Next Article in Journal
Underutilization Versus Nutritional-Nutraceutical Potential of the Amaranthus Food Plant: A Mini-Review
Previous Article in Journal
Study on the Effects of Ultrasonic Agitation on CO2 Adsorption Efficiency Improvement of Cement Paste
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ab Initio Study of Martensitic Transformation in NiTiPt High Temperature Shape Memory Alloys

1
School of Materials Science and Engineering, Beihang University, Beijing 100191, China
2
School of Physics and Electronic Science, Zunyi Normal College, Zunyi 563006, China
*
Author to whom correspondence should be addressed.
Appl. Sci. 2021, 11(15), 6878; https://doi.org/10.3390/app11156878
Submission received: 1 July 2021 / Revised: 22 July 2021 / Accepted: 23 July 2021 / Published: 26 July 2021

Abstract

:
The crystal structures and martensitic transformation of Ti 50 Ni 50 x Pt x alloys (x = 0, 6.25, 8.33, 10.42, 12.5, 18.75, 25) were studied by means of density functional theory (DFT). The computational results indicate that the lattice parameters of Ti-Ni-Pt alloys continuously increase with increasing the Pt content. It is found that at ≤ 12.5 at.% Pt, the martensite structure is monoclinic B19 phase, and the energy differences between parent and martensite phases ( Δ E ) decrease slightly with a minimum observed at 6.25 at.% Pt. However, when the Pt content is increased to around 15 at.%, the most stable martensite phase is the orthorhombic B19 structure, and the Δ E increases sharply with Pt concentration. It was found that the phase transition temperatures are closely related to the energy differences Δ E between parent and martensite phases. The electronic structures of martensite B19 and B19 phases are also discussed.

1. Introduction

NiTi shape memory alloys (SMAs) exhibit the shape memory effect and superelasticity which are attributed to reversible thermoelastic martensitic transformations (MTs). The NiTi alloys have found diverse commercial use in biomedicine and engineering applications since 1963 [1,2]. The structure, mechanics, and martensitic phase transformation in binary NiTi shape memory alloy were widely studied [3,4,5,6,7]. As is known, the dependence of the phase transformation temperature on alloy composition is of utmost importance with regard to the application of NiTi SMAs [2,8]. Some alloying additions, for example, Fe, Co, Al, Mn, V, or Cr will decrease the martensite start temperatures ( M s ) of NiTi-based alloys [8,9]. Unfortunately, some special applications such as aerospace, automotive and power generation, and chemical processing industries demand a high transition temperature, which are greatly limited due to the low M s .
Over the past decades, considerable efforts have been made in developing high-temperature shape memory alloys (HTSMAs) with significantly elevated transformation temperatures based on ternary NiTiX (X = Pd, Pt, Au, Hf, and Zr) compositions [10]. Experimental studies indicate additions of hafnium above 3 at.% increase the transformation temperatures of Ti-Ni-Hf alloys [11]. For NiTi alloys, replacing titanium or nickel with Zr or Pd above 10 at.% can increase the transformation temperatures [12,13]. There are many reports on the crystal structures [11,14,15,16,17], phase transformations [18,19,20,21,22,23,24], shape-memory behaviors [12,25,26,27,28,29], and precipitate phases [30,31,32,33,34] in NiTiX (X = Pd, Hf, and Zr) alloys.
Of these HTSMAs, the (Ni,Pt)Ti system, owing to the highest potential use temperature, in spite of the high costs, has been studied as the promising SMAs for demanding applications, such as in the aerospace, automotive, power generation, and chemical processing industries. Yet, compared to the NiTiX (X = Pd, Hf and Zr) alloys, Ti-Ni-Pt alloys have not received a similar level of attention. Recently, Rios et al. [35] reported that the martensitic transformation temperature of the TiNiPt system increases sharply when Pt content exceeds 10 at.%. The M s can be increased to near 1000 ° C in TiPt alloy. Moreover, NiTiPt alloys show the low thermal hysteresis [36,37], good work output [38], and excellent oxidation resistance [39].
The elastic property of the B2 phase in NiTiPt has been investigated by density functional theory (DFT) [40,41]. To the best of our knowledge, few studies discuss the mechanism of phase transition. No theoretical study on the martensitic transformations of TiNiPt has been published. DFT has been used successfully to examine the influence of the Hf, Pd, Cu, or Zr on NiTi properties [42,43,44,45,46,47]. Consequently, the MTs of ternary TiNiPt HTSMAs as a function of the Pt content is investigated in the present research by first principles.

2. Model and Methodology

NiTi B2, orthorhombic B19, and monoclinic B19 structures taken from our previous work [47] are adopted in the present study. Note that we employ the parent NiTi body-centered tetragonal (bct) lattice. The structural parameters of bct are a = a 0 , b = c = 2 a 0 , here a 0 is the lattice constants of cubic B2 NiTi structure. In addition, we focus on the B19 and B19 structures for the martensite phase reported in the literature [3,5,48]. The structures and parameters of B19 and B19 are very close to each other, except the monoclinic distortion from 90 to β = 98.26 . Considering the symmetry of the cell, 3 × 2 × 2 supercells of the B2, B19, and B19 are constructed containing 48 atoms for the present simulation.
The different site occupations of Ni atoms in the austenite B2 3 × 2 × 2 supercell Ni 24 Ti 24 structure, shown in Figure 1, are indicated by smaller blue spheres and labeled with numbers. A survey of the literature [9] reveals that the addition of Pt prefers to occupy the Ni rather than Ti sites in TiNiPt alloys. In this study, replacing nickel with platinum, we adopt the Ti 50 Ni 50 x Pt x alloys (x = 0, 6.25, 8.33, 10.42, 12.5, 18.75, 25) alloys as our models. Geometry optimizations were implemented for all possible doping structures. The most stable structures were calculated with the principle of the lowest energy. Table 1 presents the site occupation of Pt in the parent phase of Ti-rich Ni-Ti-Pt alloys.
The present calculations are performed within the framework of DFT as implemented in the Vienna ab initio simulation package [49,50,51] (VASP) code, applying the generalized gradient approximation [52] (GGA) and a projected augmented wave [53] (PAW) basis. We chose a 500 eV plane-wave energy cutoff and the 4 × 4 × 4 k-points [54] for the 3 × 2 × 2 supercells of the B2, B19, and B19 phases in TiNiPt alloys.
In order to analyze the phase stabilities of B2, B19, and B19 of Ni-Ti-Pt alloys with the increasing of the Pt content, we evaluate the energies of formation, E f [55,56,57]. Formation energies per atom of Ti 50 Ni 50 x Pt x alloys are calculated as:
E f = E tot ( T i 50 N i 50 x P t x ) 50 E Ti ( 50 x ) E Ni x E Pt 100
where E tot ( T i 50 N i 50 x P t x ) is the total energy of TiNiPt alloys. E Ni , E Ti , and E Pt denote the energies per atom of the pure Ni, Ti, and Pt in their bulk states, respectively.
The transformation behaviors as a function of the doping concentration can be described by the total energy differences Δ E A M between the austenite and martensite phases. According to the laws of thermodynamics, the Gibbs free energy G are defined as:
G = U T S + P V
where U, T, S, P, and V are the internal energy, temperature, entropy, pressure, and volume, respectively. From the point of the thermodynamics of phase transition, at the equilibrium temperature T m , the free energies of the austenitic and martensitic phases are equal, i.e.,
Δ G A M = Δ U A M T m Δ S A M + P Δ V A M = 0 .
For solid state phase transition,
P Δ V A M 0 .
Thus,
Δ U A M = T m Δ S A M .
Generally, the DFT calculations are implemented at the temperature of 0 K. The Δ U A M is approximately equal to the total energy differences Δ E A M . So,
Δ E A M = T m Δ S A M
Δ E A M T m ,
T m usually increases with the increasing of Δ E A M . Therefore, the Δ E A M between B2 and B19 (or B19 ) structures can be used to discuss the phase transition temperature T m . A larger energy differences Δ E A M indicate the higher martensite start temperature M S , as has been shown in the literature [47,55,58,59,60].
A schematic diagram of our study is presented in Figure 2. We first construct the computational models, then solve the Kohn–Sham equations, obtain the lattice parameters, and the total energies as a function of the doping concentration. We evaluate the phase stabilities of TiNiPt alloys via calculating the formation energies E f . At last, we describe the martensitic transformation through the energy differences between austenite and martensite phases in TiNiPt alloys.

3. Results and Discussion

3.1. Crystal Structures of TiNiPt Alloys

At first, the calculated lattice parameters of B2, B19, and B19 phases of NiTi presented in Figure 3, which are in good agreement with the experimental and theoretical data [48,61,62,63,64]. Generally the lattice volume depends on both the size of the constituent atoms and the number of valence electrons of the intermetallic crystal [65]. In ternary TiNiPt alloys, the valence electrons per atom are calculated as [11,42,66]:
e v a = f Ni e v Ni + f Ti e v Ti + f Pt e v Pt
where f Ni , f Ti , and f Pt represent the atomic fractions of elements, and e v Ni , e v Ti , and e v Pt are the corresponding number of valence electrons, i.e., e v Ti = 4 , e v Ni = 10 , and e v Pt = 10 , according to the valence electron configurations: Ti-3d 2 4s 2 , Ni-3d 8 4s 2 , and Pt-5d 9 6s 1 .
Substituting Ni with Pt, the number of valence electrons per atom of Ti 50 Ni 50 x Pt x alloys remains a constant 7. Thus, the supercell volume mainly depends on the size of the constituent atoms. The atomic radii of Ni, Ti, and Pt are 1.25 Å, 1.45 Å, and 1.36 Å, respectively. The lattice parameters a, b, c and the lattice volumes of the NiTiPt alloys expand consistently with increasing Pt content in B2, B19, and B19 structures (see Figure 3). This can be attributed to the larger atomic size of Pt than of Ni atoms.

3.2. Phase Stabilities of Ti 50 Ni 50 x Pt x Alloys

Before rationalizing the strong dependence of the martensite start temperature on alloy composition in ternary Ni-Ti-Pt shape memory alloys from first-principles calculations, we examine the transformation pathways of Ti 50 Ni 50 x Pt x alloys (x = 0, 6.25, 8.33, 10.42, 12.5, 18.75, 25).
In Figure 4, we plot the calculated E f as a function of the Pt concentration. The binary NiTi alloy can be considered as a TiNiPt alloy containing zero at.% Pt. Above all, as can be seen from Figure 4, in equiatomic NiTi binary alloys, the B2 phase, which is only stable at high temperatures, exhibits the highest value of E f ; the B19 martensitic structure with the lowest formation energy is the most stable one. They are in excellent agreement with the previous experimental observations [67] and computational results [48].
Furthermore, the overall trend as presented in Figure 4 shows that the E f of B2, B19, and B19 decline almost linearly with the increasing of Pt doping concentration, which indicates that the stability of Ti 50 Ni 50 x Pt x alloys becomes better when Ni is replaced by Pt. This can be explained by the electronic hybridization of Pt with Ti and Ni (see Figure 5), and will be discussed in more detail in Section 3.3. We can also note that B2 exhibits the highest formation energy which is one of the most unstable structures among all the structures in the TiNiPt system.
Finally, as shown in Figure 4, when Pt content is between 0 and 12.5 at.%, E f of monoclinic B19 phases are all negative and lower than that of orthorhombic B19 and B2 phases, thus the most stable phase is the monoclinic martensitic crystal structure. However, as Pt content reaches to near 15 at.%, the orthorhombic phases are more stable compared with the monoclinic phases. It can be seen that the E f curve of B19 is a little steeper than that of B19 , thus a crossover exists for Pt content about 15 at.%, which can be attributed to the different bond lengths of Ni-Ti in B19 and B19 . The shortest Ni-Ti bond length of B19 (2.514 Å) is smaller than that of B19 (2.564 Å), which leads to slightly more repulsive interaction in B19 compared to that of B19 with the doping of Pt atoms.
Therefore, it can be expected that the transformation path would be B2 to B19 at Pt content less than about 15 at.%, while B2 to B19 at Pt content greater than 15 at.%. As the experimental observations [10] show that, at ≤ 10 at.% Pt, the martensite structure is monoclinic B19 , at higher levels of Pt, at least 16 at.% or greater, the martensite orthorhombic B19 is formed. This means that our calculated result is consistent with the experimental result.

3.3. Electronic Structures

In order to investigate the electronic mechanism behind the martensitic transformation pathways, the partial density of states (PDOSs) of the B19 and B19 structures of Ti 50 Ni 50 x Pt x alloys are calculated and plotted in Figure 5. For NiTi, the electronic states below the Fermi level are mostly contributed by the Ni d states, while electronic states above the Fermi level come mainly from Ti d states. This feature is consistent with other DFT study [46]. Furthermore, the hybridization between Ti d and Ni d states in the structure of B19 is stronger than that of B19. Moreover, compared with the Ti d of B19 NiTi, the high energy peaks above Fermi level of B19 shift towards the Fermi level. This suggests that B19 is more stable than B19 in NiTi, in accordance with the results of E f (see Figure 4).
Figure 5 shows that, due to the Pt dopants, a new set of peaks centered at −4.5 eV emerges in the PDOSs of the B19 and B19 of Ti 50 Ni 50 x Pt x serials. Moreover, these set of peaks become stronger and broader with the increase of Pt content. The doping of Pt gives rise to the strong resonant states between the Ni and its nearest neighbor Pt atoms in the energy region from −3 eV to −1.8 eV, and enhances the interaction between Ti and Ni (Pt) atoms. The peaks of Ti atoms and Ni-Pt resonated atoms shift towards lower energies slightly. These indicate that both the B19 and B19 phases become more stable as the Pt increases, in line with the data of E f . The electronic interaction in B19 is slightly stronger than that of in B19.
However, for the x = 18.75 alloy, the peaks of Ni d states below the Fermi level located at about −1.8 eV in the PDOSs of B19 are higher than those of B19 , which implies that the hybridization between Ni (Pt) d and Ti d states in the B19 phase is stronger than that of B19 . This reveals that the B19 orthorhombic phase is more stable compared with the B19 monoclinic phase, which is in accordance with the formation energy results that the orthorhombic phase is the martensitic stable structure for Ti 50 Ni 50 x Pt x alloy, as x > 15.

3.4. Energy Differences between the Austenite and Martensite

After studying the phase transition path of Ti 50 Ni 50 x Pt x alloys (x = 0, 6.25, 8.33, 10.42, 12.5, 18.75, 25), we set out to research the concentration dependence of M s of these alloys using the density functional theory. At first, we invoke the total energy differences, Δ E , between B2 and martensite (B19 or B19 ) phases. According to the previous studies [47,55,58,59,60], there is a strong relationship between M s and Δ E from first-principles calculations, i.e., a larger Δ E corresponds to a higher M s . All these energies are normalized per atom. They are calculated (per atom) as:
Δ E B 19 = E B 2 E B 19
Δ E B 19 = E B 2 E B 19
with E B 2 , E B 19 , and E B 19 being the total energies per atom of B2, B19, and B19 structures of Ti 50 Ni 50 x Pt x alloys, respectively.
At first, the energy differences Δ E B 19 = E B 2 E B 19 = 34 meV/atom and Δ E B 19 = E B 2 E B 19 = 44 meV/atom for an equiatomic binary NiTi alloy are obtained, which is consistent with other theoretical values [48,61,64,68]. The Δ E B 19 and Δ E B 19 are both positive, and the Δ E B 19 is a little higher than Δ E B 19 which revealed that the monoclinic B19 is more energetically stable than orthorhombic B19 structures.
Then, the calculated concentration dependence of Δ E B 19 and Δ E B 19 for Ti 50 Ni 50 x Pt x alloys (x = 0, 6.25, 8.33, 10.42, 12.5, 18.75, 25) are presented in Figure 6. Finally, in order to confirm the experimental observations of alloy composition dependence of M s in Figure 7, we plot the total energy differences, Δ E , as well as M s as a function of the Pt concentration in the Ti 50 Ni 50 x Pt x alloys. M s of Ni-Ti-Pt is cited from Ref. [35].
As presented in Figure 6, the Δ E B 19 and Δ E B 19 are both positive. There is a crossover at the Δ E curve for Pt near 15 at.% concentration, which corresponds to that of E f in B19 and B19 (see Figure 4), most likely due to the marginal difference in the Ni-Ti bond length previously mentioned. For Pt < 15 at.%, the Δ E B 19 is higher than Δ E B 19 , which indicates that the monoclinic B19 phase is more stable than orthorhombic B19 structures; whereas for Pt > 15 at.%, the Δ E B 19 is higher than Δ E B 19 , which suggests orthorhombic B19 structures are more stable than the monoclinic B19 phase. These are in line with the present computed results of E f of Ti 50 Ni 50 x Pt x alloys (x = 0, 6.25, 8.33, 10.42, 12.5, 18.75, 25) (see Figure 4).
According to the data of Δ E of the most stable martensite phase, Δ E and M s of Ti 50 Ni 50 x Pt x alloys are presented in Figure 7 as a function of Pt content. First and foremost, it can be found from Figure 7 that the overall trends of the Δ E and M s of Ti 50 Ni 50 x Pt x alloys with the increasing of the Pt concentration are remarkably similar to each other, which confirms that the present calculated results are fully in line with the experimental observations. For Pt ≤ 8.33 at.%, martensite start temperatures and Δ E are less sensitive to composition, and decrease slightly with a minimum observed at 6.25 at.% Pt; for 8.33 at.% < Pt < 12.5 at.%, Δ E and M s increase slightly. However, at higher levels of Pt, especially exceeding about 15 at.%, Δ E and M s increase sharply with Pt content. This can be explained as follows, with the dramatically increasing of M s , correspondingly, more energy is needed for the martensitic transformation. Taking into account good coincidences with the results of experimental studies, the resulting model can be a good tool for simplifying and rationalizing experimental studies of the NiTiPt system.

4. Conclusions

In the present paper we explored the strong dependence of the martensite start temperature M s on alloy composition in ternary Ni-Ti-Pt shape memory alloys by means of density functional theory. Based on the crystal structure optimizations of Ti 50 Ni 50 x Pt x (x = 0, 6.25, 8.33, 10.42, 12.5, 18.75, 25) shape memory alloys, Pt additions substituted for Ni in NiTi with a supercell approach are calculated. From the results obtained in the present work the following conclusions can be drawn:
The lattice parameters a, b, c and supercell volume of Ti 50 Ni 50 x Pt x alloys increase with the increasing of Pt content. Since the number of valence electrons of Pt is equal to that of Ni, the larger atomic radius of Pt compared to Ni results in a bigger supercell volume.
As the calculated results of formation energies presented, for Pt > 15 at.%, the monoclinic B19 martensite crystal structure becomes unstable and an orthorhombic B19 crystal structure is formed, resulting from a gradual destabilization/stabilization of B19 /B19, which is in excellent agreement with the experimental data.
The computational results of the total energy differences Δ E between parent and martensite phases are as follows: for Pt < 10 at.%, the substitution of Pt for Ni has relatively little effect on Δ E (with a minimum at 6.25 at.% Pt), which shows only a moderate dependence of Δ E on alloy composition; for Pt ≥ 10 at.%, the Δ E increase linearly with increasing Pt content, which brings about the dramatical enhancement of M s .

Author Contributions

Conceptualization, X.Y. and J.S.; methodology, X.Y.; writing—original draft preparation, X.Y.; and writing—review and editing, X.Y. and J.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (NSFC, grants no. 51371017 and no. 11864047), and the Natural Science Foundation of Technology Department (QKHLH [2015]7021).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Buehler, W.J.; Wiley, R.C. TiNi-ductile intermetallic compound. ASM-Trans. 1962, 55, 269–276. [Google Scholar]
  2. Otsuka, K.; Ren, X. Physical metallurgy of Ti-Ni-based shape memory alloys. Prog. Mater. Sci. 2005, 50, 511–678. [Google Scholar] [CrossRef]
  3. Guda Vishnu, K.; Strachan, A. Size effects in NiTi from density functional theory calculations. Phys. Rev. B 2012, 85, 014114. [Google Scholar] [CrossRef] [Green Version]
  4. Haskins, J.B.; Lawson, J.W. Finite temperature properties of NiTi from first principles simulations: Structure, mechanics, and thermodynamics. J. Appl. Phys. 2017, 121, 205103. [Google Scholar] [CrossRef]
  5. Kumar, P.; Waghmare, U.V. First-principles phonon-based model and theory of martensitic phase transformation in NiTi shape memory alloy. Materialia 2020, 9, 100602. [Google Scholar] [CrossRef]
  6. Lang, L.; Payne, A.; Valencia-Jaime, I.; Verstraete, M.J.; Bautista-Hernández, A.; Romero, A.H. Assessing Nickel Titanium Binary Systems Using Structural Search Methods and Ab Initio Calculations. J. Phys. Chem. C 2021, 125, 1578–1591. [Google Scholar] [CrossRef]
  7. Zarkevich, N.A.; Johnson, D.D. Stable atomic structure of NiTi austenite. Phys. Rev. B 2014, 90, 060102. [Google Scholar] [CrossRef] [Green Version]
  8. Frenzel, J.; Wieczorek, A.; Opahle, I.; Maaß, B.; Drautz, R.; Eggeler, G. On the effect of alloy composition on martensite start temperatures and latent heats in Ni-Ti-based shape memory alloys. Acta Mater. 2015, 90, 213–231. [Google Scholar] [CrossRef]
  9. Bozzolo, G.; Noebe, R.D.; Mosca, H.O. Site preference of ternary alloying additions to NiTi: Fe, Pt, Pd, Au, Al, Cu, Zr and Hf. J. Alloys Compd. 2005, 389, 80–94. [Google Scholar] [CrossRef] [Green Version]
  10. Ma, J.; Karaman, I.; Noebe, R.D. High temperature shape memory alloys. Int. Mater. Rev. 2010, 55, 257–315. [Google Scholar] [CrossRef]
  11. Zarinejad, M.; Liu, Y.; White, T.J. The crystal chemistry of martensite in NiTiHf shape memory alloys. Intermetallics 2008, 16, 876–883. [Google Scholar] [CrossRef]
  12. Bigelow, G.S.; Padula, S.A.; Garg, A.; Gaydosh, D.; Noebe, R.D. Characterization of Ternary NiTiPd High-Temperature Shape-Memory Alloys under Load-Biased Thermal Cycling. Metall. Mater. Trans. A 2010, 41, 3065–3079. [Google Scholar] [CrossRef] [Green Version]
  13. Inoue, S.; Sawada, N.; Namazu, T. Effect of Zr content on mechanical properties of Ti-Ni-Zr shape memory alloy films prepared by dc magnetron sputtering. Vacuum 2009, 83, 664–667. [Google Scholar] [CrossRef]
  14. Bozzolo, G.; Mosca, H.O.; del Grosso, M.F. Energy of formation, lattice parameter and bulk modulus of (Ni, X)Ti alloys with X = Fe, Pd, Pt, Au, Al, Cu, Zr, Hf. Intermetallics 2008, 16, 668–675. [Google Scholar] [CrossRef]
  15. Mosca, H.O.; Bozzolo, G.; del Grosso, M.F. Atomistic modeling of ternary additions to NiTi and quaternary additions to Ni-Ti-Pd, Ni-Ti-Pt and Ni-Ti-Hf shape memory alloys. Phys. B Conden. Matter 2012, 407, 3244–3247. [Google Scholar] [CrossRef]
  16. Potapov, P.L.; Shelyakov, A.V.; Gulyaev, A.A.; Svistunov, E.L.; Matveeva, N.M.; Hodgson, D. Effect of Hf on the structure of Ni-Ti martensitic alloys. Mater. Lett. 1997, 32, 247–250. [Google Scholar] [CrossRef]
  17. Suresh, K.S.; Kim, D.I.; Bhaumik, S.K.; Suwas, S. Evolution of microstructure and texture in Ni49.4Ti38.6Hf12 shape memory alloy during hot rolling. Intermetallics 2013, 42, 1–8. [Google Scholar] [CrossRef]
  18. Hsieh, S.F.; Wu, S.K. A Study on Ternary Ti-rich TiNiZr Shape Memory Alloys. Mater Charact. 1998, 41, 151–162. [Google Scholar] [CrossRef]
  19. Koenig, D.; Zarnetta, R.; Savan, A.; Brunken, H.; Ludwig, A. Phase transformation, structural and functional fatigue properties of Ti-Ni-Hf shape memory thin film. Acta Mater. 2011, 59, 3267–3275. [Google Scholar] [CrossRef]
  20. Meng, X.L.; Cai, W.; Chen, F.; Zhao, L.C. Effect of aging on martensitic transformation and microstructure in Ni-rich TiNiHf shape memory alloy. Scr. Mater. 2006, 54, 1599–1604. [Google Scholar] [CrossRef]
  21. Moshref-Javadi, M.; Seyedein, S.H.; Salehi, M.T.; Aboutalebi, M.R. Age-induced multi-stage transformation in a Ni-rich NiTiHf alloy. Acta Mater. 2013, 61, 2583–2594. [Google Scholar] [CrossRef]
  22. Tong, Y.; Chen, F.; Tian, B.; Li, L.; Zheng, Y. Microstructure and martensitic transformation of Ti49Ni51−xHfx high temperature shape memory alloys. Mater. Lett. 2009, 63, 1869–1871. [Google Scholar] [CrossRef]
  23. Tong, Y.; Liu, Y.; Miao, J. Phase transformation in NiTiHf shape memory alloy thin films. Thin Solid Films 2008, 516, 5393–5396. [Google Scholar] [CrossRef]
  24. Tong, Y.; Liu, Y.; Miao, J.; Zhao, L. Characterization of a nanocrystalline NiTiHf high temperature shape memory alloy thin film. Scr. Mater. 2005, 52, 983–987. [Google Scholar] [CrossRef]
  25. Acar, E.; Karaca, H.E.; Tobe, H.; Noebe, R.D.; Chumlyakov, Y.I. Characterization of the shape memory properties of a Ni45.3Ti39.7Hf10Pd5 alloy. J. Alloys Compd. 2013, 578, 297–302. [Google Scholar] [CrossRef]
  26. Belbasi, M.; Salehi, M.T.; Mousavi, S.A.A.A.; Ebrahimi, S.M. A study on the mechanical behavior and microstructure of NiTiHf shape memory alloy under hot deformation. Mater. Sci. Eng. A 2013, 560, 96–102. [Google Scholar] [CrossRef]
  27. Meng, X.L.; Cai, W.; Fu, Y.D.; Li, Q.F.; Zhang, J.X.; Zhao, L.C. Shape-memory behaviors in an aged Ni-rich TiNiHf high temperature shape-memory alloy. Intermetallics 2008, 16, 698–705. [Google Scholar] [CrossRef]
  28. Meng, X.L.; Zheng, Y.F.; Cai, W.; Zhao, L.C. Two-way shape memory effect of a TiNiHf high temperature shape memory alloy. J. Alloys Compd. 2004, 372, 180–186. [Google Scholar] [CrossRef]
  29. Saghaian, S.M.; Karaca, H.E.; Tobe, H.; Souri, M.; Noebe, R.; Chumlyakov, Y.I. Effects of aging on the shape memory behavior of Ni-rich Ni50.3Ti29.7Hf20 single crystals. Acta Mater. 2015, 87, 128–141. [Google Scholar] [CrossRef]
  30. Karaca, H.E.; Saghaian, S.M.; Ded, G.; Tobe, H.; Basaran, B.; Maier, H.J.; Noebe, R.D.; Chumlyakov, Y.I. Effects of nanoprecipitation on the shape memory and material properties of an Ni-rich NiTiHf high temperature shape memory alloy. Acta Mater. 2013, 61, 7422–7431. [Google Scholar] [CrossRef]
  31. Meng, X.L.; Cai, W.; Zheng, Y.F.; Zhao, L.C. Phase transformation and precipitation in aged Ti-Ni-Hf high-temperature shape memory alloys. Mater. Sci. Eng. A 2006, 438-440, 666–670. [Google Scholar] [CrossRef]
  32. Santamarta, R.; Arróyave, R.; Pons, J.; Evirgen, A.; Karaman, I.; Karaca, H.E.; Noebe, R.D. TEM study of structural and microstructural characteristics of a precipitate phase in Ni-rich Ni-Ti-Hf and Ni-Ti-Zr shape memory alloys. Acta Mater. 2013, 61, 6191–6206. [Google Scholar] [CrossRef]
  33. Yang, F.; Coughlin, D.R.; Phillips, P.J.; Yang, L.; Devaraj, A.; Kovarik, L.; Noebe, R.D.; Mills, M.J. Structure analysis of a precipitate phase in an Ni-rich high-temperature NiTiHf shape memory alloy. Acta Mater. 2013, 61, 3335–3346. [Google Scholar] [CrossRef]
  34. Yang, F.; Kovarik, L.; Phillips, P.J.; Noebe, R.D.; Mills, M.J. Characterizations of precipitate phases in a Ti-Ni-Pd alloy. Scr. Mater. 2012, 67, 145–148. [Google Scholar] [CrossRef]
  35. Rios, O.; Noebe, R.D.; Biles, T.; Garg, A.; Palczer, A.; Scheiman, D.; Seifert, H.J.; Kaufman, M. Characterization of Ternary NiTiPt High-Temperature Shape Memory Alloys. Proc. SPIE 2005, 5761, 376–387. [Google Scholar]
  36. Zarnetta, R.; Takahashi, R.; Young, M.L.; Savan, A.; Furuya, Y.; Thienhaus, S.; Maaß, B.; Rahim, M.; Frenzel, J.; Brunken, H.; et al. Identification of quaternary shape memory alloys with near-zero thermal hysteresis and unprecedented functional stability. Adv. Funct. Mater. 2010, 20, 1917–1923. [Google Scholar] [CrossRef]
  37. Zhang, Z.; James, R.D.; Müller, S. Energy barriers and hysteresis in martensitic phase transformations. Acta Mater. 2009, 57, 4332–4352. [Google Scholar] [CrossRef] [Green Version]
  38. Meisner, L.L.; Sivokha, V.P. The effect of applied stress on the shape memory behavior of TiNi-based alloys with different consequences of martensitic transformations. Phys. B Conden. Matter 2004, 344, 93–98. [Google Scholar] [CrossRef]
  39. Smialek, J.L.; Humphrey, D.L.; Noebe, R.D. Comparative Oxidation Kinetics of a NiPtTi High Temperature Shape Memory Alloy. Oxid. Met. 2010, 74, 125–144. [Google Scholar] [CrossRef]
  40. Chovan, D.; Nolan, M.; Tofail, S.A.M. First principles simulations of elastic properties of radiopaque NiTiPt. J. Alloys Compd. 2015, 630, 54–59. [Google Scholar] [CrossRef]
  41. Tan, C.L.; Tian, X.H.; Ji, G.J.; Gui, T.L.; Cai, W. Elastic property and electronic structure of TiNiPt high-temperature shape memory alloys. Solid State Commun. 2008, 147, 8–10. [Google Scholar] [CrossRef]
  42. Gou, L.; Liu, Y.; Ng, T.Y. An investigation on the crystal structures of Ti50Ni50−xCux shape memory alloys based on density functional theory calculations. Intermetallics 2014, 53, 20–25. [Google Scholar] [CrossRef]
  43. Hu, Q.M.; Yang, R.; Lu, J.M.; Wang, L.; Johansson, B.; Vitos, L. Effect of Zr on the properties of (TiZr)Ni alloys from first-principles calculations. Phys. Rev. B 2007, 76, 224201. [Google Scholar] [CrossRef]
  44. Ma, L.; Wang, X.; Shang, J.X. Effect of Pd in NiTi on the martensitic transformation temperatures and hysteresis: A first-principles study. Acta Phys. Sin. 2014, 63, 233103. [Google Scholar]
  45. Singh, N.; Talapatra, A.; Junkaew, A.; Duong, T.; Gibbons, S.; Li, S.; Thawabi, H.; Olivos, E.; Arróyave, R. Effect of ternary additions to structural properties of NiTi alloys. Comput. Mater. Sci. 2016, 112, 347–355. [Google Scholar] [CrossRef] [Green Version]
  46. Tan, C.L.; Cai, W.; Tian, X.H. First-principles study on the effect of Hf content on martensitic transformation temperature of TiNiHf alloy. Chin. Phys. 2006, 15, 2718–2723. [Google Scholar]
  47. Yang, X.; Ma, L.; Shang, J. Martensitic transformation of Ti50(Ni50−xCux) and Ni50(Ti50−xZrx) shape-memory alloys. Sci. Rep. 2019, 9, 3221. [Google Scholar] [CrossRef] [Green Version]
  48. Kibey, S.; Sehitoglu, H.; Johnson, D.D. Energy landscape for martensitic phase transformation in shape memory NiTi. Acta Mater. 2009, 57, 1624–1629. [Google Scholar] [CrossRef]
  49. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  50. Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  51. Kresse, G.; Hafner, J. Ab initio molecular dynamics for open-shell transition metals. Phys. Rev. B 1993, 48, 13115–13118. [Google Scholar] [CrossRef]
  52. Perdew, J.P.; Wang, Y. Accurate and simple analytic representation of the electron-gas correlation energy. Phys. Rev. B 1992, 45, 13244–13249. [Google Scholar] [CrossRef] [PubMed]
  53. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  54. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  55. Chen, J.; Li, Y.; Shang, J.; Xu, H. First principles calculations on martensitic transformation and phase instability of Ni-Mn-Ga high temperature shape memory alloys. Appl. Phys. Lett. 2006, 89, 231921. [Google Scholar] [CrossRef]
  56. Alonso, P.; Rubiolo, G. Relative stability of bcc structures in ternary alloys with Ti50Al25Mo25 composition. Phys. Rev. B 2000, 62, 237. [Google Scholar] [CrossRef]
  57. Holec, D.; Friák, M.; Dlouhy, A.; Neugebauer, J. Ab initio study of point defects in NiTi-based alloys. Phys. Rev. B 2014, 89, 014110. [Google Scholar] [CrossRef] [Green Version]
  58. Yamaguchi, K.; Ishida, S.; Asano, S. Electron concentration and structural transformation of Ni2MnGa-based shape memory alloys. Mater. Trans. 2002, 43, 846–851. [Google Scholar] [CrossRef] [Green Version]
  59. Chakrabarti, A.; Biswas, C.; Banik, S.; Dhaka, R.S.; Shukla, A.K.; Barman, S.R. Influence of Ni doping on the electronic structure of Ni2MnGa. Phys. Rev. B 2005, 72, 073103. [Google Scholar] [CrossRef] [Green Version]
  60. Wang, X.; Shang, J.X.; Wang, F.H.; Chen, Y. Origin of the strain glass transition in Ti50(Ni50−xDx) alloys. J. Alloys Compd. 2016, 678, 325–328. [Google Scholar] [CrossRef]
  61. Huang, X.; Ackland, G.J.; Rabe, K.M. Crystal structures and shape-memory behaviour of NiTi. Nat. Mater. 2003, 2, 307–311. [Google Scholar] [CrossRef] [PubMed]
  62. Kudoh, Y.; Tokonami, M.; Miyazaki, S.; Otsuka, K. Crystal-structure of the martensite in Ti-49.2at%Ni alloy analyzed by the single-crystal X-ray-diffration method. Acta Metall. 1985, 33, 2049–2056. [Google Scholar] [CrossRef]
  63. Philip, T.; Beck, P.A. CsCl-type ordered structures in binary alloys of transition elements. Trans. AIME J. Metal. 1957, 209, 1269–1271. [Google Scholar] [CrossRef]
  64. Zhang, J.M.; Guo, G.Y. Microscopic Theory of the Shape Memory Effect in TiNi. Phys. Rev. Lett. 1997, 78, 4789–4792. [Google Scholar] [CrossRef]
  65. Simon, A. Intermetallic Compounds and the Use of Atomic Radii in Their Description. Angew. Chem. Int. Ed. 1983, 22, 95–113. [Google Scholar] [CrossRef]
  66. Zarinejad, M.; Liu, Y. Dependence of Transformation Temperatures of NiTi-based Shape-Memory Alloys on the Number and Concentration of Valence Electrons. Adv. Funct. Mater. 2008, 18, 2789–2794. [Google Scholar] [CrossRef]
  67. Otsuka, K.; Ren, X. Recent developments in the research of shape memory alloys. Intermetallics 1999, 7, 511–528. [Google Scholar] [CrossRef]
  68. Ye, Y.; Chan, C.; Ho, K. Structural and electronic properties of the martensitic alloys TiNi, TiPd, and TiPt. Phys. Rev. B 1997, 2, 8–22. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Ni occupancy in the austenite B2 3 × 2 × 2 supercell Ni 24 Ti 24 structure. Smaller blue spheres denote Ni atoms; larger gray spheres denote Ti atoms. The 3 × 2 × 2 supercells of B19 and B19 are also used, and structures are similar to that of B2.
Figure 1. Ni occupancy in the austenite B2 3 × 2 × 2 supercell Ni 24 Ti 24 structure. Smaller blue spheres denote Ni atoms; larger gray spheres denote Ti atoms. The 3 × 2 × 2 supercells of B19 and B19 are also used, and structures are similar to that of B2.
Applsci 11 06878 g001
Figure 2. Research schematic diagram.
Figure 2. Research schematic diagram.
Applsci 11 06878 g002
Figure 3. Pt content dependence of (a) ‘a’ lattice parameter, (b) ‘b’ lattice parameter (c) ‘c’ lattice parameter, and (d) unit cell volume in B2, B19 and B19 structures of Ti 50 Ni 50 x Pt x alloys.
Figure 3. Pt content dependence of (a) ‘a’ lattice parameter, (b) ‘b’ lattice parameter (c) ‘c’ lattice parameter, and (d) unit cell volume in B2, B19 and B19 structures of Ti 50 Ni 50 x Pt x alloys.
Applsci 11 06878 g003
Figure 4. The formation energies, E f of the austenite B2, martensite B19 and B19 phases as a function of Pt content in ternary Ti 50 Ni 50 x Pt x alloys.
Figure 4. The formation energies, E f of the austenite B2, martensite B19 and B19 phases as a function of Pt content in ternary Ti 50 Ni 50 x Pt x alloys.
Applsci 11 06878 g004
Figure 5. Partial density of states (PDOSs) of martensite (a) B19 and (b) B19 phase of Ti 50 Ni 50 x Pt x (x = 0, 12.5, 18.75) alloys for the Ti, Ni, and Pt site. The vertical line denotes the Fermi level, which is located at 0 eV.
Figure 5. Partial density of states (PDOSs) of martensite (a) B19 and (b) B19 phase of Ti 50 Ni 50 x Pt x (x = 0, 12.5, 18.75) alloys for the Ti, Ni, and Pt site. The vertical line denotes the Fermi level, which is located at 0 eV.
Applsci 11 06878 g005
Figure 6. The energy differences between austenite and martensite ( Δ E ) as a function of the doping concentration in Ti 50 Ni 50 x Pt x alloys.
Figure 6. The energy differences between austenite and martensite ( Δ E ) as a function of the doping concentration in Ti 50 Ni 50 x Pt x alloys.
Applsci 11 06878 g006
Figure 7. Compositional dependence of Δ E and phase transformation start temperature ( M s ) of Ti 50 Ni 50 x Pt x alloys. The data of M s refer to the right axis and those of Δ E to the left axis. M s is cited from Ref. [35].
Figure 7. Compositional dependence of Δ E and phase transformation start temperature ( M s ) of Ti 50 Ni 50 x Pt x alloys. The data of M s refer to the right axis and those of Δ E to the left axis. M s is cited from Ref. [35].
Applsci 11 06878 g007
Table 1. The site preference of Pt additions in the austenite B2 3 × 2 × 2 supercell structure at different doping concentrations, n is the number of Pt atoms in the supercell.
Table 1. The site preference of Pt additions in the austenite B2 3 × 2 × 2 supercell structure at different doping concentrations, n is the number of Pt atoms in the supercell.
nPt Content (at.%)Site Occupation of Pt
36.258, 11, 24
48.338, 11, 24, 18
510.425, 8, 11, 24, 18
612.502, 5, 8, 11, 24, 18
918.753, 5, 8, 9, 11, 16, 17, 22, 24
1225.003, 5, 6, 8, 10, 11, 14, 16, 17, 19, 22, 24
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yang, X.; Shang, J. Ab Initio Study of Martensitic Transformation in NiTiPt High Temperature Shape Memory Alloys. Appl. Sci. 2021, 11, 6878. https://doi.org/10.3390/app11156878

AMA Style

Yang X, Shang J. Ab Initio Study of Martensitic Transformation in NiTiPt High Temperature Shape Memory Alloys. Applied Sciences. 2021; 11(15):6878. https://doi.org/10.3390/app11156878

Chicago/Turabian Style

Yang, Xiaolan, and Jiaxiang Shang. 2021. "Ab Initio Study of Martensitic Transformation in NiTiPt High Temperature Shape Memory Alloys" Applied Sciences 11, no. 15: 6878. https://doi.org/10.3390/app11156878

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