Next Article in Journal
Highly Efficient Spatial Three-Level CoP@ZIF-8/pNF Based on Modified Porous NF as Dual Functional Electrocatalyst for Water Splitting
Previous Article in Journal
Utilizing Imine Bonds to Create a Self-Gated Mesoporous Silica Material with Controlled Release and Antimicrobial Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Magnetocaloric Effect, Structure, Spinodal Decomposition and Phase Transformations Heusler Alloy Ni-Mn-In

1
Kotelnikov Institute of Radioengineering and Electronics of Russian Academy of Sciences, 125009 Moscow, Russia
2
M.N. Miheev Institute of Metal Physics of Ural Branch of Russian Academy of Sciences, 620108 Ekaterinburg, Russia
3
Department of Physics, The Ural State Agrarian University, 620075 Yekaterinburg, Russia
4
IRC Nanotechnology, Research Park, St. Petersburg State University, 199034 St. Petersburg, Russia
5
Scientific-Practical Materials Research Center of National Academy of Sciences of Belarus, 220072 Minsk, Belarus
6
Department of Mechanical and Aerospace Engineering, Western Michigan University, Kalamazoo, MI 49008, USA
*
Author to whom correspondence should be addressed.
Nanomaterials 2023, 13(8), 1385; https://doi.org/10.3390/nano13081385
Submission received: 18 March 2023 / Revised: 6 April 2023 / Accepted: 9 April 2023 / Published: 16 April 2023

Abstract

:
Ni46Mn41In13 (close to 2-1-1 system) Heusler alloy was studied by magnetization measurement dependence on the temperature in magnetic fields of up to 13.5 T. The magnetocaloric effect measured by the direct method in quasi-adiabatic conditions showed a maximum value of ∆Tad = −4.2 K at a temperature T = 212 K in a magnetic field of 10 T in the region of martensitic transformation. The structure of the alloy was studied by transmission electron microscopy (TEM) as a function of the temperature and the thickness of the sample foil. In the temperature range from 353 to 215 K, at least two processes were established. The results of the study indicate that the concentration stratification occurs according to the mechanism of spinodal decomposition (conditionally spinodal decomposition) into nanoscale regions. At a temperature of 215 K and lower, martensitic phase with 14 M modulation is observed in the alloy at thicknesses greater than 50 nm. Some austenite is also observed. In foils with thickness of less than 50 nm in a temperature range from 353 to 100 Km only the initial austenite, which has not transformed, was found.

1. Introduction

Heusler alloys of various systems, particularly those based on Ni-Mn-In, are promising materials with multifunctional properties for the future design of thermally and magnetically controlled micro- and nanoactuators and solid-state magnetic refrigerators. By varying both the chemical and phase composition, it is possible to change the characteristic temperatures of the metamagnetostructural transformation of thermoelastic martensitic type transformation, which greatly expands the possibilities of application. Therefore, the study of such materials, especially their crystalline structure, is necessary to achieve the maximum values of the beneficial effects and increase the strength and stability of materials during operation [1,2]. Heusler alloys represent a wide range of materials, which are commonly understood as chemical compounds with structural ordering, such as L21, that transitions upon cooling to a low symmetry phase (martensite). So Heusler alloys Ni-Mn-In, both in stoichiometric and non-stoichiometric compositions, often demonstrate a sequence of structural and magnetic phase transformations whose temperatures can be carefully tailored giving rise to attractive multifunctional properties in a useful temperature range [1,2,3].
One of the important properties of considerable interest of researchers in Heusler alloys of the Ni-Mn-X family (X = In, Sn, Sb) is the magnetocaloric effect (MCE), which demonstrates high values in the region of martensitic transformation [4,5]. It is known [6] that the MCE is characterized by a change in temperature ∆Tad under adiabatic conditions (direct measurements) and a change in entropy ∆Smag in isothermal conditions (indirect measurements) at the application of an external magnetic field and can be both conventional and inverse. A certain composition of the Ni-Mn-X (X = In, Sn, Sb) alloys are shown to exhibit an inverse MCE in the vicinity of the structural transformation of martensitic phase to austenitic phase, while the alloys of the system Ni-Mn-Ga have a conventional MCE.
In general, as the temperature decreases the high-temperature austenitic state of the nonstoichiometric Heusler alloys NixMnyInz can undergo a series of disorder–order transitions A2→B2→L21, depending on the occupancy of the positions of the X, Y, and Z sublattices [7]. At temperatures above 1220 K, an A2-type phase is observed, in which the atoms of the X, Y, and Z elements are randomly distributed over the sites of the crystal lattice. As the sample is cooled, the degree of disorder decreases and, upon cooling to 990 K, the alloy transforms into a B2-type phase, in which the atoms of element X occupy positions at the corners of the crystal lattice, and elements Y and Z occupy positions with equal probability at the center of the cubic cell. Upon further cooling, the alloy transforms from the partially ordered state B2 to the ordered phase L21. Any deviation from this stoichiometry leads to decomposition, and the alloys prefer to have an energetically favorable structure with a 2-1-1 stoichiometry. The rest of the alloys can form different structures with different compositions.
The focus of the present study is on the structure of Heusler alloys of non-stoichiometric compositions, since it has not yet been shown reliably whether it is possible to stabilize in them a single-phase, austenitic L21-type ordered phase. The B2→L21 transformation in these compounds requires very long annealing at low temperatures, and diffusion processes proceed more slowly below 600 K; as a result, a certain amount of a disordered phase may be present at room temperature along with a highly ordered phase.
Work on the study of order–disorder transitions in alloys based on Heusler Ni–Mn–In showed that, for example, in Ni45Co5Mn50−xInx, the transformation temperature B2→L21 is approximately 896 K [8,9] and the portion of the L21 phase increases with cooling. This means that the structural order present at room temperature depends on the cooling conditions, in particular on the rate [10].
In addition, non-stoichiometric Heusler alloys with an excess of manganese have structural metastability, which manifests itself in the coexistence of two phases; namely, they transform into a composite made of L21(Ni50Mn25X25) + L10(Ni50Mn50) [11]. The decomposition process occurs during slow cooling or during annealing of ternary Heusler alloys below the B2→L21 transformation temperature. In [12], it is shown that during the decomposition of the Ni50Mn45In5 alloy at 650, 700, and 750 K, inclusions of the L21 phase with Ni50Mn25In25 composition are formed, embedded in a matrix of the L10 phase with Ni50Mn50 composition. These phases are in different magnetic states: the inclusions are paramagnetic, whereas the matrix is antiferromagnetic. The interaction between them leads to a phenomenon called «shell ferromagnetism» [12].
This paper presents the study, by transmission electron microscopy (TEM), of the structural features of Ni46Mn41In13 nonstoichiometric Heusler alloy that affect the magnetostructural transformations in this alloy. The study of the thermoelastic metamagnetostructural transition in both bulk samples and nanoscale foils of this alloy, under high magnetic fields and temperature, is also an aim of the present work.

2. Experimental

Polycrystalline samples of the Ni46Mn41In13 alloy (e/a = 7.86) were arc melted and subsequently vacuum annealed, as described in the [13]. X-ray diffraction studies at room temperature were carried out using Cu K-alpha radiation on a Rigaku Smartlab X-ray diffractometer. The magnetization was measured by the induction method on the cryogenics setup in magnetic fields up to 13.5 T. Measurements of the magnetocaloric effect were made by the direct method in stationary magnetic fields 2, 5, and 10 T on the original setup described in [14]. Electron microscopy studies, including electron diffraction and TEM imaging in bright and dark fields, were carried out with an electron beam accelerating voltage of 200 kV. Temperature in situ experiments were carried out using a dual-axis TEM cryoanalytic holder Gatan model 636 compatible with the Temperature Controller model 900 SmartSet cold stage controller.
Foils for TEM were obtained by mechanical thinning with abrasives of various fineness, and further thinning was carried out electrochemically in an ethanol solution of 30% nitric acid at room temperature. The final thinning was carried out in an ion polishing unit with argon ions. The samples for the structural study were ultrathin TEM foils obtained by standard methods. Electron microscopy studies and foils preparation for TEM were carried out at the IRC Nanotechnology, Research Park, St. Petersburg State University, 199034 St. Petersburg, Russia.

3. Research Results and Discussion

3.1. Characteristic of Metamagnetostructural Transformation and Magnetocaloric Effect

Using DSC (data present in [13]) methods and determining the temperature dependences of the magnetization (the results are present in Figure 1) in low and high fields, the start and finish temperatures of the direct (Ms and Mf) and reverse (As and Af) metamagnetostructural transformation and the Curie temperature (Tc) were determined.
The magnetization data of the sample confirms the presence of a metamagnetostructural phase transformation. Study by magnetization in a field of 0.018 T: Ms = 230 K, Mf = 213 K, As = 226 K, Af = 245 K, Tc = 323 K (Figure 1a).
The data of the isofield dependence of the magnetization show that the sensitivity of the material to the magnetic field was kMs = 8.9 K/T (Figure 1b), that is., an applied external magnetic field of 1 T, shifts the temperature of the metamagnetostructural transformation by 8.9 K toward lower temperatures. Isofield dependence 1–3 T in [13].
Isothermal curves of magnetization in magnetic fields up to ±13.5 T with an increase in temperature from 5 K, 200 K, 273 K, and 319 K are shown in Figure 1c. It can be seen here that, at a temperature of 5 K, in the region of magnetic frustration the martensitic phase behaves as an antiferromagnet, and the value of the magnetization corresponds to the value of the ZFC regime in isofield measurements (Figure 1b). At a temperature of 200 K, an induced magnetic field metamagnetostructural phase transition is observed. When a magnetic field is applied, ferromagnetic austenite begins to nucleate in the parent martensitic phase. With an increase in the magnetic field from 0 to 6.5 T, the entire martensite passes into the austenite phase, and then up to 13.5 T, the process of technical magnetization of ferromagnetic austenite occurs. When the magnetic field is removed at 4.6 T, on the contrary, a martensitic transition begins, when antiferromagnetic martensite appears in ferromagnetic austenite. It is worth noting that with a second increase in the magnetic field in the range from 0 to 4 T, the magnetization loop passes above the initial one. This is due to the fact that at 200 K, the previously induced austenite does not completely returns to the martensite phase. At a temperature of 273 K, ferromagnetic properties appear in the austenite region. At a temperature of 319 K near the Curie temperature, the dependence of the magnetization acquires a paramagnetic character.
Figure 2 shows the dependences of the magnetocaloric effect ∆Tad on temperature in magnetic fields of 2, 5, and 10 T. It can be seen that, in the temperature range below 160 K, a slight conventional magnetocaloric effect is observed in the region of magnetic frustration in the martensite phase. In the temperature range of 160–240 K, in the region of the metamagnetostructural phase transition martensite–austenite, an inverse magnetocaloric effect appears, and in the temperature range of the existence of ferromagnetic austenite, a positive increase in temperature, that is, a conventional magnetocaloric effect. It should be noted that an increase in the applied magnetic field causes a shift in the maximum values of the MCE to the low-temperature martensite phase. The obtained dependence is consistent with the data of direct measurements of the magnetocaloric effect in Heusler alloys of the Ni-Mn-In system [15,16,17,18,19,20,21,22,23]. Namely, in [24], the maximum value of the magnetocaloric effect was obtained equal to ∆Tad = −5.5 K at a temperature T = 219 K in a magnetic field of 8 T, versus ∆Tad = −4.2 K at a temperature T = 212 K in a magnetic field of 10 T obtained in our article. In a magnetic field of 5 T, the authors obtained a value of ∆Tad = −4.8 K at a temperature of T = 224 K, against ∆Tad = −4.1 K at a temperature of T = 222 K obtained by us.
The behavior of the temperature dependence ∆Tad(T) in the Ni46Mn41In13 compound under study in the region of the magnetic field-induced phase transition from antiferromagnetic martensite to ferromagnetic austenite and the presence of a large inverse MCE can be due to a significant difference in the values of magnetization M(T) in the martensitic phase compared to the austenitic phase. In turn, the difference in magnetization can be explained by the fact that in Mn-based Heusler alloys, magnetic moments are localized mainly on Mn atoms and the exchange interaction strongly depends on the Mn–Mn distance [25,26]. Therefore, any change in the configuration of the crystal lattice during the martensitic transition can change the strength of exchange interactions and, as a result, lead to a change in the magnetic properties of the alloy [27].

3.2. Structural Studies

According to the isofield dependences of the magnetization on temperature, it was found that the metamagnetostructural phase transition in this Ni46Mn41In13 bulk sample has a high sensitivity of characteristic temperatures to the magnetic field KMs = 8.9 K/T in fields of 1–3 T, as mentioned above, which is several times higher than the sensitivity for Heusler alloys of the Ni-Mn-Sb system and for magnetostructural phase transitions in the Ni-Mn-Ga systems. Therefore, it is expected that when the sample size is reduced to the size of thin films obtained from bulk samples, their sensitivity to an external magnetic field will not change (or even increase), and macro deformation, as a result of magnetically and thermally induced martensitic transition, will be large, which is extremely important for creating micro- and nanomanipulators, devices, and other products whose activation requires medium-power magnetic fields. (0.5–3 T, for example), and the value of the magnetocaloric effect will practically not change.
It is known that the thin-film state, in comparison with the bulk state, expands the temperature range of the austenitic phase for Ni-Mn-In systems [13]. Similar deviations in the behavior of the Ni-Mn-Ga system in the thin-film and bulk states were also observed for the Ni-Mn-Ga system [28,29]. However, information about the specific microstructures of various states of such systems (primarily with thermoelastic martensitic and metamagnetostructural transitions) is rather scarce, and this information is extremely necessary, both for applied purposes and for modeling functional materials.
Figure 3 shows X-ray diffraction data for a Ni46Mn41In13 sample at room temperature (300 K), that is, above the martensitic transformation temperature of the alloy. It can be seen that all the main Bragg lines are «blurred», which indicates the non-single-phase nature and the presence of internal stresses. The peak in the diffraction pattern (111) is almost invisible, which indicates incomplete ordering. Characteristic diffraction lines (200) and (220) are both present, demonstrating the formation of B2-type ordering; a (331) peak associated with the emergence of the L21 phase is also observed.
In order to unambiguously determine the structure in nonstoichiometric Heusler alloys from the point of view of «order–disorder», one can consider a stoichiometric (as simpler) compound in B2 and L21 ordering types. The results of such a study are published in the article [30] for the Ni2MnAl alloy. The authors assume that, since the B2→L21 transformation in Ni2MnAl occurs at a temperature at which diffusion proceeds quite slowly, the high-temperature B2 phase is retained at room temperature in a metastable state. This, in turn, prevents the exclusively formation of the thermodynamically more favorable phase L21.
In addition, upon cooling, order–disorder transformations can overlap, and the alloy will tend to decompose and, as was shown in [11], transform into a two-phase composite L21(Ni50Mn25X25) + L10(Ni50Mn50). Determination of the effects of ordering and secondary nanoscale phases is very difficult from X-rays. In addition, at room temperature, there may be a small amount of the martensite phase, and some peaks can be attributed to the martensitic phase with low symmetry. Thus, X-ray diffraction data alone cannot unambiguously determine a small degree of disorder in an ordered structure or a low degree of order in a disordered structure [31]; therefore, a detailed analysis of structural studies by other methods is required.
In connection with the foregoing analysis, we studied in detail the Ni46Mn41In13 alloy by transmission electron microscopy. The micrographs of the alloy in the austenitic state (Figure 4a) show both regions of the completely ordered L21 phase (Figure 4b) and regions characterized by a tweed structure with diffuse scattering and additional reflections on the electron diffraction patterns (Figure 4c).
The electron diffraction patterns of the alloy in the austenitic state (Figure 5) showed consistently characteristic reflections of the B2 structure with a = 3.085 Å and superstructural reflections of the (111) plane, confirming the presence of the L21 superstructure with a = 6.170 Å. In the microdiffraction pattern (Figure 5b) with [110]L21/B2 zone axis, both the main reflections B2 + L21 with high intensity and the superstructural reflections B2 and L21 with low intensity (the grating is highlighted in blue, indexes are for L21) are present. Separately, it should be noted that not a single electron diffraction pattern revealed solely a B2 structure, so it may be concluded that the sample always contained ordered regions. In addition, a characteristic feature of this sample is that in all diffraction patterns of austenite, there are additional reflections localized by three, creating a grating (highlighted in red), one of the directions of which is parallel to [110]L21.
On the bright and dark field electron microscopy images of the austenite structure, a tweed contrast is observed with a distance between the stripes of ~10 nm (Figure 5a), and on the electron diffraction patterns near the main reflections of the L21 phase (except for the superstructural (111)L21 and (001)B2) regions diffuse scattering in the form of two radial diffuse strands intersecting at an angle of ~115° along the directions [2 24 ¯ ]* and [2 2 ¯ 4]*. The tweed contrast bands are in accordance with the diffuse scattering bands: each system of contrast bands is orthogonal to one of the diffuse scattering directions. In addition to the tweed contrast, parallel very thin bands along <220>* with a distance of ~1.35 nm between them are observed (Figure 5a). This indicates that the grating contains micromodulations with a period of nine interplanar distances in the <004>* direction (d004 ≈ 0.15 nm).
The tweed contrast is observed in alloys and compounds in which the martensitic-type transformation is accompanied by pre-transition phenomena: the displacements of atoms towards the future phase. Displacements of atoms from equilibrium positions cause lattice deformations, which manifest themselves in the appearance of modulated structures, in particular, a tweed microstructure. Such tweed domains appearing in the pre-martensitic state probably serve as sites for martensite nucleation. In such a case, satellite reflections and characteristic diffuse scattering appear in the diffraction patterns. The tweed structure observed in this case is the result of the incorporation of the intermediate phase into the initial cubic phase. Such a structure in [32] is called not just pre-martensitic tweed, but an intermediate tweed phase, believing that the so-called pre-martensitic phase is actually an independent phase preceding the martensitic transformation.
At moderate cooling rates, nonstoichiometric Ni46Mn41In13 can decompose into regions that differ both in chemical composition and levels of ordering, resulting in the formation of a modulated periodic structure. Such a structure develops as a result of self-organization of phases under the action of deformations, that is, displacements of atoms from equilibrium positions towards the future phase. The results obtained suggest that the phase transformation begins with the formation of local regions of L21 oriented along directions that create a tweed contrast (orthogonal to the directions of diffuse strands along [2 24 ¯ ]* and [2 2 ¯ 4]*). Figure 6a clearly shows local moiré patterns (circled) arising from fully or partially coherent regions of the ordered phase ~5 nm in size. In some areas of the foil, almost square L21 structural nanodomains ~3–5 nm in size, aligned along the directions of the tweed, are observed (Figure 6b).
Dark-field images obtained from the region in Figure 6a in the main lattice reflections of the L21 phase, which have diffused scattering strands, correspond to one (Figure 6d) or two (Figure 6e) sets of tweed contrast bands. The dark-field image in the triple reflection, which we associate with decay products, is a classical modulated structure resulting from stratification by the spinodal mechanism [33] (Figure 6f). The triplet of reflections in electron diffraction patterns (Figure 4c and Figure 5b) is made of the main reflection and two extra reflections, the origin of which is associated with “punctures” of the Ewald sphere by diffuse strands that do not belong to the given section of the reciprocal lattice. Intensity bands located obliquely to the diffraction plane are depicted on the electron diffraction patterns as satellites near the Bragg reflections.
High resolution TEM analysis revealed the microstructure of the tweed with alternating dark and light domains of 3–5 nm (Figure 7). Striped contrast extending across the entire field of Figure 7b, is formed by the projections of the {220}L21 planes with the interplanar spacing. These bands run continuously through all domains, indicating that the boundaries between domains are coherent.
It should be recalled that the structure of non-stoichiometric compositions of Heusler alloys is metastable, even small inhomogeneities in composition or temperature changes can lead to the formation of various microstructures. For example, depending on the degree of atomic ordering, the Ni46Mn41In13 compound is expected to contain structural domains and antiphase boundaries (APBs) formed during the transition from the partially ordered B2 phase to the ordered L21. In the next part of the work, we will analyze dark-field images obtained of both superstructural and basic reflections of the L21 phase (Figure 8 and Figure 9). In the dark-field image (Figure 8b) obtained of the (1 1 ¯ 1)L21 superstructural reflection, bright regions correspond to zones with a large volume fraction of the L21 phase, whereas dark regions correspond to B2-type ordered zones. It must be considered that some areas of L21 and B2 may intersect at different heights and give an image with different gray intensities depending on the ratio L21/B2 [34]. Since the ordering process in this compound is accompanied by decomposition, this explains why pronounced APBs are not observed in the superstructural reflection of the ordered phase, which will be less energetically favorable compared to the separation-matrix interfaces.
If a dark-field image is obtained in the (004)L21 reflection (Figure 8c), that is, in a reflection with diffuse scattering strands, then a contrast is observed from the APB separating the tweed domains. The dark-field image obtained in the triplet of reflections is shown in Figure 8d. The bright areas in the micrograph are periodic, evenly distributed curved strips in the form of “worms” 100–200 nm long. One can define this structure as a visualization of alternating domains of a partially ordered B2 phase and an ordered L21 phase. Figure 9 shows dark-field APB images obtained in various reflections from a region characterized by a tweed structure (the region is circled in blue in Figure 9a).
Evidence that antiphase boundaries are magnetic domain wall stoppers was presented for the Ni50Mn25Al12.5Ga12.5 alloy (L21 phase) in [35]. It was shown that the antiphase boundaries observed by the dark field method coincided almost perfectly with the magnetic domain walls detected using Lorentz microscopy. Zuo et al. [36] demonstrated periodic stripe magnetic domains in Ni50Mn35In15 observed using in situ Lorentz transmission electron microscopy, which are indeed very similar to the domains in Figure 8d.
© structure in the form of “worms” is similar to the classical modulated structure that occurs during concentration separation, resulting from stress fields, indicating local lattice distortions [37]. Moreover, the X-ray diffraction pattern of the sample (Figure 3) suggests the possibility of the presence of two phases of different composition, and structural metastability, as already mentioned, is an inherent property of Heusler alloys. The foregoing makes it possible to consider the coexistence of order–disorder in these alloys in unity with structural instability.
Next, for the alloy under study, the distribution of the elements Mn, Ni, and In was obtained, and the local concentration of these elements along the scanning line was determined. Scanning from the edge into the depth of the sample with a step of ~300 nm was performed as shown in Figure 10a. The distribution of the elements (see inset), with the exception of the very edge of the foil, did not reveal fluctuations in the composition throughout the entire scanning range. Scanning with a step of ~150 nm, commensurate with the size of the worm-shaped strips (in the inset at the bottom of Figure 10c), revealed a slight fluctuation in the composition of both Ni, Mn, and In (in the inset at the top of Figure 10b). This indicates that the transition to a two-phase composite in this alloy does not occur according to the scenario proposed in [11]. However, this does not rule out the presence of nanoscale regions resulting from separation by the spinodal mechanism into regions with different contents of indium, manganese, or nickel. The alloy Ni46Mn41In13 exhibits a phase decomposition tendency, even without additional low temperature annealing, only in the process of cooling with a furnace provided a suitable system for studying the early stage of decomposition. Microstructural observations indicate that the decay mechanism is indeed spinodal rather than nucleation and growth. The TEM method observes structures that correspond to nanoscale decay uniformly distributed inside the grain. Moreover, since the decay is considered not only as a chemical stratification, but also from the point of view of ordering, in this case we can speak more about conditional spinodal decomposition, rather than about classical spinodal decomposition. In addition, the diversity of structures indicates a high defectiveness, which hinders the formation of a full-fledged phase transformation of the martensitic type, which is typical not only for Heusler alloys [38].

3.3. Austenite–Martensite Boundary

The features of the fine structure of the wedge-shaped foil of the Ni46Mn41In13 alloy observed in this work showed that the martensitic transformation is blocked at a distance of about 600 nm from the edge of the sample at a plate thickness of less than 50 nm, where the formation of a martensitic structure does not occur even at liquid nitrogen temperature [13]. This size effect, shown in Figure 11a provided us with the opportunity to observe the austenite–martensite interface in the Ni46Mn41In13 alloy (Figure 11b,c). The boundary is not sharp and is a fairly wide transition layer ~150 nm wide. Such blurring indicates a high level of stresses, especially on the austenite side (Figure 11d). It can be seen how the banded contrast changes going from austenite (upper right corner) to martensite (lower left corner) through the transition layer. Of the two directions of the thin banded austenite contrast, only one direction remains with the same periodicity cane be seen at first, and then the distance between the bands increases, and a coarse black-and-white banded contrast is observed on the martensite side.
Electron diffraction patterns belonging to the austenite region, the martensite region, and directly to the interface were obtained (indicated in Figure 12a). Figure 12b brings together two sets of reflections; reflections from the electron diffraction pattern of austenite with zone axis [110]L21 are superimposed on the electron diffraction pattern obtained from the martensite region. It can be seen from the combined austenite–martensite diffraction patterns that, as the transformation progresses, all reflections of the L21 phase (reflections of the (220), (002), (004), and (224) types), except for the superstructural ones (111), split in the direction of one of the two diffuse scattering strands near reflections of the (004), (224), and (444) types. In this case, the corresponding planes of the L21 phase transform into the planes of martensite. The appearance of extra reflections on the electron diffraction patterns, at distances that are 1/7 of the distance between the main structural reflections, characterizes the martensitic phase as a modulated 14 M, and the martensitic transformation occurs according to the L21 → 14 M type, which is consistent with the findings of other authors [39,40,41]. Banding in bright-field images of martensite is perpendicular to the direction of the splitting of reflections in the electron diffraction patterns. For reflections belonging to the austenite phase, ordered according to type B2 (in Figure 12b, these reflections are of the (110) and (020) types with two satellites circled in red). When approaching the austenite–martensite interface, one of the three reflections first disappears, then they all vanish. This suggests that these reflections still belong to a phase with a different type of ordering, and not to secondary phases or oxides of manganese and nickel.
Figure 13a shows the electron diffraction pattern of the austenite structure with zone axis [001]L21 and the electron diffraction pattern of the martensite structure, which has a modulated lattice with a periodicity of seven planes along the [ 11 ¯ 0]L21 direction (Figure 13b), corresponding to 14 M. In this case, the martensite structure is in the form of plates with an internal streaky structure (Figure 13c). Banding in dark-field images of martensite is perpendicular to the direction of splitting of reflections in electron diffraction patterns.
The appearance of such a variety of structures obviously suggests that the crystal structure of the martensite phase will strongly depend on the degree of ordering of the austenite phase. In addition, the coexistence of regions L21 and B2 in austenite apparently leads to the impossibility of complete martensitic transformation in the entire volume of the material, since the propagation of martensitic transformation is hindered in a medium inhomogeneity in the microstructure.

4. Summary

The Ni46Mn41In13 Heusler alloy was studied by the measurement of the dependence of magnetization on temperature in magnetic fields up to 13.5 T. The dependence of magnetization showed a shift of the characteristic temperatures in magnetic field of KMs = 8.9 K/T. With a further increase in the magnetic field, the hysteresis expands, and in a field of 13.5 T, blocking of the transformation is observed, and only the austenite (strongly magnetic phase) is observed in the structure. The magnetocaloric effect, measured by the direct method in the temperature range of 140–240 K and magnetic fields of 2, 5, and 10 T, demonstrated the maximum value in a magnetic field of 10 T of the inverse effect −4.2 K in the region of martensitic transformation.
The structure of the non-stoichiometric Ni46Mn41In13 Heusler alloy was studied by TEM methods in the temperature range of 353–215 K. At least two processes were established; one is where the cubic parent phase undergoes shear transformation in combination with diffusion, creating a tweed contrast consisting of a coherent mixture of the parent composition and L21 ordered domains. The second process is associated with stratification into two phases that differ in composition and ordering. Both processes are a single ordering and decay phenomenon in which two phase transformations occur simultaneously, namely, atomic ordering and solid solution decomposition. It is not determined which of these transformations is the leading one. Both processes coexist and occur under the influence of each other. This makes it possible to consider the coexistence of order–disorder in these alloys in unity with structural instability and martensitic transformation.
A qualitative analogy was observed between the effects of an increase in the external magnetic field and a decrease in the thickness of the alloy sample on the manifestation of the martensitic transition in the alloy. Both factors lead to the same qualitative result: a decrease in temperature and subsequent complete blocking of the martensitic phase transition.
The results obtained are considered a single phenomenon in which two processes occur simultaneously, namely, atomic ordering and solid solution decomposition. As a result, a structure is formed consisting of ordered domains separated by decay particles. Such a process occurs, for example, in some alloys of the gold–copper–silver system [42] and is a complex continuous “ordering-decomposition” reaction. Both processes coexist and occur under the influence of each other. This makes it possible to consider the coexistence of order–disorder in these alloys in unity with structural instability and martensitic transformation. The results of this study indicate that the concentration stratification occurs by the mechanism of spinodal decomposition (conditionally spinodal decomposition) into nanoscale regions.

Author Contributions

D.D.K. writing—original draft preparation, investigation. E.I.K. writing—review and editing, investigation. A.V.M. investigation. A.S.L. investigation. D.V.D. investigation. V.I.M. investigation. A.S.K. investigation. V.G.S. investigation. V.V.K. investigation. P.A.-G. investigation. All authors have read and agreed to the published version of the manuscript.

Funding

The study was supported by the Russian Science Foundation (project no. 20-79-10197). Link to information about the project: https://rscf.ru/en/project/20-79-10197/, accessed on 8 April 2023.

Data Availability Statement

Not applicable.

Acknowledgments

The authors are grateful to Vladimir Zolotorev for the sample preparation. Thanks are expressed to Andrei Alexandrovich Lomov for X-ray diffraction, Valiev Institute of Physics and Technology RAS. Electron microscopy studies, including electron diffraction and TEM imaging in bright and dark fields, were carried out with an electron beam accelerating voltage of 200 kV. The samples for the structural study were ultrathin TEM foils obtained by standard methods. Electron microscopy studies and foils preparation for TEM were carried out at the IRC Nanotechnology, Research Park, St. Petersburg State University, 199034 St. Petersburg, Russia. The study was supported by the Russian Science Foundation (project no. 20-79-10197). Link to information about the project: https://rscf.ru/en/project/20-79-10197/, accessed on 8 April 2023.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Elphick, K.; Frost, W.; Samiepour, M.; Kubota, T.; Takanashi, K.; Sukegawa, H.; Mitani, S.; Hirohata, A. Heusler alloys for spintronic devices: Review on recent development and future perspectives. Sci. Technol. Adv. Mater. 2021, 22, 235–271. [Google Scholar] [CrossRef] [PubMed]
  2. Mashirov, A.V.; Irzhak, A.V.; Tabachkova, N.Y.; Milovich, F.O.; Kamantsev, A.P.; Zhao, D.; Liu, J.; Kolesnikova, V.G.; Koledov, V.V. Magnetostructural Phase Transition in Micro-and Nanosize Ni–Mn–Ga–Cu Alloys. IEEE Magn. Lett. 2019, 10, 1–4. [Google Scholar] [CrossRef]
  3. Mashirov, A.V.; Kamantsev, A.P.; Kuznetsov, D.D.; Koledov, V.V.; Shavrov, V.G. Surface Energy upon the Martensitic Phase Transition in Microsized Samples of Ni–Mn–Ga–Cu Heusler Alloy. Bull. Russ. Acad. Sci. Phys. 2021, 85, 751–754. [Google Scholar] [CrossRef]
  4. Bhobe, P.A.; Priolkar, K.R.; Nigam, A.K. Room temperature magnetocaloric effect in Ni–Mn–In. Appl. Phys. Lett. 2007, 91, 242503. [Google Scholar] [CrossRef]
  5. Gottschall, T.; Skokov, K.P.; Frincu, B.; Gutfleisch, O. Large reversible magnetocaloric effect in Ni-Mn-In-Co. Appl. Phys. Lett. 2015, 106, 021901. [Google Scholar] [CrossRef]
  6. Tishin, A.M.; Spichkin, Y.I. The Magnetocaloric Effect and Its Applications; CRC Press: Boca Raton, FL, USA, 2016; p. 486. [Google Scholar]
  7. Wang, Y.; Salas, D.; Duong, T.C.; Medasani, B.; Talapatra, A.; Ren, Y.; Chumlyakov, Y.I.; Karaman, I.; Arróyave, R. On the fast kinetics of B2–L21 ordering in Ni-Co-Mn-In metamagnetic shape memory alloys. J. Alloys Compd. 2019, 781, 479–489. [Google Scholar] [CrossRef]
  8. Ito, W.; Nagasako, M.; Umetsu, R.Y.; Kainuma, R.; Kanomata, T.; Ishida, K. Atomic ordering and magnetic properties in the Ni45Co5Mn36.7In13.3 metamagnetic shape memory alloy. Appl. Phys. Lett. 2008, 93, 232503. [Google Scholar] [CrossRef]
  9. Ito, W.; Imano, Y.; Kainuma, R.; Sutou, Y.; Oikawa, K.; Ishida, K. Martensitic and magnetic transformation behaviors in Heusler-type NiMnIn and NiCoMnIn metamagnetic shape memory alloys. Metall. Mater. Trans. A 2007, 38, 759–766. [Google Scholar] [CrossRef]
  10. Sánchez-Alarcos, V.; Recarte, V.; Pérez-Landazábal, J.I.; Cuello, G.J. Correlation between atomic order and the characteristics of the structural and magnetic transformations in Ni–Mn–Ga shape memory alloys. Acta Mater. 2007, 55, 3883–3889. [Google Scholar] [CrossRef]
  11. Entel, P.; Gruner, M.E.; Acet, M.; Hucht, A.; Çakır, A.; Arróyave, R.; Karaman, I.; Duong, T.C.; Talapatra, A.; Bruno, N.M.; et al. Probing glassiness in Heuslers via density functional theory calculations. Frustrated Mater. Ferroic Glas. 2018, 275, 153–182. [Google Scholar]
  12. Çakır, A.; Mehmet, A.; Michael, F. Shell-ferromagnetism of nano-Heuslers generated by segregation under magnetic field. Sci. Rep. 2016, 6, 28931. [Google Scholar] [CrossRef] [PubMed]
  13. Kuznetsov, D.D.; Kuznetsova, E.I.; Mashirov, A.V.; Loshachenko, A.S.; Danilov, D.V.; Shandryuk, G.A.; Shavrov, V.G.; Koledov, V.V. In Situ TEM Study of Phase Transformations in Nonstoichiometric Ni46Mn41In13 Heusler Alloy. Phys. Solid State 2022, 64, 15–21. [Google Scholar] [CrossRef]
  14. Kuznetsov, A.S.; Mashirov, A.V.; Aliev, A.M.; Petrov, A.O.; Anikin, M.S.; Musabirov, I.I.; Amirov, A.A.; Kon, I.A.; Koledov, V.V.; Shavrov, V.G. The Magnetocaloric Effect upon Adiabatic Demagnetization of a Polycrystalline DyNi2 Alloy. Phys. Met. Metallogr. 2022, 123, 397–401. [Google Scholar] [CrossRef]
  15. Aliev, A.M.; Batdalov, A.B.; Khanov, L.N.; Mashirov, A.V.; Dil’mieva, E.T.; Koledov, V.V.; Shavrov, V.G.E. Degradation of the magnetocaloric effect in Ni49.3Mn40.4In10.3 in a cyclic magnetic field. Phys. Solid State 2020, 62, 837–840. [Google Scholar] [CrossRef]
  16. Khanov, L.N.; Batdalov, A.B.; Mashirov, A.V.; Kamantsev, A.P.; Aliev, A.M. Magnetocaloric Effect and Magnetostriction in a Ni49.3Mn40.4In10.3 Heusler Alloy in AC Magnetic Fields. Phys. Solid State 2018, 60, 1111–1114. [Google Scholar] [CrossRef]
  17. Pramanick, S.; Chatterjee, S.; Giri, S.; Majumdar, S.; Koledov, V.; Mashirov, A.; Aliev, A.; Batdalov, A.; Hernando, B.; Rosa, W.; et al. Multiple magneto-functional properties of Ni46Mn41In13 shape memory alloy. J. Alloys Compd. 2013, 578, 157–161. [Google Scholar] [CrossRef]
  18. Fayzullin, R.; Buchelnikov, V.; Mashirov, A.; Zhukov, M. Phase transformations and magnetocaloric effect in Ni-Mn-(Co)-In Heusler alloys. Phys. Procedia 2015, 75, 1259–1264. [Google Scholar] [CrossRef]
  19. Fayzullin, R.R.; Mashirov, A.V.; Buchelnikov, V.D.; Koledov, V.V.; Shavrov, V.G.; Taskaev, S.V.; Zhukov, M.V. Direct and inverse magnetocaloric effect in Ni1.81Mn1.64In0.55, Ni1.73Mn1.80In0.47, and Ni1.72Mn1.51In0.49 Co0.28 Heusler alloys. J. Commun. Technol. Electron. 2016, 61, 1129–1139. [Google Scholar] [CrossRef]
  20. Gottschall, T.; Skokov, K.P.; Fries, M.; Taubel, A.; Radulov, I.; Scheibel, F.; Benke, D.; Riegg, S.; Gutfleisch, O. Making a cool choice: The materials library of magnetic refrigeration. Adv. Energy Mater. 2019, 9, 1901322. [Google Scholar] [CrossRef]
  21. Ghorbani Zavareh, M.; Salazar Mejía, C.; Nayak, A.K.; Skourski, Y.; Wosnitza, J.; Felser, C.; Nicklas, M. Direct measurements of the magnetocaloric effect in pulsed magnetic fields: The example of the Heusler alloy Ni50Mn35In15. Appl. Phys. Lett. 2015, 106, 071904. [Google Scholar] [CrossRef]
  22. Liu, J.; Gottschall, T.; Skokov, K.P.; Moore, J.D.; Gutfleisch, O. Giant magnetocaloric effect driven by structural transitions. Nat. Mater. 2012, 11, 620–626. [Google Scholar] [CrossRef] [PubMed]
  23. Gottschall, T.; Skokov, K.P.; Scheibel, F.; Acet, M.; Zavareh, M.G.; Skourski, Y.; Wosnitza, J.; Farle, M.; Gutfleisch, O. Dynamical effects of the martensitic transition in magnetocaloric heusler alloys from direct Δ T ad measurements under different magnetic-field-sweep rates. Phys. Rev. Appl. 2016, 5, 024013. [Google Scholar] [CrossRef]
  24. Batdalov, A.B.; Khanov, L.N.; Mashirov, A.V.; Koledov, V.V.; Aliev, A.M. On the nature of the magnetocaloric effect in the Ni46Mn41In13 Heusler alloy in cyclic magnetic fields. J. Appl. Phys. 2021, 129, 123901. [Google Scholar] [CrossRef]
  25. Brown, P.J.; Gandy, A.P.; Ishida, K.; Kainuma, R.; Kanomata, T.; Neumann, K.U.; Ziebeck KR, A. The magnetic and structural properties of the magnetic shape memory compound Ni2Mn1.44Sn0.56. J. Phys. Condens. Matter 2006, 18, 2249. [Google Scholar] [CrossRef]
  26. Krenke, T.; Duman, E.; Acet, M.; Wassermann, E.F.; Moya, X.; Mañosa, L.; Ouladdiaf, B. Magnetic superelasticity and inverse magnetocaloric effect in Ni-Mn-In. Phys. Rev. B 2007, 75, 104414. [Google Scholar] [CrossRef]
  27. Krenke, T.; Duman, E.; Acet, M.; Acet, M.; Wassermann, E.F.; Moya, X.; Mañosa, L.; Planes, A. Inverse magnetocaloric effect in ferromagnetic Ni–Mn–Sn alloys. Nat. Mater. 2005, 4, 450–454. [Google Scholar] [CrossRef]
  28. Aseguinolaza, I.R.; Orue, I.; Svalov, A.V.; Wilson, K.; Müllner, P.; Barandiarán, J.M.; Chernenko, V.A. Martensitic transformation in Ni–Mn–Ga/Si (100) thin films. Thin Solid Film. 2014, 558, 449–454. [Google Scholar] [CrossRef]
  29. Chopra, H.D.; Ji, C.; Kokorin, V.V. Magnetic-field-induced twin boundary motion in magnetic shape-memory alloys. Phys. Rev. B 2000, 61, R14913. [Google Scholar] [CrossRef]
  30. Acet, M.; Duman, E.; Wassermann, E.F.; Manosa, L.; Planes, A. Coexisting ferro-and antiferromagnetism in Ni2MnAl Heusler alloys. J. Appl. Phys. 2002, 92, 3867–3871. [Google Scholar] [CrossRef]
  31. Graf, T.; Casper, F.; Winterlik, J.; Balke, B.; Fecher, G.H.; Felser, C. Crystal structure of new Heusler compounds. Z. Für Anorg. Und Allg. Chem. 2009, 635, 976–981. [Google Scholar] [CrossRef]
  32. Liu, J.; Liu, Z.; Jin, X. Origin of tweed in Au–Cu–Al alloys. Philos. Mag. 2014, 94, 56–72. [Google Scholar] [CrossRef]
  33. Müller, C.M.; Sologubenko, A.S.; Gerstl, S.S.; Spolenak, R. On spinodal decomposition in Cu–34 at.% Ta thin films–An atom probe tomography and transmission electron microscopy study. Acta Mater. 2015, 89, 181–192. [Google Scholar] [CrossRef]
  34. Bruno, N.M.; Salas, D.; Wang, S.; Roshchin, I.V.; Santamarta, R.; Arroyave, R.; Karaman, I. On the microstructural origins of martensitic transformation arrest in a NiCoMnIn magnetic shape memory alloy. Acta Mater. 2018, 142, 95–106. [Google Scholar] [CrossRef]
  35. Yano, T.; Murakami, Y.; Kainuma, R.; Shindo, D. Interaction between magnetic domain walls and antiphase boundaries in Ni2Mn (Al, Ga) studied by electron holography and Lorentz microscopy. Mater. Trans. 2007, 48, 2636–2641. [Google Scholar]
  36. Zuo, S.; Liu, Y.; Zhang, Y.; Xiong, J.; Liu, J.; Qiao, K.; Shen, B. In situ TEM study on diversified martensitic transition behaviour in Ni50Mn35In15 alloys. Nanoscale 2019, 11, 4999–5004. [Google Scholar] [CrossRef]
  37. Kuznetsova, E.I.; Blinova, Y.V.; Sudareva, S.V.; Krinitsina, T.P.; Bobylev, I.B.; Romanov, E.P. Spinodal decomposition of the nonstoichiometric YBa2Cu3O7-[delta] compound at room temperature. Phys. Met. Metallogr. 2006, 102, 213. [Google Scholar] [CrossRef]
  38. Bobylev, I.B.; Kuznetsova, E.I.; Krinitsina, T.P.; Zyuzeva, N.A.; Sudareva, S.V.; Romanov, E.P. Restoration of the microstructure of Ba2YCu3O7−δ at T > 900 °C after low-temperature decomposition. Phys. Met. Metallogr. 2011, 112, 165–169. [Google Scholar] [CrossRef]
  39. Kaletina, Y.V.; Kabanova, I.G.; Frolova, N.Y.; Gundyrev, V.M.; Kaletin, A.Y. Crystallographic specific features of the martensitic structure of Ni47Mn42In11 alloy. Phys. Solid State 2017, 59, 2008–2015. [Google Scholar] [CrossRef]
  40. Kaletina, Y.V.; Kabanova, I.G.; Frolova, N.Y.; Kaletin, A.Y. Crystal structure peculiarities of martensite in the Ni47Mn42In11alloy that underwent forward and reverse phase transformations. Phys. Met. Metallogr. 2018, 119, 383–387. [Google Scholar] [CrossRef]
  41. Erkartal, B.; Duppel, V.; Niemann, R.; Schultz, L.; Fähler, S.; Schürmann, U.; Kienle, L. Structure and composition of magnetocaloric Ni–Mn–In–Co thin films on the nanoscale—A TEM study. Adv. Eng. Mater. 2012, 14, 710–715. [Google Scholar] [CrossRef]
  42. Potekaev, A.I. Long-period states of ordered metal alloys. 1. Analysis of structural features. Russ. Phys. J. 1995, 38, 549–562. [Google Scholar] [CrossRef]
Figure 1. Graph showing the dependence of the sample’s magnetization (M) on temperature (T): (a)—in a field of 0.018 T; (b)—in fields of 0.05, 1, 3.5, 10, and 13.5 T; (c)—the dependence magnetization (M) on magnetic field (T) at temperatures 5, 200, 273, 319 K.
Figure 1. Graph showing the dependence of the sample’s magnetization (M) on temperature (T): (a)—in a field of 0.018 T; (b)—in fields of 0.05, 1, 3.5, 10, and 13.5 T; (c)—the dependence magnetization (M) on magnetic field (T) at temperatures 5, 200, 273, 319 K.
Nanomaterials 13 01385 g001aNanomaterials 13 01385 g001b
Figure 2. Temperature dependence of ∆Tad obtained upon sequential heating sample in magnetic fields of 2, 5 and 10 T.
Figure 2. Temperature dependence of ∆Tad obtained upon sequential heating sample in magnetic fields of 2, 5 and 10 T.
Nanomaterials 13 01385 g002
Figure 3. X-ray diffraction pattern of the Ni46Mn41In13 alloy at room temperature.
Figure 3. X-ray diffraction pattern of the Ni46Mn41In13 alloy at room temperature.
Nanomaterials 13 01385 g003
Figure 4. Micrograph of the Ni46Mn41In13 alloy (a) and microdiffraction patterns (b,c) from regions with different structures.
Figure 4. Micrograph of the Ni46Mn41In13 alloy (a) and microdiffraction patterns (b,c) from regions with different structures.
Nanomaterials 13 01385 g004
Figure 5. Bright-field (a) TEM image of the Ni46Mn41In13 alloy in the austenitic state at room temperature and the corresponding microelectron diffraction pattern (b) obtained from the entire area. The grating L21 is shown in blue, with the [110] zone axis.
Figure 5. Bright-field (a) TEM image of the Ni46Mn41In13 alloy in the austenitic state at room temperature and the corresponding microelectron diffraction pattern (b) obtained from the entire area. The grating L21 is shown in blue, with the [110] zone axis.
Nanomaterials 13 01385 g005
Figure 6. Structures of the Ni46Mn41In13 alloy in the austenitic state at room temperature: (a,b) bright-field TEM images; (c) microelectron diffraction pattern obtained from area (a); (df) dark-field images in reflections (2 24 ¯ )L21, (2 2 ¯ 0)L21, and triple reflex, respectively. Zone axis [110]L21.
Figure 6. Structures of the Ni46Mn41In13 alloy in the austenitic state at room temperature: (a,b) bright-field TEM images; (c) microelectron diffraction pattern obtained from area (a); (df) dark-field images in reflections (2 24 ¯ )L21, (2 2 ¯ 0)L21, and triple reflex, respectively. Zone axis [110]L21.
Nanomaterials 13 01385 g006
Figure 7. High resolution photomicrographs from the tweed texture area of the Ni46Mn41In13 alloy in the austenitic state at room temperature (a) and selected enlarged fragment of the photo (b).
Figure 7. High resolution photomicrographs from the tweed texture area of the Ni46Mn41In13 alloy in the austenitic state at room temperature (a) and selected enlarged fragment of the photo (b).
Nanomaterials 13 01385 g007
Figure 8. Bright-field (a) and dark-field (bd) TEM images of the structure of the Ni46Mn41In13 alloy in the austenitic state and the corresponding microelectron diffraction pattern on the inset, obtained from the entire region (a), zone axis [110]L21, (b) in reflection (1 1 ¯ 1)L21, (c) in the reflection (004)L21, (d) in the triplet of reflections).
Figure 8. Bright-field (a) and dark-field (bd) TEM images of the structure of the Ni46Mn41In13 alloy in the austenitic state and the corresponding microelectron diffraction pattern on the inset, obtained from the entire region (a), zone axis [110]L21, (b) in reflection (1 1 ¯ 1)L21, (c) in the reflection (004)L21, (d) in the triplet of reflections).
Nanomaterials 13 01385 g008
Figure 9. Bright-field image of the structure of the Ni46Mn41In13 alloy (a) (inset: SEAD electron diffraction pattern from the region in the upper right corner (tweed contrast) with the zone axis [110]L21 and dark-field images in reflections ( 2 ¯ 20)L21 (b) and ( 4 ¯ 40)L21 (c).
Figure 9. Bright-field image of the structure of the Ni46Mn41In13 alloy (a) (inset: SEAD electron diffraction pattern from the region in the upper right corner (tweed contrast) with the zone axis [110]L21 and dark-field images in reflections ( 2 ¯ 20)L21 (b) and ( 4 ¯ 40)L21 (c).
Nanomaterials 13 01385 g009
Figure 10. STEM images of local content of the elements O, In, Mn, Ni (on the graphs from bottom to top) along the scanning line with a large step (a) and a smaller step (b). The arrows indicate the scan lines. Distribution of indium (c) in the area circled in (b).
Figure 10. STEM images of local content of the elements O, In, Mn, Ni (on the graphs from bottom to top) along the scanning line with a large step (a) and a smaller step (b). The arrows indicate the scan lines. Distribution of indium (c) in the area circled in (b).
Nanomaterials 13 01385 g010
Figure 11. A–M boundary (100 K) at different magnifications (upper right corner—austenite, lower left corner—martensite), bright-field images.
Figure 11. A–M boundary (100 K) at different magnifications (upper right corner—austenite, lower left corner—martensite), bright-field images.
Nanomaterials 13 01385 g011aNanomaterials 13 01385 g011b
Figure 12. Structures of the Ni46Mn41In13 alloy at the A–M interface at 215 K: bright-field image (a), combined austenite–martensite electron diffraction pattern (b), electron diffraction patterns from the austenite region (c), A–M (“B” in circle) boundary (d) and martensite (e). Zone axis [110]L21.
Figure 12. Structures of the Ni46Mn41In13 alloy at the A–M interface at 215 K: bright-field image (a), combined austenite–martensite electron diffraction pattern (b), electron diffraction patterns from the austenite region (c), A–M (“B” in circle) boundary (d) and martensite (e). Zone axis [110]L21.
Nanomaterials 13 01385 g012
Figure 13. Structure of the Ni46Mn41In13 alloy at 208 K: (a)—electron diffraction pattern of the austenite structure, zone axis [001]L21; (b)—electron diffraction pattern of the martensite structure, zone axis [0 1 ¯ 0]14M; (c)—fine structure of martensite, dark-field image in the (00 14 ¯ )M reflection (circled).
Figure 13. Structure of the Ni46Mn41In13 alloy at 208 K: (a)—electron diffraction pattern of the austenite structure, zone axis [001]L21; (b)—electron diffraction pattern of the martensite structure, zone axis [0 1 ¯ 0]14M; (c)—fine structure of martensite, dark-field image in the (00 14 ¯ )M reflection (circled).
Nanomaterials 13 01385 g013
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

Kuznetsov, D.D.; Kuznetsova, E.I.; Mashirov, A.V.; Loshachenko, A.S.; Danilov, D.V.; Mitsiuk, V.I.; Kuznetsov, A.S.; Shavrov, V.G.; Koledov, V.V.; Ari-Gur, P. Magnetocaloric Effect, Structure, Spinodal Decomposition and Phase Transformations Heusler Alloy Ni-Mn-In. Nanomaterials 2023, 13, 1385. https://doi.org/10.3390/nano13081385

AMA Style

Kuznetsov DD, Kuznetsova EI, Mashirov AV, Loshachenko AS, Danilov DV, Mitsiuk VI, Kuznetsov AS, Shavrov VG, Koledov VV, Ari-Gur P. Magnetocaloric Effect, Structure, Spinodal Decomposition and Phase Transformations Heusler Alloy Ni-Mn-In. Nanomaterials. 2023; 13(8):1385. https://doi.org/10.3390/nano13081385

Chicago/Turabian Style

Kuznetsov, D. D., E. I. Kuznetsova, A. V. Mashirov, A. S. Loshachenko, D. V. Danilov, V. I. Mitsiuk, A. S. Kuznetsov, V. G. Shavrov, V. V. Koledov, and P. Ari-Gur. 2023. "Magnetocaloric Effect, Structure, Spinodal Decomposition and Phase Transformations Heusler Alloy Ni-Mn-In" Nanomaterials 13, no. 8: 1385. https://doi.org/10.3390/nano13081385

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