Next Article in Journal
Lead-Free Multiferroic Barium-Calcium Zirconate-Titanate & Doped Nickel Ferrite Composites
Next Article in Special Issue
Investigation on Magnetization, Magnetocalory, Magnetoresistance, and Electric Properties of Ni-Mn Based Heusler Alloy
Previous Article in Journal / Special Issue
Investigation on Performance and Kerf Characteristics during Cryogenic-Assisted Suspension-Type Abrasive Water Jet Machining of Acrylonitrile Butadiene Rubber
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Martensitic Transformation and Magnetic Properties of Ni-Mn Quinary Heusler Alloy

by
Sandeep Nambiar S.
1,
Murthy B. R. N.
1,*,
Sathyashankara Sharma
1 and
Prasanna A. A.
2
1
Department of Mechanical and Industrial Engineering, Manipal Institute of Technology, Manipal Academy of Higher Education, Manipal 576104, Karnataka, India
2
Department of Physics, Malnad College of Engineering, Hassan 573202, Karnataka, India
*
Author to whom correspondence should be addressed.
J. Compos. Sci. 2023, 7(1), 1; https://doi.org/10.3390/jcs7010001
Submission received: 20 October 2022 / Revised: 9 November 2022 / Accepted: 30 November 2022 / Published: 20 December 2022
(This article belongs to the Special Issue Advanced Polymeric Composites and Hybrid Materials)

Abstract

:
Ni-Mn-based quinary Heusler alloys have seldom been investigated with respect to their martensitic transformation and mechanical properties for near room temperature transformation. In the current work, we identified and investigated martensitic transformation near room temperature, and the martensitic properties of Ni-Mn-Sn-Fe-In-based quinary Heusler alloys. Alloys prepared in an argon-rich vacuum arc melting furnace. During X-ray diffraction (XRD) analysis, it was identified that the L21 cubic structure austenite phase of the alloy transforms into L10 orthorhombic martensite phase in the case of alloys with greater Fe substitution. The martensitic transformation zone of the alloy is also shifted to the near-room-temperature range of 15–28 °C by changing the stoichiometry of the alloy composition. Magnetic measurements like field heating (FH), field cooling (FC) and zero field cooling (ZFC) indicate the presence of a dual magnetic phase in the alloy, while magnetic susceptibility testing also helped to establish claims regarding the magnetic measurement results.

1. Introduction

With more focus being directed towards the preservation of a greener and cleaner environment, there has been a lot of discussion and focus on existing refrigeration technologies. The existing system works using a traditional process, and makes use of various refrigerants such as chlorofluorocarbons (CFCs) and hydrofluorocarbons (HFCs), which are also referred to as greenhouse gases [1,2,3,4]. These greenhouse gasses tend to cause significant damage to the environment, and thus various researchers have worked extensively on the development of alternate technologies. One alternative to existing refrigeration systems is the use of an energy-efficient refrigeration technique using Heusler alloy, which is a type of ferromagnetic shape memory (FSM) alloy [5]. These alloys have attracted attention due to their excellent functional properties, which include magnetocaloric effect, shape memory effect, enhanced mechanical properties, and inverse magnetocaloric effect. Previous works on such alloys have shown that the annealing temperature, stoichiometric changes, and magnetic field applied play major roles in determining the phase, magnetic and structural properties of Heusler alloys, making them a suitable candidate for various applications [6,7,8,9].
The inferior mechanical properties of Ni-Mn-based Heusler alloys have been the greatest obstacle to their production, thus restricting their applicability [10]. These properties are attributed to their being a highly organized intermetallic with limited dislocation movement tolerance. To overcome this characteristic of the alloy systems, a gamma phase was introduced, strengthening the grain boundary and preventing the development of fractures along the grain boundaries [11]. Among the alloys’ properties, the magnetocaloric effect is of great importance. A change can be observed in the caloric properties when varying the magnetic field applied to it. The majority of Heusler alloys constituted from Ni and Mn have been shown to exhibit inverse transformation properties with applied magnetic field.
Conventional Heusler alloys have been developed for application in the subzero and higher temperature ranges, but very few efforts have been directed towards the development of near-room-temperature martensitic transforming quinary Heusler alloys. Ni-Mn-Co-In-Sn-based alloys have been studied, but mechanical studies and studies on their correlation with magnetic properties have not been performed in detail. Thus, there is a need to achieve near-room-temperature transformation using elements like Ni-Mn-Sn-Fe and In, which undergo a transformation from austenitic phase to a magnetic martensitic phase in a near-room-temperature transformation region [12,13,14]. While our previous work on the same alloy revealed the influence of Fe variation on the alloy’s mechanical properties, this current work helps in understanding the effect of the variation of Fe concentration on martensitic and magnetic properties of the alloy [15]. In this work, the relationship between the structure, microstructure and the magnetic properties of the alloy is established. Little other work has been reported on four-element alloy systems, and these results represent a complete study of the mechanical and magnetic properties of this alloy, thus making the present work of great importance. This paper identifies the constituent elements by considering valance electron ratios, alloy preparation technique, and the heat treatment temperature used, and various compositional, microstructural and magnetic tests are performed to establish the desired results.

2. Materials and Methods

Using the necessary stoichiometry, vacuum arc melting is used to manufacture alloys by melting the 99.9% pure components. Prepare the hearth chamber by removing any signs of oxygen. Initial hearth preparation takes three to five minutes. It was discovered that the electron per atom (e/a) ratio is crucial for influencing the martensitic transition. Due to the high volatility of Mn, 2% of its weight is added extra during the melting process. The electron per atom (e/a) ratio necessary for controlling the martensitic transition is established. Here, the number of valence electrons per atom for Ni, Mn, Sn, Fe, and In are correspondingly 10, 7, 4, 8, and 3. The proportion of atomic weight is written in the form at%. Electron concentrations in the outer shells may be computed using Formula (1). This helps to preserve the composition of the alloy systems at around room temperature and to keep the e/a of the alloy between 7.80 and 7.70, such that it undergoes martensitic transformation near room temperature.
e/a = (10 × (Ni at%)+7 × (Mn at%) + 4 × (Sn at%) + 8 × (Fe at%) + 3 × (In at%))/100
The furnace chamber is evacuated and filled with 200 mm Hg argon gas. Ionized gas heats the sample when the cathode is close enough. The sample is rotated between melting cycles to ensure uniformity and forms a final alloy in the form of a 5 g disc. To reduce mass loss during melting, a phenomenon observed in previous works on Ni-Mn-based alloys owing to evaporation of low-melting-point metals, pure metals were placed in a crucible in decreasing order of their melting temperatures (1538, 1455, 1246, 231.9 and 156.6 °C for Fe, Ni, Mn, Sn and In) [15,16]. This helps melt low-melting-temperature metals by absorbing heat from the higher-melting-temperature metal on top. After melting, the alloy’s (cast as a disc) mass loss was within 1%. This was estimated by a compositional test conducted on the alloy as discussed in the results section. Once the ingots of Ni50−xFexMn30Sn20−yIny where 1 ≤ x ≤ 4; 2 ≤ y ≤ 8 alloy have been obtained, they are subjected to annealing at 800 °C for 48 h in an evacuated vacuum-free setup.
Furthermore, the samples undergo X-ray Diffraction (XRD) to identify the structural transitions that occur in the alloys. A high-resolution X-ray diffractometer (Rigaku Miniflex 600) is used to capture the XRD patterns of the different alloys using filtered CuK light at 0.15400 nm, 600 W CuK with a Ni filter. The system includes a low-noise silicon strip detector, a changeable slit, and a 150 mm goniometer. Continuous scanning is performed over a diffraction angle 2θ ranging from 5 to 80°, which spans the alloy samples diffraction peaks. Using filtered CoK radiation of 0.17900 nm, data are obtained at ambient temperature and at chosen temperatures between 24 and 126 °C to determine the phase and, if existent, polymorphism change. Experiments at high temperatures are performed to determine the lattice expansion and/or phase shift of a martensite-to-austenite alloy. Subsequently, Differential Scanning Calorimetry (DSC) is used to determine the structural transition temperature; a Mettler-Toledo DSC 822e is used [15,16,17,18,19].
A master alloy piece is put in the aluminum chamber for the DSC test, and a reference sample is inserted in the neighboring chamber. Experiments are performed at temperatures between −73 and 200 °C using liquid nitrogen and fine air. A computer interfaced with the apparatus is used to heat and cool the specimen at 10 °C/min. During thermal breakdown, phase shift, or phase transition, a material’s DSC thermogram produces a unique peak. The alloys’ magnetic properties are studied using physical property measurement system (PPMS) and superconducting quantum interference device (SQUID) magnetometers. Quantum Design’s PPMS 14 T is used on a vertical scanning magnetometer (VSM), electrical transport option (ETO), resistivity, heat capacity, and thermal conductivity to assess its magnetic qualities. n. Temperatures are varied from −271 to 126 °C, and the magnetic fields are 14 kOe. The specimens’ readings are taken for zero-field cooled (ZFC), field cooling (FC) and field heating (FH) protocols. Field emission scanning electron microscopy (FESEM) along with energy dispersive X-ray spectroscopy (EDS) are employed to determine the microstructure and composition of the formed alloy. Inductively coupled plasma–optical emission spectrometry (ICP-OES) is employed to correlate the compositional results from EDS [20,21,22].

3. Results and Discussion

3.1. Compositonal Analysis

The compositional analysis of the alloy was performed using two techniques, as previously discussed. The EDX and ICP-OES techniques were used to determine the composition. Approximately 1 g of alloy sample was dissolved fully in 20 mL of a mixture of nitric acid and hydrochloric acid in the ratio of 1:3 (molar), with a representative batch being taken for analysis by gently heating the solution in an 800 mL beaker. Then, 2 mL of HF was added to the solution, which was then diluted further with HCl to a total volume of 100 mL, a 20 mL sample was obtained from this solution and diluted into a 200 mL solution in hydrochloric acid to bring the signal strength to within the measurement range. This solution was fed into the apparatus for the purpose of comparing its absorption spectra to that of a reference solvent. The software interfaced to the machine was used to obtain the spectral intensities in certain spectral lines. This enabled the Ni, Mn, Sn, Fe and In concentrations of the master alloy to be investigated.
The results of the test are shown in Table 1, and Figure 1 depicts the histography of the EDX with respect to the composition of the alloy set. Here, it can be seen that the homogeneity of the alloying elements can be established in all four compositions, with the variation in alloy composition corresopnding to 1 ≤ x ≤ 4 and 2 ≤ y ≤ 8 [19,20].

3.2. Structural Analysis Using XRD

XRD patterns of the alloys measured the range 10° < 2θ < 90° for diffraction angle 2θ are shown in Figure 2, where the variation in the peaks with changes in stoichiometry can be clearly observed. For the alloy where x = 4 and y = 8, six peaks are formed in its L10 martensitic phase.
The XRD analysis reveals that the alloy composition where x = 1 and y = 2 is in austenite phase. When the concentrations of constituent elements added to the alloys are increased, a coexisting phase can be observed at x = 3 and y = 6, while a complete martensite transformation of the alloy occurs at x = 4 and y = 8. The decrease in unit cell volume in the alloy’s concentration has previously been observed by other researchers with Ni-Mn-based constituents [19]. The lattice parameters of the alloys are shown in Table 2.
Upon examining the transformation in the alloy sets in various temperature ranges, it can be observed that the diffractogram depicts a tetragonal L10 martensite crystal structure. To be more precise, the alloy possesses tetragonal martensite phase at temperatures of 15 °C or below. At 40 °C, the presence of a (220) diffraction peak at 2θ = 47.26° confirms the presence of austenite phase together with martensite peak in this alloy. The DSC analysis provides further confirmation of these outcomes [20,21,22,23,24].

3.3. Microstructure

Figure 3 presents FESEM images of well-ordered martensite nanostrips generated from parallel arrays. The occurrence of these microstructure patterns upon directional cooling implies that a liquid phase breaks up and hardens rapidly into several strips. During the martensite (M)-austenite (A) transition, the specific lattice volume decreases, resulting in a compressive shear stress that causes the martensite phase to break up locally into discrete and deep martensite strips. A largely non-diffusional M-A transition occurs with a restoring force to avoid the compositional increase. Therefore, the strips may be easily cooled out of dynamic equilibrium at a critical temperature. Slipping and/or twinning in local structures may result in self-accommodating behavior of the entities spaced apart, allowing strain energy to be relieved in a pure martensite structure. A frictional force develops in the direction opposite to that of the interfacial motion, which causes the strips to divide into finer substrips as the M and A decrease adiabatically. A closer look at the FESEM images indicates that the alloy is made up of clusters of substrips that depict a nano-twinned structure (a mirror-like stacking defect of the closely packed atomic layers) in the martensite L10 phase.
Increasing Fe concentration in the alloy resulted in the creation of a phase in the alloy, as observed in the XRD study. The images clearly show that the strips vanish with decreasing concentrations of Fe [24,25,26,27,28,29].

3.4. Calorimetric Transition

To investigate how martensite strips and local stresses impact the caloric signals associated with magnetic and structural transitions, the heat production of alloys of tiny crystallites was monitored during heating and cooling at temperatures ranging from −75 to 120 °C. The DSC findings for the four alloys (x = 1, 2, 3, and 4; y = 2, 4, 6, and 8) are shown in Figure 4. Exothermic peak formation occurred during the transition from martensite to austenite, at the midpoint, TA = (As + Af)/2, where As, Af, Ms, Mf indicate the austenite start and finish and martensite start and finish temperatures, respectively, as shown in Table 3.
The transitions seem to have occurred above room temperature; one appears to be close to room temperature. The peaks do not dominate, owing to the abrupt change in temperature at the moment of temperature reversal during the measurement [29,30].

3.5. Thermomagnetic Curves

The microstructure of the alloys suggests the creation of structures resembling thin plates. As seen in Figure 5, each of these plates functions as an independent domain. Thus, it is necessary to investigate the relationship between the microstructure and thermodynamics of the domains and their related magnetic properties. Figure 5 presents the magnetization curves obtained under applied magnetic fields of 5 mT and 5 T, in accordance with the three conventional thermal protocols: (a) Zero Field Cooled (ZFC); (b) Field Cooled (FC); and (c) Field Heating (FH). These curves were used to identify the material [31]. Focusing on the curves recorded at 5 mT, it can be seen that all samples have certain characteristics. At 126 °C, the specimens are in the paramagnetic phase and in an austenitic state, and hence the magnetization is almost equivalent to zero, which is similar to the results obtained by other researchers for three-element alloys. The curves reveal that σ = 4 emu/g when the specimens are heated to −173 °C, and a small rise in the curve can be observed at a blocking temperature Tb of −73 °C, where it can be observed that σ = 4.2 emu/g. As the warming is further continued, σ decreases further to around 0.2 emu/g over the temperature range of 0 to 26 °C. As the temperatures reaches close to the martensitic transformation temperature (MT), a peak in magnetization can be observed in the case of the specimen with x = 1. With increasing values of x, a change in trend can be observed, and prominent peaks can be seen in the plots. This corroborates the DSC results previously obtained. The As and Af temperatures are 28 and 40 °C, respectively, where the σ peak can be observed at 3.1 emu/g. Meanwhile, in the case of the FC cycle, M reverts back from A with an enhanced peak of σ = 3.9 emu/g at Ms and Mf values of 27 and 26 °C, respectively. At lower temperature ranges, the ferromagnetic states are reordered near −73 °C, with σ = 6.5 emu/g, and it becomes steady at lower temperatures.
The results for specimens in vast fields show an increase in σ at temperatuers between 26 and 40 °C. There is also a tendency to see specimens with a greater σ value of around 10 emu/g when x is 4. At lower temperatures, the ZFC curve departs from the FC and FH curves to a comparable extent for all specimens, demonstrating magnetic irreversibility. This has been connected to the presence of ferromagnetic particles or grains in a nonmagnetic matrix in Heusler alloys. There are several differences between the ZFC-FC curves acquired under the application of a 5 T magnetic field and those generated under a low magnetic field. There is a continuous increase in magnetization close to the Curie temperature, and just below that, the magnetization undergoes a hysteretic shift at temperatures compatible with the MT that is comparable to that found by DSC analysis [23,24].

3.6. Magnetic Susceptibility

To further understand the dynamics of the alloy’s martensitic transition, AC susceptibility measurements were taken. Measurements were carried out at B = 0.1 mT with a frequency of (f) 20 Hz and at temperatures ranging from −223 to 76 °C, as shown in Figure 6. There is a hysteretic peak at temperatures corresponding to the MT in the samples with x = 1 and 2, which in good agreement with the values obtained from the DSC measurements. Peaks can be observed at −72, −73, −74 and −76 °C.
Larger peaks can also be observed at 22, 24, 27 and 28 °C, with these results being similar to the results obtained for the ZFC-FC curves. It can also be observed that, with the substitution of x = 4, an enhanced curve in the temperature range of 20 °C for 10 kHz. There is an abrupt decrease in χac value with increasing Sn content. When the Sn content is at its lowest at y = 4, there is a huge increase in χac value. This is mainly due to the change in the magnetic anisotropy from cubic to uniaxial in the latter alloy [31,32,33,34,35,36].

4. Conclusions

Ni-Mn-based Heusler alloys formed using the vacuum arc melting technique tend to show martensitic transformation properties. These microstructure patterns suggest that liquid phase hardens in strips upon directed cooling. During the transition, the specific lattice volume decreases, causing a compressive shear stress that breaks apart martensite into deep strips. The addition of Fe up to Fe4 generates nickel martensite, and an FCC phase structure, as seen from the XRD data. The Ni-Fe martensite with an FCC structure exhibits high strength and is robust. This newly generated phase is opposite to the iron-carbon phase. The images show that the alloy is made up of clusters of sub strips with a nano-twinned structure in the martensite L10 phase. The XRD analysis shows that increasing the Fe content in the alloy creates a phase in which these strips disappear, exhibiting decreasing material concentrations. The structure of the alloy transforms from L21 austenitic phase to L10 martensitic phase with the change in the constituent composition of the alloy. The inclusion of Fe can be seen to have helped in obtaining a transformation temperature of around 19–32 °C, as per the results obtained from the DSC tests. These temperature and mechanical properties are appropriate for use in the formation of alloys at larger scale for application. The FC-FH-ZFC curves show the magnetic phase in the alloys in the transformation temperature range. These results also corroborate the results obtained during the AC susceptibility test.

Author Contributions

Conceptualization, S.N.S. and P.A.A.; methodology, S.N.S.; software, S.N.S.; validation, P.A.A., S.S. and M.B.R.N.; formal analysis, S.N.S.; investigation, S.N.S.; resources, S.N.S.; data curation, S.N.S.; writing—original draft preparation, S.N.S.; writing—review and editing, S.N.S.; visualization, S.N.S.; supervision, M.B.R.N. and S.S.; project administration, M.B.R.N. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

INUP: Centre for Nano Science and Engineering (CeNSE), Indian Institute of Science, Bangalore: Karnataka, India.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Everhart, W.; Newkirk, J. Mechanical properties of Heusler alloys. Heliyon 2019, 5, e01578. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Heusler, F. Über Manganbronze und über die Synthese magnetisierbarer Legierungen aus Unmagnetischen Metallen; Angewandte Chemie: Berlin, Germany, 1904; pp. 260–264. [Google Scholar]
  3. Society, R. Available online: https://royalsocietypublishing.org/ (accessed on 13 May 2022).
  4. Graf, T.; Felser, C.; Parkin, S.S. Progress in Solid state chemistry simple rules for the understanding of Heusler compounds. Prog. Solid State Chem. 2011, 39, 1–50. [Google Scholar] [CrossRef]
  5. Zhang, Y.; Billman, J.; Shamberger, P.J. Size effects in the martensitic transformation hysteresis in Ni–Mn–Sn Heusler alloy film. Acta Mater. 2019, 180, 116–125. [Google Scholar] [CrossRef]
  6. Czaja, P.; Przewoźnik, J.; Fitta, M. Heat treatment effect on the evolution of magnetic properties of martensite in magnetic shape memory Ni48Mn39. 5Sn9. 5Al3 Heusler alloy ribbons. Mater. Res. Bull. 2021, 135, 111120. [Google Scholar] [CrossRef]
  7. Liu, Z.H.; Liu, G.D.; Wu, G.H.; Meng, F.B.; Li, Y.X. The influence of heat treatment on the magnetic and phase transformation properties of quaternary Heusler alloy Ni50Mn8Fe17Ga25 ribbons. Scr. Mater. 2004, 51, 1011–1015. [Google Scholar] [CrossRef]
  8. Popa, F.; Chicinas, H.F.; Marinca, T.F.; Chicinas, I. Influence of mechanical alloying and heat treatment processing on the Ni2MnSn Heusler alloy structure. J. Alloys Compd. 2017, 716, 137–143. [Google Scholar] [CrossRef]
  9. Bradley, A.J.; Rodgers, J.W. The crystal structure of the Heusler alloys. Proc. R. Soc. London. Ser. A 1934, 144, 340–359. [Google Scholar]
  10. Liu, H.-X.; Yan, H.-L.; Jia, N.; Tang, S.; Cong, D.; Yang, B.; Li, Z.; Zhang, Y.; Esling, C.; Zhao, X.; et al. Machine-learning-assisted discovery of empirical rule for inherent brittleness of full Heusler alloys. J. Mater. Sci. Technol. 2022, 131, 1–13. [Google Scholar] [CrossRef]
  11. Zhang, Y.; Kumar, S.V.; Xiang, W.; Wu, Z.; Sun, Z. Martensitic transformation and mechanical properties of grain refined Ni–Co–Mn–Sn Heusler alloys via Cr doping. Mater. Sci. Eng. A 2021, 804, 140777. [Google Scholar] [CrossRef]
  12. Notonegoro, H.A.; Ferdian, D.; Prasetyo, Y.; Manaf, A. Microstructure analysis of synthesized iron substituted Ni44Fe6Mn32Al18 Heusler alloy. AIP Conf. Proc. 2016, 1729, 20036. [Google Scholar] [CrossRef]
  13. Yu, B.; Liu, M.; Egolf, P.W.; Kitanovski, A. A review of magnetic refrigerator and heat pump prototypes built before the year. Int. J. Refrig. 2010, 33, 1029–1060. [Google Scholar] [CrossRef]
  14. Dubenko, I.; Ali, N.; Stadler, S.; Zhukov, A.; Zhukova, V.; Hernando, B.; Granovsky, A. Magnetic, magnetocaloric, magnetotransport, and magneto-optical properties of Ni–Mn–In-based Heusler alloys: Bulk, ribbons, and microwires. Nov. Funct. Magn. Mater. 2016, 231, 41–82. [Google Scholar] [CrossRef] [Green Version]
  15. Nambiar, S.S.; Murthy, B.R.N.; Sharma, S.; Prasanna, A.A.; Chelvane, A.J. Microstructure and mechanical properties of annealed quinary Ni-Mn-Sn-Fe-In Heusler alloy. Eng. Sci. 2022, 17, 303–308. [Google Scholar]
  16. Álvares da Silva, G.H.; Otubo, J. Designing NiTiAg shape memory alloys by vacuum arc remelting: First practical insights on melting and casting. Shape Mem. Superelasticity 2018, 4, 402–410. [Google Scholar] [CrossRef]
  17. Ghosh, A.; Mandal, K. Effect of structural disorder on the magnetocaloric properties of Ni-Mn-Sn alloy. Appl. Phys. Lett. 2014, 104, 031905. [Google Scholar] [CrossRef]
  18. Xuan, H.C.; Xie, K.X.; Wang, D.H.; Han, Z.D.; Zhang, C.L.; Gu, B.X.; Du, Y.W. Effect of annealing on the martensitic transformation and magnetocaloric effect in Ni44.1Mn44.2Sn11.7 ribbons. Appl. Phys. Lett. 2008, 92, 242506. [Google Scholar] [CrossRef]
  19. Kumar, S.V.; Singh, R.; Raja, M.M.; Kumar, A.; Bysakh, S.; Mahendran, M. Microstructure and nanomechanical properties of Mn-rich Ni–Mn–Ga thin films. Intermetallics 2016, 71, 57–64. [Google Scholar] [CrossRef]
  20. Peruman, K.V.; Chokkalingam, R.; Mahendran, M. Annealing effect on phase transformation in nano structured Ni–Mn–Ga ferromagnetic shape memory alloy. Phase Transit. 2010, 83, 509–517. [Google Scholar] [CrossRef]
  21. Taubel, A.; Gottschall, T.; Fries, M.; Riegg, S.; Soon, C.; Skokov, K.P.; Gutfleisch, O. A Comparative Study on the Magnetocaloric Properties of Ni-Mn-X(-Co) Heusler Alloys. Phys. Status Solidi (b) 2018, 255, 1700331. [Google Scholar] [CrossRef]
  22. Babita, I.; Raja, M.M.; Gopalan, R.; Chandrasekaran, V.; Ram, S. Phase transformation and magnetic properties in Ni–Mn–Ga Heusler alloys. J. Alloys Compd. 2007, 432, 23–29. [Google Scholar] [CrossRef]
  23. Titov, I.; Acet, M.; Farle, M.; González-Alonso, D.; Mañosa, L.; Planes, A.; Krenke, T. Hysteresis effects in the inverse magnetocaloric effect in martensitic Ni-Mn-In and Ni-Mn-Sn. J. Appl. Phys. 2012, 112, 073914. [Google Scholar] [CrossRef]
  24. Prasanna, A.A.; Ram, S.; Fecht, H.-J. Consecutive magnetic and magnetocaloric transitions in herringbone nanostructured Heusler Mn50Ni41Sn9 alloy. J. Nanosci. Nanotechnol. 2013, 13, 5351–5359. [Google Scholar] [CrossRef] [PubMed]
  25. Li, Z.; Jing, C.; Chen, J.; Yuan, S.; Cao, S.; Zhang, J. Observation of exchange bias in the martensitic state of Ni50Mn36Sn14 Heusler alloy. Appl. Phys. Lett. 2007, 91, 112505. [Google Scholar] [CrossRef]
  26. Czaja, P.; Chulist, R.; Szczerba, M.J.; Przewoźnik, J.; Olejnik, E.; Chrobak, A.; Maziarz, W.; Cesari, E. Magnetostructural transition and magnetocaloric effect in highly textured Ni-Mn-Sn alloy. J. Appl. Phys. 2016, 119, 165102. [Google Scholar] [CrossRef]
  27. Long, Y.; Zhang, Z.Y.; Wen, D.; Wu, G.H.; Ye, R.C.; Chang, Y.Q.; Wan, F.R. Phase transition processes and magnetocaloric effects in the Heusler alloys NiMnGa with concurrence of magnetic and structural phase transition. J. Appl. Phys. 2012, 98, 046102. [Google Scholar] [CrossRef] [Green Version]
  28. Basso, V.; Sasso, C.; Skokov, K.; Gutfleisch, O.; Khovaylo, V.V. Hysteresis and magnetocaloric effect at the magnetostructural phase transition of Ni-Mn-Ga and Ni-Mn-Co-Sn Heusler alloys. Phys. Rev. B 2012, 85, 014430. [Google Scholar] [CrossRef]
  29. Yüzüak, E. The magnetothermal characterization of Ni-Cu-Mn-Sn alloy. Mater. Res. Bull. 2021, 142, 111398. [Google Scholar] [CrossRef]
  30. Guo, J.; Zhong, M.; Zhou, W.; Zhang, Y.; Wu, Z.; Li, Y.; Zhang, J.; Liu, Y.; Yang, H. Grain Size Effect of the γ Phase Precipitation on Martensitic Transformation and Mechanical Properties of Ni–Mn–Sn–Fe Heusler Alloys. Materials 2021, 14, 2339. [Google Scholar] [CrossRef]
  31. Zhao, X.G.; Tong, M.; Shih, C.W.; Li, B.; Chang, W.C.; Liu, W.; Zhang, Z.D. Microstructure, martensitic transitions, magnetocaloric, and exchange bias properties in Fe-doped Ni-Mn-Sn melt-spun ribbons. J. Appl. Phys. 2013, 113, 913. [Google Scholar] [CrossRef]
  32. Stevens, E.; Toman, J.; Kimes, K.; Chernenko, V.; Wojcik, A.; Maziarz, W.; Chmielus, M. Microstructural Evaluation of Magnetocaloric Ni-Co-Mn-Sn Produced by Directed Energy Deposition. Microsc. Microanal. 2016, 22, 1774–1775. [Google Scholar] [CrossRef] [Green Version]
  33. Deltell, A.; Escoda, L.; Saurina, J.; Suñol, J.J. Martensitic Transformation in Ni-Mn-Sn-Co Heusler Alloys. Metals 2015, 5, 695–705. [Google Scholar] [CrossRef] [Green Version]
  34. Moya, X.; Mañosa, L.; Planes, A.; Krenke, T.; Duman, E.Y.Ü.P.; Acet, M.; Wassermann, E.F. Calorimetric study of the inverse magnetocaloric effect in ferromagnetic Ni–Mn–Sn. J. Mag. Mag. Mater. 2007, 316, 572–574. [Google Scholar] [CrossRef]
  35. Kamantsev, A.P.; Koledov, V.V.; Mashirov, A.V.; Dilmieva, E.T.; Shavrov, V.G.; Cwik, J.; Los, A.S.; Nizhankovskii, V.I.; Rogacki, K.; Tereshina, I.S.; et al. Magnetocaloric and thermomagnetic properties of Ni2.18Mn0.82Ga Heusler alloy in high magnetic fields up to 140 kOe. J. Appl. Phys. 2015, 117, 163903. [Google Scholar] [CrossRef] [Green Version]
  36. Fukushima, K.; Sano, K.; Kanomata, T.; Nishihara, H.; Furutani, Y.; Shishido, T.; Ito, W.; Umetsu, R.; Kainuma, R.; Oikawa, K.; et al. Phase diagram of Fe-substituted Ni–Mn–Sn shape memory alloys. Scr. Mater. 2009, 61, 813–816. [Google Scholar] [CrossRef]
Figure 1. EDS histogram representing the composition of the alloys in Ni50−xFexMn30Sn20−yIny where 1 ≤ x ≤ 4; 2 ≤ y ≤ 8 alloys are represented as (ad), respectively.
Figure 1. EDS histogram representing the composition of the alloys in Ni50−xFexMn30Sn20−yIny where 1 ≤ x ≤ 4; 2 ≤ y ≤ 8 alloys are represented as (ad), respectively.
Jcs 07 00001 g001aJcs 07 00001 g001b
Figure 2. XRD representing the composition of the alloys in Ni50−xFexMn30Sn20−yIny where 1 ≤ x ≤ 4; 2 ≤ y ≤ 8 alloys are represented in (ad), respectively.
Figure 2. XRD representing the composition of the alloys in Ni50−xFexMn30Sn20−yIny where 1 ≤ x ≤ 4; 2 ≤ y ≤ 8 alloys are represented in (ad), respectively.
Jcs 07 00001 g002aJcs 07 00001 g002b
Figure 3. FESEM images of specimens Ni50−xFexMn30Sn20−yIny where 1 ≤ x ≤ 4; 2 ≤ y ≤ 8 alloy (a) x = 1 (b) x= 2 (c) x = 3 (d) x = 4.
Figure 3. FESEM images of specimens Ni50−xFexMn30Sn20−yIny where 1 ≤ x ≤ 4; 2 ≤ y ≤ 8 alloy (a) x = 1 (b) x= 2 (c) x = 3 (d) x = 4.
Jcs 07 00001 g003aJcs 07 00001 g003b
Figure 4. DSC plots for the sets of alloys.
Figure 4. DSC plots for the sets of alloys.
Jcs 07 00001 g004
Figure 5. Magnetization vs. temperature curves for samples in magnetic fields of 5 mT and 5 T measured under ZFC (blue), FH (Black) and FC (red) where for Ni50−xFexMn30Sn20−yIny 1 ≤ x ≤ 4; 2 ≤ y ≤ 8 alloy and (a) x = 1; (b) x = 2; (c) x = 3; (d) x = 4.
Figure 5. Magnetization vs. temperature curves for samples in magnetic fields of 5 mT and 5 T measured under ZFC (blue), FH (Black) and FC (red) where for Ni50−xFexMn30Sn20−yIny 1 ≤ x ≤ 4; 2 ≤ y ≤ 8 alloy and (a) x = 1; (b) x = 2; (c) x = 3; (d) x = 4.
Jcs 07 00001 g005aJcs 07 00001 g005b
Figure 6. AC susceptibility vs. temperature curves for Ni50−xFexMn30Sn20−yIny 1 ≤ x ≤ 4; 2 ≤ y ≤ 8 alloy, where (a) x = 1; (b) x = 2; (c) x = 3; (d) x = 4.
Figure 6. AC susceptibility vs. temperature curves for Ni50−xFexMn30Sn20−yIny 1 ≤ x ≤ 4; 2 ≤ y ≤ 8 alloy, where (a) x = 1; (b) x = 2; (c) x = 3; (d) x = 4.
Jcs 07 00001 g006
Table 1. Composition analysis of arc melted alloy using the ICP-OES and EDX techniques.
Table 1. Composition analysis of arc melted alloy using the ICP-OES and EDX techniques.
Alloy ConstituentsMethodComposition (%)
xyNiFeMnSnIn
ICP-OES37.226.5531.543.900.70
12EDX37.426.6331.273.930.77
ICP-OES37.1526.4227.397.091.19
24EDX37.0426.5227.287.041.12
ICP-OES37.1226.8523.6510.531.45
36EDX37.026.9123.5910.681.37
ICP-OES36.9526.9520.9613.711.68
48EDX36.8826.9220.9113.741.75
The values are accurate to within ±0.1% error.
Table 2. Structural parameters estimated for XRD pattern of the alloys.
Table 2. Structural parameters estimated for XRD pattern of the alloys.
xyPhasee/aLattice Parameters
a (nm)c (nm)c/aV (nm3)
12L21(A)7.860.60240.602410.218
24L21(A)7.820.6340.63410.219
36L10(M)7.750.6390.63910.220
48L10(M)7.710.7880.6950.880.431
Table 3. Transformation temperatures as per DSC curve.
Table 3. Transformation temperatures as per DSC curve.
AlloyMs/°CMf/°CAf/°CAs/°CTA/°C
13222.932.622.927.75
23213.532.113.322.7
325.213.22713.120.05
425.31325.31319.15
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Nambiar S., S.; B. R. N., M.; Sharma, S.; A. A., P. Martensitic Transformation and Magnetic Properties of Ni-Mn Quinary Heusler Alloy. J. Compos. Sci. 2023, 7, 1. https://doi.org/10.3390/jcs7010001

AMA Style

Nambiar S. S, B. R. N. M, Sharma S, A. A. P. Martensitic Transformation and Magnetic Properties of Ni-Mn Quinary Heusler Alloy. Journal of Composites Science. 2023; 7(1):1. https://doi.org/10.3390/jcs7010001

Chicago/Turabian Style

Nambiar S., Sandeep, Murthy B. R. N., Sathyashankara Sharma, and Prasanna A. A. 2023. "Martensitic Transformation and Magnetic Properties of Ni-Mn Quinary Heusler Alloy" Journal of Composites Science 7, no. 1: 1. https://doi.org/10.3390/jcs7010001

Article Metrics

Back to TopTop