Next Article in Journal
Two-Dimensional Crystalline Gridding Networks of Hybrid Halide Perovskite for Random Lasing
Previous Article in Journal
Texture and Microstructure Evolution of Ultra-High Purity Cu-0.1Al Alloy under Different Rolling Methods
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Tunable Martensitic Transformation and Magnetic Properties of Sm-Doped NiMnSn Ferromagnetic Shape Memory Alloys

1
Department of Physics, University of Central Punjab, Sargodha 40100, Pakistan
2
School of Materials Science and Engineering, Nanjing University of Science and Technology, Nanjing 210000, China
3
Department of Physics, University of Kotli Azad Jammu and Kashmir, Kotli 11100, Pakistan
4
Department of Physics, University of Lahore, Lahore 40050, Pakistan
5
Applied Physics, Computer and Instrumentation Centre, PCSIR Laboratories Complex Ferozpur Road, Lahore 54600, Pakistan
6
Department of Electrical Engineering, Sejong University, 209 Neungdong-ro, Seoul 05006, Korea
7
Division of Global Business, University College, Konkuk University, Seoul 05029, Korea
8
Department of Physics, Konkuk University, Seoul 05029, Korea
*
Authors to whom correspondence should be addressed.
Crystals 2021, 11(9), 1115; https://doi.org/10.3390/cryst11091115
Submission received: 12 August 2021 / Revised: 8 September 2021 / Accepted: 9 September 2021 / Published: 13 September 2021

Abstract

:
NiMnSn ferromagnetic shape memory alloys exhibit martensitic transformation at low temperatures, restricting their applications. Therefore, this is a key factor in improving the martensitic transformation temperature, which is effectively carried out by proper element doping. In this research, we investigated the martensitic transformation and magnetic properties of Ni43Mn46-x SmxSn11 (x = 0, 1, 2, 3) alloys on the basis of structural and magnetic measurements. X-ray diffraction showed that the crystal structure transforms from the cubic L21 to the orthorhombic martensite and gamma (γ) phases. The reverse martensitic and martensitic transformations were indicated by exothermic and endothermic peaks in differential scanning calorimetry. The martensitic transformation temperature increased considerably with Sm doping and exceeded room temperature for Sm = 3 at. %. The Ni43Mn45SmSn11 alloy exhibited magnetostructural transformation, leading to a large magnetocaloric effect near room temperature. The existence of thermal hysteresis and the metamagnetic behavior of Ni43Mn45SmSn11 confirm the first-order magnetostructural transition. The magnetic entropy change reached 20 J·kg−1·K−1 at 266 K, and the refrigeration capacity reached ~162 J·Kg−1, for Ni43Mn45SmSn11 under a magnetic field variation of 0–5 T.

1. Introduction

The materials showing martensitic transformation (MT) exhibit various multifunctional phenomena, such as the magnetocaloric effect (MCE) [1,2], exchange bias (EB) [3,4,5], magnetothermal conductivity (MC) [6,7] and magnetoresistance (MR) [8,9,10]. This coupled transition is obtained around the MT temperature (Tt). These multifunctional properties show application in vast areas such as magnetic refrigeration, energy harvesting, magneto-mechanical devices and sensors [11,12,13,14]. The MCE is a magneto-thermodynamic phenomenon in which the temperature is changed in the material when exposed to an external non-constant magnetic field [15]. MCE-based magnetic refrigeration has many advantages over conventional gas cooling technology, such as being environmentally friendly, having high refrigerant efficiency and low cast measures, occupying less space, possessing low mechanical vibration and being harmless [16,17,18,19]. MCE materials are utilized in magnetic refrigeration technology. The materials showing first-order magnetic transitions are widely used for magnetic cooling and investigated due to their large MCE, e.g., LeFeSi [20,21,22], MnFePAs [23], GdSiGe [24], MMnX (M = Ni or Co, X = Ge or Si) [25] and NiMnZ (Z = Sn, In, Sb) ferromagnetic shape memory alloys (FMSMAs) [26].
FMSMAs exhibit first-order magnetostructural transformation from high-temperature austenite to low-temperature martensite upon cooling, with considerable thermal hysteresis and other effects such as lattice distortion and abrupt changes in magnetization. The magnetic field induces a first-order magnetostructural transition in these alloys, generating the MCE [27]. It is promising to tune the Tt in a wide temperature range, endowing FMSMAs with the ability to exhibit large multifunctional effects under a low applied magnetic field for magnetic applications. It is reported that the Tt can be tuned by changing the valence electron concentration (e/a) [28], grain size [29], external parameters such as the magnetic field and applied pressure [30], volume of the unit cell [14] and heat treatment [31], while adjustable elemental composition and preparation conditions are very important to achieve the MCE for the desired applications [25,26,27].
NiMnSn-based alloys undergo a reverse martensitic transformation due to magnetostructural coupling and exhibit a variety of multifunctional properties, which indicates a potential for many applications. The utilization of MCE materials strongly depends upon the mechanical properties and magnetostructural coupling. Unfortunately, the highly brittle nature and poor mechanical properties of NiMnSn alloys are major drawbacks. Additionally, NiMnSn alloys show a low working temperature due to coupling between MT and ferromagnetic transition [32].
Only a few studies have reported on the enhancement of the working temperature of these alloys, which is the main point of application. Moreover, it is well known that a large MCE has been reported in rare-earth metal-based materials, i.e., Gd5Si2Ge2 [24]. Meanwhile, it has also been reported that the addition of rare-earth elements could enhance the Tt in these types of alloys. Therefore, it can be anticipated that the addition of proper rare-earth elements in NiMnSn FMSMAs can tune the Tt to a higher temperature, resulting in the MCE near room temperature (RT). There is still a lack of information and knowledge about the rare-earth element Sm after it has been doped. In this research, we doped Sm in Ni43Mn46-x SmxSn11 (x = 0, 1, 2, 3) alloys which increases the Tt considerably, and a large MCE is obtained near RT for the Ni43Mn45SmSn11 alloy sample.

2. Materials and Methods

Polycrystalline Ni43Mn46-x SmxSn11 (x = 0, 1, 2, 3) alloy samples were prepared by arc melting 99.99% pure Ni, Mn, Sn and Sm using a water-cooled copper crucible in an argon atmosphere. The melting process was repeated four times to ensure the purity and uniformity of the alloys. Homogeneity is an important factor to enhance the properties of the materials [31]. Therefore, ingots were annealed at 1123 K for 2 days followed by quenching in ice water. The elemental composition of the samples was determined by X-ray energy-dispersive spectroscopy (EDS, Thermo system 7, Malvern UK), and powder X-ray diffraction (XRD, Bruker, D8 Advance, Madison, WI, US) was used to identify the crystal structure at room temperature. The structural transitions were investigated using differential scanning calorimetry (Mettler Toledo, DSC 3, Nanterre, France), while the magnetic properties were examined by a magnetic property measurement system (MPMS-7, Quantum Design, San Diego, CA, USA) and physical property measurement system (PPMS, Quantum Design, Dynacool, San Diego, CA, USA). The so-called loop method was used to measure isothermal magnetization (MH) curves to avoid irreversibility [33].

3. Results and Discussion

Energy-dispersive X-ray spectroscopy (EDS) was used to check the appropriate amount and homogeneity of the constituent elements in the prepared alloys. The EDS results of the Ni43Mn46–xSmxSn11 (x = 0, 1, 2, 3) alloys are shown in Table 1, confirming the presence of all precursor elements. The obtained compositions of the alloys were found to be very close to the nominal composition. Here, a small deviation in the composition of the elements is very common due to various factors such as weighing the precursor elements, melting and annealing. The electronic concentration (e/a) was measured using Equation (1) [34].
e/a = (7 × at. % Mn + 10 × at. % Ni + 4 × at. % Sn + 8 × at. % Sm)/10
It is well known that martensites with twin boundaries play an important role in the magnetic field-induced strain of FMSMAs, depending upon the twin boundary migration. The X-ray diffraction patterns of the Ni43Mn46-x SmxSn11 (x = 0, 1, 2, 3) alloy samples are shown in Figure 1.
The reflection peaks corresponding to the cubic austenite phase are indicated by green squares. On comparing the NiMnSn alloys reported by Krenke et al. [35], the main reflection peaks corresponding to the martensitic phase are indicated by red circles, and extra peaks which indicate the presence of an extra γ phase are represented by blue triangles. The XRD pattern for x = 0 shows typical diffraction peaks from the austenite phase, indicating a single cubic L21 crystal structure. As MT is closely linked to the crystal structure, the presence of the L21 cubic structure for x = 0 indicates that its Tt is below the RT. For the x = 1 and 2 Sm-doped samples, a mixed complicated structure of the cubic L21 phase, orthorhombic martensite phase and γ phase is exhibited. The coexistence of austenite and martensite phases indicates that the Tt lies near RT for these samples. Ni43Mn43Sn11Sm3 presents a two-phase structure, i.e., orthorhombic martensite phase and γ phase. The disappearance of the austenite phase in Ni43Mn43Sm3Sn11 indicates that the Tt is increased from RT. This is in good agreement with the DSC measurement, which is discussed in the next session. For x = 0, the alloy sample crystallizes in the cubic L21 structure with lattice parameters a = b = c = 5.986 Å. In the case of x = 1 and 2, there is the appearance of an orthorhombic martensite phase with dominating cubic austenite with lattice parameters a = b = c = 5.784 Å. It is well known that the crystal structure of NiMnSn alloys is sensitive to the composition. With Sm doping, the 220 peak shifts minutely towards a higher angle, and the cell volume decreases. When Sm is increased to 3 at. %, the crystal structure changes completely to the 10 M orthorhombic phase with lattice parameters a = 476 Å, b = 5.64 Å and c = 26 Å.
Furthermore, it can also be seen that as the Sm content is increased, the γ peak becomes stronger. Therefore, it can be speculated that the appearance of unknown peaks is due to the presence of Sm, as reported earlier for rear-earth metals [36]. A detailed investigation about the γ phase may be obtained from the temperature-dependent XRD and TEM analyses. Therefore, XRD patterns show that with increasing Sm contents, the structural transformation from austenite to martensite takes place, which is responsible for an increase in the Tt.
The stability of the martensite temperature is a key factor in the application temperature of FMSMAs based on twin boundary migration [34]. Therefore, it is very important to tune the Tt temperature in a wide stable temperature range. The thermal-induced structural transformations in the Ni43Mn46-x SmxSn11 (x = 0, 1, 2, 3) alloys were studied by DSC analysis from 160 to 360 K at a ramp rate of 10 K/min during heating and cooling cycles, as shown in Figure 2a. The existence of large exothermic/endothermic peaks during the heating/cooling process is assigned to the occurrence of reverse martensitic/martensitic transformation in the alloys. These peaks confirm the structural transformation between the austenite and martensite phases.
The first-order nature of the transformation is evident by the presence of thermal hysteresis among the heating and cooling processes [37]. The characteristic MT temperatures are the austenite start temperature As, austenite finish temperature Af, martensite start temperature Ms and martensite finish temperature Mf. The Tt in heating and cooling processes can be calculated using the equations TtH = (As + Af)/2 and TtC = (Ms + Mf)/2. The characteristic temperatures of the MT of the Ni43Mn46-xSmxSn11 (x = 0, 1 and 3) alloys are listed in Table 2. The sample without Sm doping (x = 0) is found far below the RT (As = 200 K, Af = 214 K, Ms = 197 K and Mf = 184 K). With a small amount of Sm doping (x = 1), the Tt increases, with a range of 62 to 69 K, i.e., the As, Af, Ms and Mf are increased by 69, 62, 64 and 65 K, respectively. An increase in the Tt has already been reported in FMSMAs by Gd (rare element substitution) [36,37,38]. With the further increase, the Mt shows an overall increasing tendency.
Many investigations on FMSMA alloys show that the composition of the constituents influences the Tt. The results reveal that the valance electron concentration (e/a) is the most significant factor affecting the martensitic transformation, and that there is a linear relationship between the Tt and e/a [34]. The second important factor that affects the MT is the change in the matrix composition. It is agreed that the substitution with atoms having different sizes from their parent atoms results in the expansion or contraction of the unit cell, and hence lattice distortion. Figure 2b shows that as the e/a is increased, the characteristic temperatures of martensitic transformation increase linearly, which is in accordance with previous reports [32].
The temperature dependence of the magnetization (M–T) curves of the Ni43Mn46-xSmxSn11 (x = 0, 1, 2 and 3) alloys were measured to investigate the magnetic phase transition. M–T curves were measured under the applied magnetic field of 0.01 T, with a temperature range of 150–400 K, during heating and cooling processes, as shown in Figure 3a–d. It can be seen that for x = 0, the alloy sample experiences a magnetic order–disorder transition of martensite at about ≈175 K, followed by reverse martensitic transformation to ferromagnetic austenite at ≈200 K during heating.
With a further increase in the temperature to about ≈310 K, the magnetization gradually decreases to zero, which corresponds to the Curie temperature of austenite. During the cooling process, however, a sudden drop in magnetization can be observed at about ≈197 K, corresponding to the occurrence of MT from the martensite to the austenite phase. Irreversibility during the heating and cooling processes is observed, which is the signature of the first-order transition and is frequently seen in these alloys [16]. It is seen that the Tt increases remarkably with the Sm addition. For x = 1, martensitic transformation occurs between the paramagnetic austenite and ferromagnetic martensite, and the Tt increases by ≈ 65 K, but the Curie temperature of austenite almost remains the same. As shown in Figure 3c,d, the behavior is different for x = 2 and 3: here, the structural transition occurs in the paramagnetic state, and magnetic transition disappears, indicating that the magnetic field cannot induce the transition in x = 2 and 3. The characteristic temperatures measured in the M–T curves agree well with the DSC measurements.
It can be seen from the DSC and M–T measurements that the Ni43Mn45SmSn11 alloy undergoes structural as well as magnetic transition near the Tt. Therefore, isothermal magnetization (M–B) curves under the applied magnetic field of 5 T around MT were measured to investigate the field-induced transition, as shown in Figure 4a. Initially, the sample was cooled down to complete the martensitic state and then slowly heated to the target temperature before starting each M–B measurement, known as the loop method [33]. This alloy exhibited a weak magnetic behavior at temperatures below the Tt (248 and 257 K), while near the Tt (260, 263 and 266 K), it showed a metamagnetic behavior, which indicates that the magnetic field can induce the transition and then transform to a strong magnetic state at 269 K. Hence, it can be concluded from the M–B curves that the magnetic field induced MT between the weak magnetic austenite and strong magnetic martensite state. A magnetization difference (ΔM) of about 38 Am2/kg under a field of 5 T was observed in this alloy. This ΔM is attributed to the field driving capacity [39].
The structural transformation between the austenite and martensite phases in the Ni43Mn45SmSn11 alloy can be driven by the magnetic field, indicated by the magnetization difference. The M–T curves of the Ni43Mn45SmSn11 alloy were measured under a field of 5 T, as shown in Figure 4b. If a comparison is made with M–T at 0.01 T, the Tt decreased by about ≈9 K, and ΔM increased by ≈37 Am2/kg, with the rate of As shifted by the magnetic field, i.e., dAs/dB, being −1.8 K·T−1, confirming the field-induced MT. This magnetic field driving capacity for the Ni43Mn45SmSn11 alloy can be attributed to the larger value of ΔM. Commonly, M–T curves at different magnetic fields can be used to obtain the values of magnetic entropy change (ΔSM). The transformation hysteresis defined as Af–Ms [39] was 18 and 16 K under the applied magnetic fields of 5 and 0.01 T. Here, we calculated the ΔSM of the Ni43Mn45SmSn11 alloy using the Clausius–Clapeyron equation (ΔS = −ΔM·dB/dT) from the M–T curves at 0.01 and 5 T. The obtained value of ΔS was 20.51 J·K−1·kg−1 (using ΔM = 37 Am2/kg, change in magnetic field (ΔB) = 4.99 T and temperature change (ΔT) = 9 K), which is in good agreement with the ΔSM obtained from the M–B curves using Maxwell relations.
The temperature-dependent ΔSM of the Ni43Mn45SmSn11 alloy was calculated by the M–B curves using the Maxwell relation given in Equation (2). The measured ΔSM for the Ni43Mn45SmSn11 alloy was 20 J·kg−1·K−1 at a temperature of 266 K, with a magnetic field variation of 0–5 T, as shown in Figure 5. The ΔSM originates from a magnetic field-induced magnetostructural transition. It can be seen that with 1% doping of Sm, the working temperature of the NiMnSn alloy increased to 266 from ~205 K, showing potential for magnetic refrigeration applications near RT. Moreover, the ΔSM obtained for Ni43Mn45SmSn11 at 266 K is comparable to many other first-order magnetocaloric materials [16,34,39,40,41].
Δ S M = 0 B δ M ( B , T ) δ T dB
The maximum value of the ΔSM derived from Equation (2) is smaller than the ΔST calculated from the DSC measurements because the magnetic field of 5 T is unable to undergo a complete phase transition.
The refrigeration capacity (RC) is a very important factor to assess the MCE and can be calculated by integration of the area under the ΔSM–T curves, using the temperature at half maximum of the peak as the upper and lower integration limits [34]. The calculated RC for the Ni43Mn45SmSn11 alloy under the 5 T field was 262 J·Kg–1. The high MCE and RC values near RT show that the Ni43Mn45SmSn11 alloy can be used in magnetic refrigeration.

4. Conclusions

The effect of Sm doping on martensitic transformation and the magnetic properties of Ni43Mn46-xSmxSn11 ferromagnetic shape memory alloys was investigated. The results show that Sm doping changes the crystal structure from cubic L21 to orthorhombic martensite, with a fraction of the γ phase. The DSC measurements showed endothermic/exothermic peaks, indicating the martensitic and reverse martensitic transformation. The Tt increased significantly with the Sm content and exceeded RT for Sm = 3 at. %. The Ni43Mn45SmSn11 alloy exhibited a magnetostructural transition, and a magnetic field-induced transition was confirmed by the high-field M–T and M–B measurements. Moreover, a magnetic entropy change (ΔSM) value of 20 J·kg−1·K−1 was reached at a temperature of 266 K, and the RC was about 166 J·kg−1, obtained for the Ni43Mn45SmSn11 alloy under an applied magnetic field of 5 T. The large ΔSM value near RT for the Ni43Mn45SmSn11 alloy shows its potential for magnetic refrigeration application.

Author Contributions

Conceptualization, N.u.H. and M.J. (Mosin Jelani); methodology, N.u.H. and I.A.S.; formal analysis, N.u.H. and K.U.R.; investigation, N.u.H., M.J. (Mosin Jelani) and I.A.S.; resources, M.F.K. and A.Q.K.; data curation, K.U.R. and A.Q.K.; writing—original draft preparation, N.u.H.; writing—review and editing, N.u.H., M.J. (Muhamma Jamil), S.R. and M.F.K.; supervision, M.F.K. and D.-k.K.; project administration, D.-k.K.; funding acquisition, M.F.K. and D.-k.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Research Foundation of Korea (NRF), ICT (2020R1G1A1012022). This research was also supported by the MOTIE (Ministry of Trade, Industry and Energy (10080581)) and KSRC (Korea Semiconductor Research Consortium) support program for the development of future semiconductor devices.

Data Availability Statement

All data is provided in the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Planes, A.; Mañosa, L.; Moya, X.; Krenke, T.; Acet, M.; Wassermann, E. Magnetocaloric effect in Heusler shape-memory alloys. J. Magn. Magn. Mater. 2007, 310, 2767–2769. [Google Scholar] [CrossRef]
  2. Khan, M.; Ali, N.; Stadler, S. Inverse magnetocaloric effect in ferromagnetic Ni 50 Mn 37+ x Sb 13− x Heusler alloys. J. Appl. Phys. 2007, 101, 053919. [Google Scholar] [CrossRef] [Green Version]
  3. Giri, S.; Patra, M.; Majumdar, S. Exchange bias effect in alloys and compounds. J. Phys. Condens. Matter 2011, 23, 073201. [Google Scholar] [CrossRef] [PubMed]
  4. Wang, B.M.; Liu, Y.; Xia, B.; Ren, P.; Wang, L. Large exchange bias obtainable through zero-field cooling from an unmagnetized state in Ni-Mn-Sn alloys. J. Appl. Phys. 2012, 111, 43912. [Google Scholar] [CrossRef]
  5. Xuan, H.C.; Cao, Q.Q.; Zhang, C.L.; Ma, S.C.; Chen, S.Y.; Wang, D.H.; Du, Y.W. Large exchange bias field in the Ni–Mn–Sn Heusler alloys with high content of Mn. Appl. Phys. Lett. 2010, 96, 202502. [Google Scholar] [CrossRef]
  6. Chandra, L.S.S.; Chattopadhyay, M.K.; Sharma, V.; Roy, S.B.; Pandey, S. Temperature dependence of thermoelectric power and thermal conductivity in ferromagnetic shape memory alloyNi50Mn34In16in magnetic fields. Phys. Rev. B 2010, 81, 195105. [Google Scholar] [CrossRef]
  7. Zhang, B.; Zhang, X.X.; Yu, S.Y.; Chen, J.L.; Cao, Z.X.; Wu, G.H. Giant magnetothermal conductivity in the Ni–Mn–In ferromagnetic shape memory alloys. Appl. Phys. Lett. 2007, 91, 012510. [Google Scholar] [CrossRef]
  8. Ghosh, A.; Mandal, K. Large magnetoresistance associated with large inverse magnetocaloric effect in Ni-Co-Mn-Sn alloys. Eur. Phys. J. B 2013, 86, 1–6. [Google Scholar] [CrossRef]
  9. Xuan, H.C.; Deng, Y.; Wang, D.H.; Zhang, C.L.; Han, Z.D.; Du, Y.W. Effect of annealing on the martensitic transformation and magnetoresistance in Ni–Mn–Sn ribbons. J. Phys. D Appl. Phys. 2008, 41. [Google Scholar] [CrossRef]
  10. Ito, W.; Ito, K.; Umetsu, R.Y.; Kainuma, R.; Koyama, K.; Watanabe, K.; Fujita, A.; Oikawa, K.; Ishida, K.; Kanomata, T. Kinetic arrest of martensitic transformation in the NiCoMnIn metamagnetic shape memory alloy. Appl. Phys. Lett. 2008, 92, 021908. [Google Scholar] [CrossRef]
  11. Liu, J.; Gong, Y.; Xu, G.; Peng, G.; Shah, I.A.; Ul Hassan, N.; Xu, F. Realization of magnetostructural coupling by modifying structural transitions in MnNiSi-CoNiGe system with a wide Curie-temperature window. Sci. Rep. 2016, 6, 1–8. [Google Scholar]
  12. Karaca, H.; Karaman, I.; Basaran, B.; Ren, Y.; Chumlyakov, Y.I.; Maier, H.J. Magnetic Field-Induced Phase Transformation in NiMnCoIn Magnetic Shape-Memory Alloys-A New Actuation Mechanism with Large Work Output. Adv. Funct. Mater. 2009, 19, 983–998. [Google Scholar] [CrossRef]
  13. Liu, J.; Scheerbaum, N.; Kauffmann-Weiss, S.; Gutfleisch, O. NiMn-Based Alloys and Composites for Magnetically Controlled Dampers and Actuators. Adv. Eng. Mater. 2012, 14, 653–667. [Google Scholar] [CrossRef]
  14. Karaman, I.; Basaran, B.; Karaca, H.E.; Karsilayan, A.I.; Chumlyakov, Y.I. Energy harvesting using martensite variant reorientation mechanism in a NiMnGa magnetic shape memory alloy. Appl. Phys. Lett. 2007, 90, 172505. [Google Scholar] [CrossRef]
  15. Hassan, N.U.; Chen, F.; Zhang, M.; Shah, I.A.; Liu, J.; Gong, Y.; Xu, G.; Xu, F. Realisation of magnetostructural coupling and a large magnetocaloric effect in the MnCoGe 1+x system. J. Magn. Magn. Mater. 2017, 439, 120–125. [Google Scholar] [CrossRef]
  16. Liu, J.; Gong, Y.; Xu, G.; Xu, F. Effect of Sb-doping on martensitic transformation and magnetocaloric effect in Mn-rich Mn50Ni40Sn10-xSbx (x= 1, 2, 3, and 4) alloys. Chin. Phys. B 2017, 26, 448–452. [Google Scholar]
  17. Jani, J.M.; Leary, M.; Subic, A.; Gibson, M.A. A review of shape memory alloy research, applications and opportunities. Mater. Des. 2014, 56, 1078–1113. [Google Scholar] [CrossRef]
  18. Guillou, F.; Porcari, G.; Yibole, H.; Van Dijk, N.; Brück, E. Taming the First-Order Transition in Giant Magnetocaloric Materials. Adv. Mater. 2014, 26, 2671–2675. [Google Scholar] [CrossRef] [Green Version]
  19. Gutfleisch, O.; Willard, M.A.; Brück, E.; Chen, C.H.; Sankar, S.G.; Liu, J.P. Magnetic materials and devices for the 21st century: Stronger, lighter, and more energy efficient. Adv. Mater. 2011, 23, 821–842. [Google Scholar] [CrossRef]
  20. Lyubina, J.; Nenkov, K.; Schultz, L.; Gutfleisch, O. Multiple Metamagnetic Transitions in the Magnetic RefrigerantLa(Fe, Si)13Hx. Phys. Rev. Lett. 2008, 101, 177203. [Google Scholar] [CrossRef]
  21. Zhang, H.; Sun, Y.; Li, Y.; Wu, Y.; Long, Y.; Shen, J.; Hu, F.; Sun, J.; Shen, B. Mechanical properties and magnetocaloric effects in La (Fe, Si) 13 hydrides bonded with different epoxy resins. J. Appl. Phys. 2015, 117, 063902. [Google Scholar] [CrossRef] [Green Version]
  22. Mandal, K.; Pal, D.; Gutfleisch, O.; Kerschl, P.; Müller, K.H. Magnetocaloric effect in reactively-milled LaFe 11.57 Si 1.43 H y intermetallic compounds. J. Appl. Phys. 2007, 102, 053906. [Google Scholar] [CrossRef]
  23. Hu, F.-X.; Shen, B.-G.; Sun, J.-R.; Cheng, Z.-H.; Rao, G.-H.; Zhang, X.-X. Influence of negative lattice expansion and metamagnetic transition on magnetic entropy change in the compound LaFe11.4Si1.6. Appl. Phys. Lett. 2001, 78, 3675–3677. [Google Scholar] [CrossRef]
  24. Pecharsky, V.K.; Gschneidner, K.A., Jr. Giant magnetocaloric effect in Gd5(Si2Ge2). Phys. Rev. Lett. 1997, 78, 4494. [Google Scholar] [CrossRef]
  25. Shah, I.A.; ul Hassan, N.; Riaz, S.; Naseem, S.; Xu, F.; Ullah, Z. Realization of Magnetostructural Transition and Magnetocaloric Properties of Ni–Mn–Mo–Sn Heusler Alloys. J. Supercond. Nov. Magn. 2019, 32, 659–665. [Google Scholar] [CrossRef]
  26. Shah, I.A.; ul Hassan, N.; Rauf, A.; Liu, J.; Gong, Y.; Xu, G.; Xu, F. Influence of Ni/Mn ratio on magnetostructural transformation and magnetocaloric effect in Ni48−xCo2Mn38+xSn12 (x = 0, 1.0, 1.5, 2.0, and 2.5) ferromagnetic shape memory alloys. Chin. Phys. B 2017, 26, 097501. [Google Scholar] [CrossRef]
  27. Hassan, N.U.; Shah, I.A.; Liu, J.; Xu, G.; Gong, Y.; Miao, X.; Xu, F. Magnetostructural Coupling and Giant Magnetocaloric Effect in Off-Stoichiometric MnCoGe Alloys. J. Supercond. Nov. Magn. 2018, 31, 3809–3815. [Google Scholar] [CrossRef]
  28. Krenke, T.; Acet, M.; Wassermann, E.F.; Moya, X.; Mañosa, L.; Planes, A. Martensitic transitions and the nature of ferromagnetism in the austenitic and martensitic states of Ni−Mn−Sn alloys. Phys. Rev. B 2005, 72, 014412. [Google Scholar] [CrossRef] [Green Version]
  29. Wu, R.; Shen, F.; Hu, F.; Wang, J.; Bao, L.; Zhang, L.; Liu, Y.; Zhao, Y.; Liang, F.; Zuo, W.; et al. Critical dependence of magnetostructural coupling and magnetocaloric effect on particle size in Mn-Fe-Ni-Ge compounds. Sci. Rep. 2016, 6, 1–10. [Google Scholar]
  30. Chatterjee, S.; Dutta, P.; Singha, P.; Giri, S.; Banerjee, A.; Majumdar, S. Emergence of compensated ferrimagnetic state in Mn2-xRu1+ xGa (x = 0.2, 0.5) alloys. J. Magn. Magn. Mater. 2021, 532, 167956. [Google Scholar] [CrossRef]
  31. Villa, F.; Nespoli, A.; Passaretti, F.; Villa, E. Microstructural and Thermo-Mechanical Characterization of Cast NiTiCu20 Shape Memory Alloy. Materials 2021, 14, 3770. [Google Scholar] [CrossRef]
  32. Tian, X.; Zhang, K.; Tan, C.; Guo, E. Influence of Doping Tb on the Mechanical Properties and Martensitic Transformation of Ni-Mn-Sn Magnetic Shape Memory Alloys. Crystals 2018, 8, 247. [Google Scholar] [CrossRef] [Green Version]
  33. Caron, L.; Ou, Z.; Nguyen, T.; Thanh, D.C.; Tegus, O.; Brück, E. On the determination of the magnetic entropy change in materials with first-order transitions. J. Magn. Magn. Mater. 2009, 321, 3559–3566. [Google Scholar] [CrossRef]
  34. Hassan, N.U.; Shah, I.A.; Jelani, M.; Naeem, M.; Riaz, S.; Naseem, S.; Xu, F.; Ullah, Z. Effect of Ni-Mn ratio on structural, martensitic and magnetic properties of Ni-Mn-Co-Ti ferromagnetic shape memory alloys. Mater. Res. Express 2018, 5, 086102. [Google Scholar] [CrossRef]
  35. Planes, A.; Mañosa, L.; Acet, M. Magnetocaloric effect and its relation to shape-memory properties in ferromagnetic Heusler alloys. J. Phys. Condens. Matter 2009, 21, 233201. [Google Scholar] [CrossRef] [Green Version]
  36. Tan, C.; Zhang, K.; Tian, X.; Cai, W. Effect of Gd addition on microstructure, martensitic transformation and mechanical properties of Ni50Mn36Sn14 ferromagnetic shape memory alloy. J. Alloy. Compd. 2016, 692, 288–293. [Google Scholar] [CrossRef]
  37. Ju, J.; Liu, H.; Shuai, L.; Liu, Z.; Kang, Y.; Yan, C.; Li, H. Martensite Transformation and Mechanical Properties of Polycrystalline Co-Ni-Al Alloys with Gd Doping. Metals 2018, 8, 848. [Google Scholar] [CrossRef] [Green Version]
  38. Li, K.; Gao, L.; Liang, Y. Martensitic Transformation and Magnetic Properties of Ni-Co-Mn-In-Gd Ferromagnetic Shape Memory Alloys. Mater. Trans. 2018, 59, 224–229. [Google Scholar] [CrossRef]
  39. Hassan, N.U.; Shah, I.A.; Khan, T.; Liu, J.; Gong, Y.; Miao, X.; Xu, F. Magnetostructural transformation and magnetocaloric effect in Mn 48− x V x Ni 42 Sn 10 ferromagnetic shape memory alloys. Chin. Phys. B 2018, 27. [Google Scholar] [CrossRef]
  40. Sharma, J.; Suresh, K. Investigation of multifunctional properties of Mn50Ni40−xCoxSn10 (x = 0–6) Heusler alloys. J. Alloy. Compd. 2015, 620, 329–336. [Google Scholar] [CrossRef]
  41. Xuan, H.; Chen, F.; Han, P.; Wang, D.; Du, Y. Effect of Co addition on the martensitic transformation and magnetocaloric effect of Ni–Mn–Al ferromagnetic shape memory alloys. Intermetallics 2014, 47, 31–35. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of Ni43Mn46-xSmxSn11 (x = 0, 1, 2, 3) alloys at room temperature.
Figure 1. XRD patterns of Ni43Mn46-xSmxSn11 (x = 0, 1, 2, 3) alloys at room temperature.
Crystals 11 01115 g001
Figure 2. (a) DSC curves of Ni43Mn46-xSmxSn11 (x = 0, 1, 2, 3) alloys during heating and cooling processes. (b) The effect of the Sm content on the phase transformation temperatures of Ni43Mn46-xSmxSn11 (x = 0, 1, 2, 3) alloys.
Figure 2. (a) DSC curves of Ni43Mn46-xSmxSn11 (x = 0, 1, 2, 3) alloys during heating and cooling processes. (b) The effect of the Sm content on the phase transformation temperatures of Ni43Mn46-xSmxSn11 (x = 0, 1, 2, 3) alloys.
Crystals 11 01115 g002
Figure 3. (ad) The temperature-dependent magnetization (M–T) curves of Ni43Mn46-xSmxSn11 (x = 0, 1, 2, 3) alloys under an applied magnetic field of 0.01 T during the heating process.
Figure 3. (ad) The temperature-dependent magnetization (M–T) curves of Ni43Mn46-xSmxSn11 (x = 0, 1, 2, 3) alloys under an applied magnetic field of 0.01 T during the heating process.
Crystals 11 01115 g003
Figure 4. (a) Isothermal magnetization (M–B) curves of Ni43Mn45SmSn11 alloy at different temperatures around martensitic transformation. (b) High-field M–T curves for Ni43Mn45SmSn11 alloy with a magnetic field of 5 T during heating (solid symbols) and cooling (open symbols) processes.
Figure 4. (a) Isothermal magnetization (M–B) curves of Ni43Mn45SmSn11 alloy at different temperatures around martensitic transformation. (b) High-field M–T curves for Ni43Mn45SmSn11 alloy with a magnetic field of 5 T during heating (solid symbols) and cooling (open symbols) processes.
Crystals 11 01115 g004
Figure 5. The temperature dependence of ΔSM of Ni43Mn45SmSn11 alloy under an applied magnetic field of 5 T measured upon heating.
Figure 5. The temperature dependence of ΔSM of Ni43Mn45SmSn11 alloy under an applied magnetic field of 5 T measured upon heating.
Crystals 11 01115 g005
Table 1. Atomic weight percentage and electronic concentration of Ni43Mn46-xSmxSn11 (x = 0, 1, 2 and 3) alloys.
Table 1. Atomic weight percentage and electronic concentration of Ni43Mn46-xSmxSn11 (x = 0, 1, 2 and 3) alloys.
SampleAtomic Weight Percentage (%)Electronic Concentration (e/a)
NiMnSnSm
x = 042.9646.0911.05079.643
x = 142.9245.1110.901.0779.713
x = 242.9043.9611.032.1179.772
x = 343.0543.0210.903.0379.794
Table 2. The measured values of characteristic temperatures for Ni43Mn46-xSmxSn11 (x = 0, 1, 2 and 3) alloys.
Table 2. The measured values of characteristic temperatures for Ni43Mn46-xSmxSn11 (x = 0, 1, 2 and 3) alloys.
xAs
(K)
Af
(K)
Ms
(K)
Mf
(K)
Tt (Heating)
(K)
Tt (Cooling)
(K)
0200214197184207190
1269276261247272254
2294305292278299285
3314327311294320302
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hassan, N.u.; Jelani, M.; Shah, I.A.; Rehman, K.U.; Khan, A.Q.; Rehman, S.; Jamil, M.; Kim, D.-k.; Khan, M.F. Tunable Martensitic Transformation and Magnetic Properties of Sm-Doped NiMnSn Ferromagnetic Shape Memory Alloys. Crystals 2021, 11, 1115. https://doi.org/10.3390/cryst11091115

AMA Style

Hassan Nu, Jelani M, Shah IA, Rehman KU, Khan AQ, Rehman S, Jamil M, Kim D-k, Khan MF. Tunable Martensitic Transformation and Magnetic Properties of Sm-Doped NiMnSn Ferromagnetic Shape Memory Alloys. Crystals. 2021; 11(9):1115. https://doi.org/10.3390/cryst11091115

Chicago/Turabian Style

Hassan, Najam ul, Mohsan Jelani, Ishfaq Ahmad Shah, Khalil Ur Rehman, Abdul Qayyum Khan, Shania Rehman, Muhammad Jamil, Deok-kee Kim, and Muhammad Farooq Khan. 2021. "Tunable Martensitic Transformation and Magnetic Properties of Sm-Doped NiMnSn Ferromagnetic Shape Memory Alloys" Crystals 11, no. 9: 1115. https://doi.org/10.3390/cryst11091115

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