phase” interface (Figure 2e), which was less pronounced during ageing without stress.
TEM studies facilitate an explanation of the dependence of the Ms temperature on the ageing conditions [2]. Firstly, in Crystals I–IV, the volume fraction and particle size of the γ′-phase change insignificantly (Figure 2a,b) and, therefore, the Ms temperature change is not associated with this phase. Secondly, in Crystals I–IV, the volume fraction and particle size of the β-phase change in relation to the ageing time at the second step. The volume fraction and size of β-phase particles are larger in Crystals III and IV, than in Crystals I and II. This leads to a decrease in the nickel concentration in the matrix and explains the increase in the Ms temperature in Crystals III and IV, in comparison with Crystals I and II [1,2,3,4,5,6,38]. Thirdly, the differences in the sizes and volume fractions of β-phase particles, as a consequence of the ageing conditions under and without stress, were not found in the study of the crystal structure. Consequently, an increase in the Ms temperature in Crystals II and IV compared to Crystals I and III, respectively, cannot be explained from the standpoint of changes in the concentration of alloying elements in the matrix, and in the morphology of the particles of the γ′- and β-phases. The increase in the Ms temperature in Crystals II and IV in comparison with Crystals I and III is due to the formation of internal oriented stress fields <σG> from the β-phase particles. A qualitative confirmation of the <σG> appearance is the formation of dislocations at the “matrix-β-phase” boundaries (Figure 2e), as was previously observed in TiNi crystals, aged under stress and containing partially non-coherent Ti3Ni4 precipitates [29,30,39,40].
As can be seen in Table 1, γ–α′ MT during heating-cooling in a free state, is associated with thermoelastic MT of the second type, in which As ≤ Ms [29,30,41]. Another feature is the large value of overcooling ΔM = Ms − Mf, with a forward γ–α′ MT and overheating of ΔA = Af − As with a reverse α′–γ MT, varying from 50 to 74 K in combination with low values of thermal hysteresis, ΔTh = Af − Ms = 18–40 K. According to the results of thermodynamic analysis for the thermoelastic MT of the second type [41,42,43,44,45], the ratio of the stored elastic energy, ΔGel, is equal to or exceeds the doubled value of the dissipated energy, 2ΔGdis, ΔGel ≥ 2ΔGdis. The values of ΔGel and ΔGdis are estimated from the relationships obtained in [33,41,42,43,44,45]:
Δ G el = Δ S ch 2 ( M s M f ) + Δ S ch 2 ( A f A s )
Δ G dis = Δ S ch 2 ( A f M s )
Ms, Mf and As, Af are, respectively, the temperatures of the start and finish of the forward γ–α′ MT during cooling and the reverse during the heating of the α′–γ MT; ΔSch is the change in entropy under MT. Estimates show that ΔGel/ΔGdis varies from 3 to 6.6 (Table 1). High values of ΔGel, as will be shown below, determine the temperature range of SE in Crystals I–IV.

3.2. SE and SME in the [001]-Oriented FeNiCoAlNb Crystals Aged under and without Stress

SE in Crystals I–IV was found within a broad temperature range (Figure 3). A comparison of the γ–α′ MT temperatures, obtained from the ρ(T) dependence (Figure 1) and the SE temperature range (Figure 3), shows that SE in Crystals I–IV takes place at Ms temperature, within the temperature range Ms < T < Af and above the Af temperature. According to well-known experimental and theoretical studies, the conditions for SE are achieved at T ≥ Af, when martensite becomes thermodynamically unstable, when the load is removed [29,30]. SE within the temperature range from Ms to Af was observed in FeNiCoAlNb single crystals [33], FeNiCoAlTaB [31,32] and FeNiCoTi polycrystals [9]. Reversible deformation within the temperature range between 77 K and Ms is not associated with the development of γ–α′ MT under stress, but is caused by the reversible motion of “austenite-martensite” interphase and twin boundaries of a cooling α′-martensite in “load–unloading” cycles. It was previously found in FeNiCoAlNb single crystals [33] and FeNiCoAlTaB polycrystals [31,32]. This was defined as having rubber-like behavior [33]. In this paper, the reversible deformation at T < Ms was not investigated.
It may be observed that σcrMs for the onset of stress-induced γ–α′ MT within the temperature range from Ms to 373 K increases, with an increase in the test temperature, and the dependence, σcrMs(T), turns out to be characteristic of alloys, undergoing MT under stress at T > Ms (Figure 3 and Figure 4). This dependence is described by the Clapeyron–Clausius relationship [29,30,45]:
d σ cr Ms dT = Δ S ch ε tr = Δ H ch ε tr T 0 = α
ΔHch and ΔSch are the enthalpy and entropy changes, respectively, at MT; T0 is the temperature of the chemical equilibrium of phases and εtr is the transformation strain at MT. Figure 4 demonstrates that the value α = dσcrMs/dT varies from 3.2 to 3.8 MPa/K in Crystals I–IV. In the case of Crystals II and IV, α = dσcrMs/dT is greater than for Crystals I and III. A comparison of the α = dσcrMs/dT values for Crystals I–IV with the known literature data, obtained on the [001]-oriented FeNiCoAlX crystals (X = Ta, Nb, Ti) under tensile strain and containing only γ′-phase particles, 3–8 nm in size and particles of the γ′- and β-phases [25,27,33,34], showed that these values, α = dσcrMs/dT, are in close proximity to one other.
Using relationship (3) and the results of an increase in the Ms temperature during ageing under stress, ΔMs = Msσ − Ms0 (Msσ and Ms0 are temperatures, respectively, in crystals aged under and without stress), one can estimate the level of internal stresses <σG> in accordance with relationship (4):
G> = (ΔMsGtr) ΔSch; ΔSchtr = α = dσcr/dT
In Crystals II and IV, the value of <σG> calculated by relationship (4) is 35–57 MPa at the ΔMs = 11–15 K and α = (3.2–3.8) MPa/K.
SE loops with a sequential increase in εtr in the “load–unloading” cycle until fracture at Ms and the temperature above Af of Crystals I–IV, are presented in Figure 5. The results obtained for Crystals I are close to those obtained earlier in [33]. In relation to Crystals I, at Ms temperature, the maximum SE was 3.3%. With an increase in the transformation strain, εtr, in the “load–unloading” cycle, an increase in the mechanical hysteresis, Δσ, is observed, and rather high values of the transformation hardening coefficient, Θ = dσ/dε, equal to 2 GPa, are found. At 273 K, the SE was 3%, the mechanical hysteresis, Δσ, increased three times relative to the Ms temperature and became equal to 170 MPa and Θ = 4.3 MPa. Among Crystals II, at an Ms temperature, the SE was 3.5%, Δσ increased to 135 MPa and Θ = dσ/dε, up to 3.2 GPa, compared to Crystals I at this temperature. At 273 K, the SE was 4%, Δσ = 190 MPa and Θ = 8 GPa. Thus, ageing under stress leads to a significant increase in Θ = dσ/dε and Δσ, compared to ageing without stress.
In Crystals III and IV, an increase in the ageing time at the second step to 4 h at 873 K, as shown above, increases the volume fraction of β-phase particles in comparison with Crystals I and II. As a result, the SE decreases, while Δσ and Θ, on the contrary, increase and an irreversible strain appears of up to 0.5% at temperatures above Af (Figure 5).
An analysis of the SE loops for Crystals I–IV in the temperature range from Ms to 373 K, shows that in all studied crystals, the stresses, σcrMs, for the onset of stress-induced γ–α′ MT, are lower than the stresses, σcrAs, of the start of the reverse α′–γ MT, when unloading (Figure 3 and Figure 5). Using the results of the thermodynamic analysis of SE loops [33,42,43,44,45], the contribution of ΔGel and ΔGdis can be determined from the values of σcrMs, σcrMf and σcrAs, σcrAf:
ΔGel = (σcrMf(T) − σcrMs(T))εtr
ΔGdis = ½(σcrMf(T) − σcrAs (T))εtr
σcrMs and σcrMf are the stresses at the start and end of the forward γ–α′ MT under load, and σcrAs and σcrAf are the stresses at the start and end of the reverse α′–γ MT, when the load is removed. The estimate of the ratio ΔGel/ΔGdis will depend on εtr, therefore, its calculations for Crystals I–IV for different temperatures, were tested in relation to the close values of εtr = 2% (Figure 3). Estimates showed that ΔGel/ΔGdis vary from 5.5 to 6.3, within the range between Ms and Af, and at T > Af.
The condition for the appearance of a closed loop at T ≥ Ms was obtained in [33] according to the assumption that σcrAf(T) > 0, and is written as follows:
ΔGch(T) + ΔGel > ΔGdis
Δ G ch ( T ) = ( T T 0 ) Δ S ch is the chemical energy change at MT. At T ≥ Ms, condition (7) is satisfied since ΔGel = (5.5–6.3)ΔGdis and the reverse α′–γ MT occurs, due to the large stored elastic energy. A high level of ΔGel determines high values of the transformation hardening coefficient Θ = dσ/dε at T ≥ Ms. This is due to the fact that β-phase particles have a non-coherent conjugation of the lattice with the high-temperature phase. These particles are stable and their atomic structure locks the MT inside the β-phase. Consequently, in the case of a reversible γ–α′ MT, the β-phase particles are only deformed elastically and can assist to the predominant nucleation of α′-martensite near the “β-phase–high-temperature phase” interphase surfaces, as was previously observed in the TiNi-based composites, containing TiC particles [46,47]. High values of Θ = dσ/dε at a low level of strain ε ≤ 5%, were observed in fcc poly- and single crystals, containing non-coherent, plastically undeformed particles during slip deformation in Cu-SiO2, Cu-Al2O3 [48,49] and upon twinning deformation in Cu-Al-Co with non-coherent CoAl particles [50]. In this case, an increase in ΔGel and, accordingly, <σG>, depending on the level of plastic deformation εpl by slip, twinning and γ–α′ MT in the absence of relaxation processes, is written as [48,49]:
< σ G > = 2 γ Gf ε pl 1 m
θ d σ G d ε pl = 2 γ fG 1 m
γ = 5 7 ν 15 ( 1 ν ) = 0.28, where ν = 0.3 is Poisson’s coefficient for martensite, G = 72 GPa is the shear modulus of martensite, f is a volume fraction of β-phase particles, which is estimated as being between 10–15%, εpl is plastic deformation and m is the Schmid factor for twinning in the [001] orientation in martensite, which is equal to 0.41 at tensile strain [33]. The experimental Θ = dσ/dε values for Crystals I–IV, which ranged from 2 to 8 GPa, were close to the calculated Θ = dσ/dε values according to relationship (9). It is interesting that in the study of SE of the [001]-oriented FeNiCoAlX crystals (X = Ta, Nb, Ti), containing only the nanosized γ′-phase particles, the value of Θ = dσ/dε was equal to 2–3 GPa, σcrMs > σcrAs, Ms < As. Therefore, MTs are transformations of the first type, according to the Tong–Wayman classification [11,12,18,21,27].
SME studies during cooling/heating, under the constant stresses of Crystals I–IV are shown in Figure 6 and Figure 7. The values of SME εSME, overcooling ΔM = Ms − Mf and overheating ΔA = Af − As, as well as thermal hysteresis ΔTh = Af − Ms, depend on the level of applied stresses, σcr, in the “cooling–heating” cycle and heat treatment (Figure 6). In the studied Crystals I–IV, the Msσ temperature of the onset of the γ–α′ MT under stress increases with an increase in the level of applied stresses, σcr, and coincides with the Msσ temperature obtained in experiments on the SE study (Figure 3 and Figure 4). The values of ΔM, ΔA, ΔTh and εtr also increase with increasing level of applied stresses, σcr. Figure 4 displays the stresses of the finish of the reverse MT, σcrAf, when the load is removed, which like the stresses for the onset of the stress-induced MT σcrMs, were obtained from the results of studying the SE temperature range (Figure 3) and the SME under stress (Figure 6). It can be seen that similar values of σcrAf and σcrMs were obtained for Crystals I–IV in both experiments. The σcrAf curves for Crystals I and II are parallel to the σcrMs curves, which indicates a weak dependence of Δσ on the test temperature and thermal hysteresis ΔTh on the level of external stresses. In Crystals III and IV, Δσ and ΔTh are greater than in Crystals I and II and increase with growth of the test temperature and external stresses, respectively. In this case, the σcrMs and σcrAf curves are not parallel to each other. Therefore, in Crystals III and IV, Δσ and ΔTh depend, respectively, on the test temperature and external stresses. The physical reason for the stronger dependence of the Δσ and ΔTh, respectively, on the test temperature and external stresses in Crystals III and IV as compared to Crystals I and II requires additional structural studies.
A comparison of Crystals I and III with Crystals II and IV reveals a number of differences. Firstly, the values of εtr and dεtr/dσ in Crystals III and IV differ significantly (Figure 6 and Figure 7). At a stress level of 300 MPa, in case of Crystals IV, εtr reaches 3.3%, whereas εtr reaches 1.4% in Crystals III. The value of εtr in these crystals is limited by their destruction. In the [001]-oriented FeNiCoAlX crystals (X = Ta, Nb, Ti), containing only the γ′-phase particles, εtr = 5.5–8.7% in experiments relating to cooling/heating under constant stress [12,18] and up to 13.5% in experiments on the SE study [5,27]. Similar results were obtained for Crystals I and II (Figure 6 and Figure 7). Therefore, as studies of the SE and SME in single crystals containing particles of the γ′- and β-phases have shown, the particles of the β-phase lead to a decrease in the plasticity associated with the stress induced γ–α′ MT.
Secondly, among Crystals I–IV at the level of constant stress of σcr = 300–350 MPa, high values of ΔM and ΔA were found, which reached 150 and 190 K; 110 and 140 K; 150 and 190 K; and 125 and 200 K, respectively, for Crystals I, II, III and IV. At the maximum level of constant stress of σcr = 300–350 MPa, the values of thermal hysteresis, ΔTh, in Crystals I, II, III and IV are equal to 150 K, 50 K, 50 K and 100 K, respectively. In all studied crystals on the εtr(T) curves under constant stress, the As temperature is lower than the Ms temperature (Figure 6). Therefore, such MTs under constant stress are related to the thermoelastic MT of the second type [41]. This indicates that in the [001]-oriented FeNiCoAlNb crystals, with particles of the γ′- and β-phases, a high level of ΔGel accumulates at the γ–α′ MT under stress. These results confirm the aforementioned conclusions regarding the second type of γ–α′ MT, obtained in the study of the temperature dependence of ρ(T) (Figure 1) and SE (Figure 3 and Figure 5).
Thirdly, ageing under stress leads to the formation of internal oriented stress fields <σG>, which is confirmed by an increase in the Ms temperature in Crystals II and IV as compared to Crystals I and III with the same ageing time at the second step. The source of these internal stress fields <σG> = 35–57 MPa are dislocations arising at the “matrix–β-phase” interface during ageing under stress (Figure 2e). Earlier, in the [001]-oriented crystals of another Fe–28Ni–17Co–11.5Al–2.5(TiNb) (at.%) (FeNiCoAlNbTi) alloys, after ageing under stress, as in Crystals II and also containing particles of γ′- and β-phases, similar values of ΔMs = 15 K and <σG> = 65 MPa were found [51].
Thus, the development of thermoelastic γ–α′ MT in the [001]-oriented FeNiCoAlX crystals (X = Ta, Nb, Ti) occurs in a disordered fcc high-temperature phase, in which the nanosized γ′-phase particles with a size d = 3–12 nm are precipitated. The fundamental role of the γ′-phase particles is that they act as “memory elements”, which determine the restoration of the initial orientation of austenite and the realization of the fact that reverse α′–γ MT is “exactly backward” [1,2,4,5,28].
The γ′-phase particles have coherent conjugation with the high-temperature phase and retain coherence with the α′-martensite. With d > dcr at γ–α′ MT, there is a loss of coherence of the γ′-phase with α′-martensite and relaxation of the elastic energy, ΔGel, stored during the MT. As a result, the transformation becomes non-thermoelastic [1,2]. In relation to Crystals I–IV, containing particles of γ′- and β-phases, the γ–α′ MT is not suppressed by the presence of the β-phase particles. In contrast to the γ′-phase particles, the β-phase particles have non-coherent conjugation with the γ-phase and with α′-martensite. This leads to a decrease in plasticity and, accordingly, to a decrease in the value of SE and SME at MT under stress. Non-coherent β-phase particles act as preferential sites for the nucleation of α′-martensite, which grows in the regions between the β-phase. The high values of Θ = dσ/dε, due to MT in these crystals, which are characteristic of the development of deformation by slip and twinning [48,49,50], indicate an increase in ΔGel with an increase in the volume fraction of α′-martensite.
The combination of a high level of ΔGel and small values of ΔGdis creates conditions for the development of SE within the temperature range from Ms to Af, which are not usually observed during the development of thermoelastic MTs. Ageing under stress at the second step leads to an increase in Ms temperature, compared to ageing without stress, which is due to the appearance of the internal oriented stress fields, <σG>. To elucidate the mechanism of <σG> formation, additional studies are required, in which it will be necessary to change the ageing conditions at the second step, in order to amend the size and volume fraction of the β-phase particles. This will facilitate the creation of unique composite materials, with a simultaneous precipitation of γ′- and β-phase particles, high transformation strain values and the development interval of the γ–α′ MT within the high-temperature region due to a decrease in nickel in the fcc high-temperature phase, as a result of the precipitation of the β-phase particles.
In the present paper, it is shown that the maximum values of the transformation strain, εtr, obtained in experiments when studying the SE and SME during cooling/heating under constant stress, are equal to 4 and 3.4%, respectively. Theoretical estimates of εtr, carried out in [5,12], show that the value of εtr for the [001]-oriented crystals under tension decreases from 8.3 to 5.5% with an increase in the tetragonality of martensite c/a from 1.1 to 1.15%. This estimate assumes that the γ–α′ MT occurs in the alloy without the γ′-phase particles or when these particles undergo elastic deformation together with the fcc matrix. The experimental values of the SME and SE under tension along the [001] direction vary from 14.5 to 5.5% for the γ′-phase particle sizes of 3 nm and 8–12 nm [5,12,18,26,27]. The non-coherent β-phase particles, the sizes of which are much larger than the γ′-phase particles, cannot accommodate transformation strain in the matrix due to elastic deformation and contribute to the generation of multiple variants of martensite from the “matrix-β-phase” interface [46,47]. As a result, εtr significantly decreases to between 3.4 and 4% and the determination of the total resource of εtr is limited by the low plasticity of Crystals I–IV due to the effect of the β-phase on transformation plasticity.

4. Conclusions

The studies of development of the thermoelastic γ–α′ MT under stress using [001]-oriented FeNiCoAlNb single crystals, containing non-coherent β-phase particles after ageing under and without stress, showed that the effect of the β-phase particles on SME and SE in single crystals and polycrystals of FeNiCoAlX alloys (X = Ta, Nb, Ti) is fundamentally different. Based on the results of the SME and SE studies, crystal structure after ageing under and without stresses and their analysis, the following conclusions can be summarized as the main findings:
  • A new two-step ageing has been proposed for the FeNiCoAlNb single crystals, oriented along the [001]-direction for tensile strain, including ageing at the first step at 973 K for 5 h without stress and at the second step, at 873 K for 2 and 4 h without and under tensile stress at 120 MPa. At the first step of ageing at 973 K for 5 h, particles of the γ′-phase are precipitated, and at the second low-temperature step at 873 K for 2 and 4 h, a structure, consisting of particles of the γ′- and β-phases is formed.
  • An increase in the ageing time at the second low-temperature step from 2 to 4 h leads to an increase in the Ms temperature, which is associated with a decrease in the nickel concentration in the matrix due to an increase in the volume fraction of the nickel-rich β-phase particles with the composition, (FeNiCo)(AlNb).
  • Ageing under a tensile stress of 120 MPa at the second low-temperature step for 2 and 4 h leads to an increase in the Ms temperature, in comparison with ageing without stress by 11–15 K, which is associated with the generation of internal oriented tensile stresses during ageing under stress.
  • In a structure containing particles of the γ′- and β-phases, conditions are created for observing SE within the temperature range from Ms to Af due to the formation of a high level of accumulated elastic energy, ΔGel, which significantly exceeds the value of the dissipated energy, ΔGdis.

Author Contributions

Conceptualization, Y.I.C. and I.V.K.; methodology, Y.I.C. and I.V.K.; investigation, I.V.K. and Z.V.P.; writing—original draft preparation, Y.I.C. and I.V.K.; writing—review and editing, Y.I.C., I.V.K., T.N. and P.K.; funding acquisition, Y.I.C. and P.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Russian Science Foundation (project No. 19-49-04101) and by (Project No. 405372848 (KR 5134/1-1)).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are available from the corresponding author on reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kokorin, V.V. Martensitic Transformations in Heterogeneous Solid Solutions. Naukova Dumka 1987, 1987, 165. (In Russian) [Google Scholar]
  2. Hornbogen, E. The effect of variables on martensitic transformation temperatures. Acta Met. 1985, 33, 595–601. [Google Scholar] [CrossRef]
  3. Hornbogen, E.; Jost, N. Alloys of iron and reversibility of martensitic transformations. J. Phys. Colloq. 1991, 1, C4-199–C4-210. [Google Scholar] [CrossRef]
  4. Maki, T. Ferrous shape memory alloys. In Book Shape Memory Material, 1st ed.; Otsuka, K., Wayman, C.M., Eds.; Cambridge University Press: Cambridge, UK, 1998; pp. 117–132. [Google Scholar]
  5. Tanaka, Y.; Himuro, Y.; Kainuma, R.; Sutou, Y.; Omori, T.; Ishida, K. Ferrous Polycrystalline Shape-Memory Alloy Showing Huge Superelasticity. Science 2010, 327, 1488–1490. [Google Scholar] [CrossRef]
  6. Choi, W.S.; Pang, E.L.; Choi, P.-P.; Schuh, C.A. FeNiCoAlTaB superelastic and shape-memory wires with oligocrystalline grain structure. Scr. Mater. 2020, 188, 1–5. [Google Scholar] [CrossRef]
  7. Cesari, E.; Chernenko, V.; Kokorin, V.; Pons, J.; Seguí, C. Physical properties of Fe-Co-Ni-Ti alloy in the vicinity of martensitic transformation. Scr. Mater. 1999, 40, 341–345. [Google Scholar] [CrossRef]
  8. Kokorin, V.V.; Kozlova, L.E.; Titenko, A.N.; Perekos, A.E.; Levchuk, Y.S. Charecteristic of thermoelastic martensitic trans-formation in ferromagnetic Fe-CoNi-Ti alloys alloyed with Cu. Phys. Metals Metallogr. 2008, 105, 564–567. [Google Scholar] [CrossRef]
  9. Gunko, L.P.; Takzei, G.A.; Titenko, A.N. Thermoplastic martensitic transformation in ferromagnetic materials and their superelastic properties. Funct. Mater. 2002, 9, 75–78. [Google Scholar]
  10. Geng, Y.; Lee, D.; Xu, X.; Nagasako, M.; Jin, M.; Jin, X.; Omori, T.; Kainuma, R. Coherency of ordered γ′ precipitates and thermoelastic martensitic transformation in FeNiCoAlTaB alloys. J. Alloy. Compd. 2015, 628, 287–292. [Google Scholar] [CrossRef]
  11. Chumlyakov, Y.I.; Kireeva, I.V.; Panchenko, E.; Kirillov, V.A.; Timofeeva, E.E.; Kretinina, I.V.; Danil’Son, Y.N.; Karaman, I.; Maier, H.J.; Cesari, E. Thermoelastic martensitic transformations in single crystals with disperse particles. Russ. Phys. J. 2011, 54, 937–950. [Google Scholar] [CrossRef]
  12. Ma, J.; Hornbuckle, B.; Karaman, I.; Thompson, G.; Luo, Z.; Chumlyakov, Y. The effect of nanoprecipitates on the superelastic properties of FeNiCoAlTa shape memory alloy single crystals. Acta Mater. 2013, 61, 3445–3455. [Google Scholar] [CrossRef]
  13. Czerny, M.; Maziarz, W.; Cios, G.; Wójcik, A.; Chumlyakov, Y.; Schell, N.; Fitta, M.; Chulist, R. The effect of heat treatment on the precipitation hardening in FeNiCoAlTa single crystals. Mater. Sci. Eng. A 2020, 784, 139327. [Google Scholar] [CrossRef]
  14. Jin, M.; Geng, Y.; Zuo, S.; Jin, X. Precipitation and its Effects on Martensitic Transformation in Fe-Ni-Co-Ti Alloys. Mater. Today Proc. 2015, 2, S837–S840. [Google Scholar] [CrossRef]
  15. Fu, H.; Li, W.; Song, S.; Jiang, Y.; Xie, J. Effects of grain orientation and precipitates on the superelasticity in directionally solidified FeNiCoAlTaB shape memory alloy. J. Alloy. Compd. 2016, 684, 556–563. [Google Scholar] [CrossRef]
  16. Fu, H.; Zhao, H.; Zhang, Y.; Xie, J. Enhancement of Superelasticity in Fe-Ni-Co-Based Shape Memory Alloys by Microstructure and Texture Control. Procedia Eng. 2017, 207, 1505–1510. [Google Scholar] [CrossRef]
  17. Krooß, P.; Holzweissig, M.J.; Niendorf, T.; Somsen, C.; Schaper, M.; Chumlyakov, Y.I.; Maier, H.J. Thermal cycling behavior of an aged FeNiCoAlTa single-crystal shape memory alloy. Scr. Mater. 2014, 81, 28–31. [Google Scholar] [CrossRef]
  18. Ma, J.; Kockar, B.; Evirgen, A.; Karaman, I.; Luo, Z.; Chumlyakov, Y. Shape memory behavior and tension–compression asymmetry of a FeNiCoAlTa single-crystalline shape memory alloy. Acta Mater. 2012, 60, 2186–2195. [Google Scholar] [CrossRef]
  19. Czerny, M.; Cios, G.; Maziarz, W.; Chumlyakov, Y.; Chulist, R. Studies on the Two-Step Aging Process of Fe-Based Shape Memory Single Crystals. Materials 2020, 13, 1724. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Krooß, P.; Niendorf, T.; Karaman, I.; Chumlyakov, Y.I.; Maier, H.J. Cyclic deformation behavior of aged FeNiCoAlTa single crystals. Funct. Mater. Lett. 2012, 5, 12500454. [Google Scholar] [CrossRef]
  21. Sehitoglu, H.; Karaman, I.; Zhang, X.; Chumlyakov, Y.; Maier, H.J. Deformation of FeNiCoTi shape memory single crystals. Scr. Mater. 2001, 44, 779–784. [Google Scholar] [CrossRef]
  22. Chumlakov, Y.I.; Kireeva, I.V.; Panchenko, E.Y.; Zakharova, E.G.; Kirillov, V.A.; Efimenko, S.P.; Sehitoglu, H. Shape memory effect in FeNiCoTi single crystals undergoing γ–α′ thermoelastic martensitic transformations. Doklady Phys. 2004, 49, 47–50. [Google Scholar] [CrossRef]
  23. Sehitoglu, H.; Efstathiou, C.; Maier, H.; Chumlyakov, Y. Hysteresis and deformation mechanisms of transforming FeNiCoTi. Mech. Mater. 2006, 38, 538–550. [Google Scholar] [CrossRef]
  24. Chumlyakov, Y.; Kireeva, I.; Kuksgauzen, I.; Poklonov, V.; Pobedennaya, Z.; Bessonova, I.; Kuksgauzen, D.; Kirillov, V.; Lauhoff, C.; Niendorf, T.; et al. Shape memory effect and superelasticity in high-strength FeNiCoAlTi single crystals hardened by nanoparticles. AIP Conf. Proc. 2019, 2167, 020064. [Google Scholar] [CrossRef]
  25. Chumlyakov, Y.I.; Kireeva, I.V.; Kuksgauzen, I.V.; Kuksgauzen, D.A.; Niendorf, T.; Krooß, P. Tension-compression asymmetry of the superelastic behavior of high-strength [001]-oriented FeNiCoAlNb crystals. Mater. Lett. 2021, 289, 129395. [Google Scholar] [CrossRef]
  26. Sehitoglu, H.; Zhang, X.Y.; Kotil, T.; Canadinc, D.; Chumlyakov, Y.; Maier, H.J. Shape memory behavior of FeNiCoTi single and polycrystals. Met. Mater. Trans. A 2002, 33, 3661–3672. [Google Scholar] [CrossRef]
  27. Chumlyakov, Y.; Kireeva, I.; Kutz, O.; Turabi, A.; Karaca, H.; Karaman, I. Unusual reversible twinning modes and giant superelastic strains in FeNiCoAlNb single crystals. Scr. Mater. 2016, 119, 43–46. [Google Scholar] [CrossRef]
  28. Dunne, D. Martens-ite. Metals 2018, 8, 395. [Google Scholar] [CrossRef] [Green Version]
  29. Otsuka, K.; Wayman, C.M. Shape Memory Materials; Cambridge University Press: Cambridge, UK, 1998; p. 284. [Google Scholar]
  30. Otsuka, K.; Ren, X. Physical metallurgy of Ti–Ni-based shape memory alloys. Prog. Mater. Sci. 2005, 50, 511–678. [Google Scholar] [CrossRef]
  31. Zhang, C.; Zhu, C.; Harrington, T.; Casalena, L.; Wang, H.; Shin, S.; Vecchio, K.S. Multifunctional Non-Equiatomic High Entropy Alloys with Superelastic, High Damping, and Excellent Cryogenic Properties. Adv. Eng. Mater. 2019, 21, 1800941. [Google Scholar] [CrossRef] [Green Version]
  32. Zhang, C.; Zhu, C.; Shin, S.; Casalena, L.; Vecchio, K. Grain boundary precipitation of tantalum and NiAl in superelastic FeNiCoAlTaB alloy. Mater. Sci. Eng. A 2019, 743, 372–381. [Google Scholar] [CrossRef]
  33. Chumlyakov, Y.; Kireeva, I.; Pobedennaya, Z.; Krooβ, P.; Niendorf, T. Rubber-like behaviour and superelasticity of [001]-oriented FeNiCoAlNb single crystals containing γ′- and β-phase particles. J. Alloy. Compd. 2021, 856, 158158. [Google Scholar] [CrossRef]
  34. Chumlyakov, Y.; Kireeva, I.; Kuksgauzen, I.; Poklonov, V.; Pobedennaya, Z.; Bessonova, I.; Kirillov, V.; Lauhoff, C.; Niendorf, T.; Krooβ, P. Effect of β- and γ’-phase particles on the shape memory effect and superelasticity in [001]-oriented FeNiCoAlTi single crystals. Mater. Lett. 2020, 260, 126932. [Google Scholar] [CrossRef]
  35. Li, D.; Chen, L. Selective variant growth of coherent Ti11Ni14 precipitate in a TiNi alloy under applied stresses. Acta Mater. 1997, 45, 471–479. [Google Scholar] [CrossRef]
  36. Li, D.; Chen, L. Morphological evolution of coherent multi-variant Ti11Ni14 precipitates in Ti-Ni alloys under an applied stress—A computer simulation study. Acta Mater. 1998, 46, 639–649. [Google Scholar] [CrossRef]
  37. Kireeva, I.; Picornell, C.; Pons, J.; Kretinina, I.; Chumlyakov, Y.; Cesari, E. Effect of oriented γ′ precipitates on shape memory effect and superelasticity in Co–Ni–Ga single crystals. Acta Mater. 2014, 68, 127–139. [Google Scholar] [CrossRef]
  38. Poster, D.A.; Easterling, K.E. Phase Transformation in Metals and Alloys; Chapman & Hall: London, UK, 1981; 165p. [Google Scholar]
  39. Khalil-Allafi, J.; Dlouhy, A.; Eggeler, G. Ni4Ti3-precipitation during aging of NiTi shape memory alloys and its influence on martensitic phase transformations. Acta Mater. 2002, 50, 4255–4274. [Google Scholar] [CrossRef]
  40. Li, J.-F.; Zheng, Z.-Q.; Li, X.-W.; Li, S.-C. Effect of compressive stress aging on transformation strain and microstructure of Ni-rich TiNi alloy. Mater. Sci. Eng. A 2009, 523, 207–213. [Google Scholar] [CrossRef]
  41. Tong, H.; Wayman, C. Characteristic temperatures and other properties of thermoelastic martensites. Acta Met. 1974, 22, 887–896. [Google Scholar] [CrossRef]
  42. Daróczi, L.; Palánki, Z.; Szabó, S.; Beke, D. Stress dependence of non-chemical free energy contributions in Cu–Al–Ni shape memory alloy. Mater. Sci. Eng. A 2004, 378, 274–277. [Google Scholar] [CrossRef]
  43. Palánki, Z.; Daróczi, L.; Beke, D.L. Method for the Determination of Non-Chemical Free Energy Contributions as a Function of the Transformed Fraction at Different Stress Levels in Shape Memory Alloys. Mater. Trans. 2005, 46, 978–982. [Google Scholar] [CrossRef] [Green Version]
  44. Beke, D.; Daróczi, L.; Samy, N.; Tóth, L.; Bolgár, M. On the thermodynamic analysis of martensite stabilization treatments. Acta Mater. 2020, 200, 490–501. [Google Scholar] [CrossRef]
  45. Wollants, P.; Roos, J.; Delaey, L. Thermally- and stress-induced thermoelastic martensitic transformations in the reference frame of equilibrium thermodynamics. Prog. Mater. Sci. 1993, 37, 227–288. [Google Scholar] [CrossRef]
  46. Vaidyanathan, R.; Bourke, M.; Dunand, D. Phase fraction, texture and strain evolution in superelastic NiTi and NiTi–TiC composites investigated by neutron diffraction. Acta Mater. 1999, 47, 3353–3366. [Google Scholar] [CrossRef]
  47. Dunand, D.C.; Mari, D.; Bourke, M.A.M.; Roberts, J.A. NiTi and NiTi-TiC composites: Part IV. Neutron diffraction study of twinning and shape-memory recovery. Met. Mater. Trans. A 1996, 27, 2820–2836. [Google Scholar] [CrossRef]
  48. Martin, J.W. Micromechanisms in Partical-Hardened Alloys; Cambridge Solid State Science Series; Cambridge University Press: Cambridge, UK, 1980; p. 197. [Google Scholar]
  49. Brown, L. Back-stresses, image stresses, and work-hardening. Acta Met. 1973, 21, 879–885. [Google Scholar] [CrossRef]
  50. Korotaev, A.D.; Chumlyakov, Y.I.; Esipenko, V.F.; Bushnev, L.S. Superelasticity effects in single crystals of Cu-15% Al-2% Co with non-coherent particles due to twinning. Phys. Status Solidi A 1984, 82, 405–412. [Google Scholar] [CrossRef]
  51. Chumlyakov, Y.; Kireeva, I.; Pobedennaya, Z.; Kuksgauzen, I.; Poklonov, V.; Krooß, P.; Niendorf, T.; Lauhoff, C.; Vollmer, M. Shape memory effect and superelasticity in high-strength FeNiCoAlTiNb single crystals. AIP Conf. Proc. 2020, 2310, 020065. [Google Scholar] [CrossRef]
Figure 1. Temperature dependence of electrical resistivity ρ(T) in the FeNiCoAlNb single crystals after two-step ageing: (a) ageing at 873 K for 2 h at the second step; (b) ageing at 873 K for 4 h at the second step.
Figure 1. Temperature dependence of electrical resistivity ρ(T) in the FeNiCoAlNb single crystals after two-step ageing: (a) ageing at 873 K for 2 h at the second step; (b) ageing at 873 K for 4 h at the second step.
Metals 11 00943 g001
Figure 2. Transmission electron microscope (TEM) investigation of structure of the [001]-oriented FeNiCoAlNb crystals after two-step ageing: (a) dark-field image of γ′-phase particles in the γ′-phase reflex after ageing at 873 K for 2 h without stress at the second step; (b) dark-field image of γ′-phase particles in the γ′-phase reflex after ageing at 873 K for 4 h under stress at the second step; (c) diffraction pattern showing reflections of γ′-phase particles that are observed in case (a,b); (d) bright-field image of β-phase particles after ageing at 873 K for 2 h at the second step without stress; (e) bright-field image of β-phase particles after ageing at 873 K for 4 h at the second step under stress of 120 MPa; (f) corresponding diffraction pattern to (b,c).
Figure 2. Transmission electron microscope (TEM) investigation of structure of the [001]-oriented FeNiCoAlNb crystals after two-step ageing: (a) dark-field image of γ′-phase particles in the γ′-phase reflex after ageing at 873 K for 2 h without stress at the second step; (b) dark-field image of γ′-phase particles in the γ′-phase reflex after ageing at 873 K for 4 h under stress at the second step; (c) diffraction pattern showing reflections of γ′-phase particles that are observed in case (a,b); (d) bright-field image of β-phase particles after ageing at 873 K for 2 h at the second step without stress; (e) bright-field image of β-phase particles after ageing at 873 K for 4 h at the second step under stress of 120 MPa; (f) corresponding diffraction pattern to (b,c).
Metals 11 00943 g002
Figure 3. Temperature range of the reversible strain in the [001]-oriented FeNiCoAlNb crystals after two-step ageing under tension: (a,b)—ageing at 873 K for 2 h at the second step; (c,d)—ageing at 873 K for 4 h at the second step. Determination of critical stresses ( σcrMs, σcrMf, σcrAs and σcrAf) and value of mechanical hysteresis Δσ is shown in (a,b).
Figure 3. Temperature range of the reversible strain in the [001]-oriented FeNiCoAlNb crystals after two-step ageing under tension: (a,b)—ageing at 873 K for 2 h at the second step; (c,d)—ageing at 873 K for 4 h at the second step. Determination of critical stresses ( σcrMs, σcrMf, σcrAs and σcrAf) and value of mechanical hysteresis Δσ is shown in (a,b).
Metals 11 00943 g003
Figure 4. Temperature dependence of critical stresses, σcrMs, for the onset of stress-induced γ–α′ MT and critical stresses, σcrAf, for the finish reverse α′–γ MT at unloading of the [001]-oriented FeNiCoAlNb crystals after two-step ageing under tensile strain. (a,b)—ageing at 873 K for 2 h at the second step; (c,d)—ageing at 873 K for 4 h at the second step; open squares and circles are the data obtained in the study of the SE temperature range, and filled squares and circles are the data obtained in the study of the SME under stress.
Figure 4. Temperature dependence of critical stresses, σcrMs, for the onset of stress-induced γ–α′ MT and critical stresses, σcrAf, for the finish reverse α′–γ MT at unloading of the [001]-oriented FeNiCoAlNb crystals after two-step ageing under tensile strain. (a,b)—ageing at 873 K for 2 h at the second step; (c,d)—ageing at 873 K for 4 h at the second step; open squares and circles are the data obtained in the study of the SE temperature range, and filled squares and circles are the data obtained in the study of the SME under stress.
Metals 11 00943 g004
Figure 5. Tensile superelastic cycles of the [001]-oriented FeNiCoAlNb crystals after two-step ageing at different temperatures: (a)—Crystals I—ageing without stress at 873 K for 2 h at the second step; (b)—Crystals III—ageing without stress at 873 K for 4 h at the second step; (c)—Crystals II—ageing under stress at 873 K for 2 h at the second step; (d)—Crystals IV—ageing under stress at 873 K for 4 h at the second step.
Figure 5. Tensile superelastic cycles of the [001]-oriented FeNiCoAlNb crystals after two-step ageing at different temperatures: (a)—Crystals I—ageing without stress at 873 K for 2 h at the second step; (b)—Crystals III—ageing without stress at 873 K for 4 h at the second step; (c)—Crystals II—ageing under stress at 873 K for 2 h at the second step; (d)—Crystals IV—ageing under stress at 873 K for 4 h at the second step.
Metals 11 00943 g005
Figure 6. Strain vs. temperature response of the [001]-oriented FeNiCoAlNb crystals after two-step ageing: (a)—ageing without stress at 873 K for 2 h at the second step; (b)—ageing under stress of 120 MPa at 873 K for 2 h at the second step; (c)—ageing without stress at 873 K for 4 h at the second step; (d)—ageing under stress of 120 MPa at 873 K for 4 h at the second step.
Figure 6. Strain vs. temperature response of the [001]-oriented FeNiCoAlNb crystals after two-step ageing: (a)—ageing without stress at 873 K for 2 h at the second step; (b)—ageing under stress of 120 MPa at 873 K for 2 h at the second step; (c)—ageing without stress at 873 K for 4 h at the second step; (d)—ageing under stress of 120 MPa at 873 K for 4 h at the second step.
Metals 11 00943 g006aMetals 11 00943 g006b
Figure 7. Transformation strain vs. tensile stress response of the [001]-oriented FeNiCoAlNb crystals after two-step ageing.
Figure 7. Transformation strain vs. tensile stress response of the [001]-oriented FeNiCoAlNb crystals after two-step ageing.
Metals 11 00943 g007
Table 1. Characteristics of γ–α′ MT in the FeNiCoAlNb single crystals after ageing under and without stress, obtained from the analysis of the temperature dependence of the electrical resistivity ρ(T): ΔTh = Af − Ms is the thermal hysteresis; ΔM = Ms − Mf is overcooling; ΔA = Af − As is overheating; ΔGel is the stored elastic energy; ΔGdis is the dissipated energy.
Table 1. Characteristics of γ–α′ MT in the FeNiCoAlNb single crystals after ageing under and without stress, obtained from the analysis of the temperature dependence of the electrical resistivity ρ(T): ΔTh = Af − Ms is the thermal hysteresis; ΔM = Ms − Mf is overcooling; ΔA = Af − As is overheating; ΔGel is the stored elastic energy; ΔGdis is the dissipated energy.
CrystalsMs, KMf, KAs, KAf, KΔTh, KΔM, KΔA, KΔGel/ΔGdis
Crystal I1851201502031865536.6
Crystal II2001301902404070503
Crystal III2141401802352174556.5
Crystal IV2251552102603570503.7
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chumlyakov, Y.I.; Kireeva, I.V.; Pobedennaya, Z.V.; Krooß, P.; Niendorf, T. Shape Memory Effect and Superelasticity of [001]-Oriented FeNiCoAlNb Single Crystals Aged under and without Stress. Metals 2021, 11, 943. https://doi.org/10.3390/met11060943

AMA Style

Chumlyakov YI, Kireeva IV, Pobedennaya ZV, Krooß P, Niendorf T. Shape Memory Effect and Superelasticity of [001]-Oriented FeNiCoAlNb Single Crystals Aged under and without Stress. Metals. 2021; 11(6):943. https://doi.org/10.3390/met11060943

Chicago/Turabian Style

Chumlyakov, Yuriy I., Irina V. Kireeva, Zinaida V. Pobedennaya, Philipp Krooß, and Thomas Niendorf. 2021. "Shape Memory Effect and Superelasticity of [001]-Oriented FeNiCoAlNb Single Crystals Aged under and without Stress" Metals 11, no. 6: 943. https://doi.org/10.3390/met11060943

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