Anomalous microstructure and magnetocaloric properties in off-stoichiometric La–Fe–Si and its hydride
He Chuna),b),c), Zhang Ming-Xiaoa),b), Shao Yan-Yana),b), Dong Jing-Dua),b), Yan A-Rua),b), Liu Jian†a),b)
Key Laboratory of Magnetic Materials and Devices, Ningbo Institute of Material Technology and Engineering, Chinese Academy of Sciences, Ningbo 315201, China
Zhejiang Province Key Laboratory of Magnetic Materials and Application Technology, Ningbo Institute of Materials Technology and Engineering, Chinese Academy of Sciences, Ningbo 315201, China
Nano Science and Technology Institute, University of Science and Technology of China, Hefei 230026, China

Corresponding author. E-mail: liujian@nimte.ac.cn

*Project supported by the National Natural Science Foundation of China (Grant No. 51371184) and the Outstanding Youth Fund of Zhejiang Province, China (Grant No. LR14E010001).

Abstract

In the present work we reported the phase formation, microstructure, magnetocaloric effect and hydrogenation behavior of La-rich La1.7Fe11.6Si1.4 alloy. In this off-stoichiometric La(Fe,Si)13 alloy, the NaZn13-type La(Fe,Si)13 matrix phase shows faceted grains, with the Cr5B3-type La5Si3 used as the secondary phase distributed intergranularly. Such a peculiar morphology quickly forms upon one day annealing. In La1.7Fe11.6Si1.4 alloy, we have observed a significant field dependence of magnetostructural transition temperature (∼ 6.3 K/T), resulting in a large and table-like entropy change (Δ S ∼ 18 J/kg·K in 2 T) over a broad temperature range (∼ 10 K). Upon hydrogenation, the maximum value of Δ S keeps almost unchanged, while the Curie temperature increases up to 350 K. These results indicate that the investigated off-stoichiometric La(Fe,Si)13 alloy is a promising magnetic material for magnetic refrigerators.

PACS: 75.30.–m; 75.50.–y
Keyword: La–Fe–Si alloys; Magnetocaloric effect; off-stoichiometric; magnetic refrigeration
1. Introduction

Currently, magnetic materials showing a large magnetocaloric effect (MCE) have become a topic of growing interest, since the magnetic refrigeration technique based on the MCE is regarded as a promising alternative to the conventional vapor-cycle refrigeration.[1, 2] A large MCE is always observed in magnetostructural transition materials where the contribution of lattice entropy to structural transition is considerable.[3] Among such materials, La– Fe– Si compounds with NaZn13-type structure (the 1:13 phase) are one of the most promising magnetic refrigerants, as they hold several advantages towards magnetic cooling applications, such as low cost and nontoxic consisting elements, high magnetic moment and MCE, relatively low driven field and little hysteresis, and tunable Curie temperature (Tc) by elemental substitution or interstitial insertion.[410] The solidified microstructure of as-cast stoichiometric La(Fe, Si)13 compounds usually consists of secondary phases including α -Fe and LaFeSi (the 1:1:1 phase). It is required at least one week homogenization around 1323 K to obtain the single 1:13 phase in La(Fe, Si)13 bulk alloys.[11] Due to this energy consuming and low efficient preparation, alternative techniques, such as melt spinning, [12] strip casting, [13] and powder metallurgical processing, [14, 15] have been developed to shorten annealing time. Till now, almost all previous studies concerned the synthesis methods, magnetic entropy change (Δ S), and adiabatic temperature change (Δ Tad) for La(Fe, Si)13-based stoichiometric compounds. Very recently, an off-stoichiometric composition LaFe8.8Si2.2 was reported, which has a single 1:13 phase directly formed in a solidification state, but shows a rather low value of Δ S (about 2 J/kg· K in 1.5 T).[16] In this paper, we present a new La-rich composition of La1.7Fe11.6Si1.4. Large Δ S of about 18 J/kg· K in 2 T in this alloy and its hydride was obtained. A peculiar microstructure and corresponding first-order magnetostructural transition, and hydrogenation behavior are investigated.

2. Experimental procedures

An ingot with the nominal composition of La1.7Fe11.6Si1.4 was prepared by induction melting in an argon atmosphere. As-cast alloys were annealed at 1323 K for different times from 1 to 7 days in a quartz ampoule filled with argon, and then quenched into ice water. Hydrogenation was performed in a furnace filled with 2 atm H2 atmosphere at 473 K for 5 hours to saturate H concentration.[17] The H concentration was determined by hot extraction method (ONH, TCH-600). The crystal structure was studied using a powder x-ray diffraction (XRD, Bruker D8, Cu-Kα radiation). The microstructure was observed by scanning electron microscope (SEM, FEI Quanta 250 FEG). The phase composition was analyzed by electron microprobe (EM, JXA-8100), which has a high accuracy about ± 0.1 wt.%. Magnetic measurements were carried out using a Quantum Design MPMS SQUID vibrating sample magnetometer.

3. Results and discussion

Apart from usually observed α -Fe and 1:1:1 phases in stoichiometric La(Fe, Si)13 compositions, an additional phase in La1.7Fe11.6Si1.4 was found by XRD presented in Fig. 1(a), which can be well indexed to be the La5Si3 phase (tetragonal, I4/mcm, a = 0.788 nm, c = 1.437 nm). Still, the α -Fe is the main phase in the as-cast sample. The phase constitution dramatically changed upon a short-time annealing for 1 day, as shown in the XRD pattern in Fig. 1(b). The phase constitution was almost unchanged upon annealing for 3 days as shown in the XRD pattern in Fig. 1(c). Full hydrogenation shifts all XRD peaks to lower angles excepted for the α -Fe phase, indicating that the values of lattice constant a increase from 1.142 nm to 1.155 nm for 1:13 phase, and for La5Si3 phase the values of lattice constant a and c increase from 0.788 nm to 0.798 nm and 1.437 nm to 1.483 nm respectively, as shown in Fig. 1(d). Figures 2(a)– 2(c) shows the microstructure of the specimens annealed at 1323 K for 0– 3 days. The as-cast sample contains the α -Fe (black phase), 1:1:1 phase (gray phase) and La5Si3 phase (white phase). From the SEM observation in Figs. 2(b) and 2(c), the 1:13 grains exhibiting a faceted morphology are surrounded by La5Si3. The volume fraction of the 1:13 phase nearly reaches 80%, whereas the La5Si3 phase was retained, but there observed little amount of α -Fe and no 1:1:1 phase. A prolonged annealing up to 7 days did not impact this microstructure very much (not shown here). According to the previously established La– Fe– Si ternary phase diagram, [18] the La5Si3 phase locates at the La rich boundary. The formation of this Si-rich and Fe-free La5Si3 phase is expected to result in the 1:13 matrix being depleted of Si, thus the concentration of Fe increases in the 1:13 matrix and the MCE is enhanced.[5] This assumption was confirmed by the result of electron microprobe, which suggests the real composition of the 1:13 matrix as La8.18Fe83.59Si8.24 ( = La1.16Fe11.83Si1.17). As the La(Fe, Si)13 compound with a lower Si content has a higher capability to absorb hydrogen due to a large weighted average Fe– Fe distance, [19] it seems reasonable that the present 1:13 hydride matrix contains high H content. The amount of H in this sample is measured to be 3100 ppm. This value is much higher than that in La1.0Fe11.8Si1.2(∼ 2320 ppm).[19] However, we are not able to exclude the H in the secondary phase La5Si3 without knowing its hydrogen absorption capability. Interestingly, hydrogen decrepitation process leads to single crystalline particles by almost keeping the initial shape and size of the non-hydrogenated 1:13 grains, as shown in Fig. 2(d). Also, some residual La5Si3 and α -Fe phases adhesive on fracture surfaces are visible. The form of single crystalline La– Fe– Si particles would seem more attractive for magnetic refrigerant applications.

Fig. 1. XRD patterns of as-cast, annealed, and hydrogenated La1.7Fe11.6Si1.4.

Fig. 2. SEM backscattered electron images of as-cast bulk La1.7Fe11.6Si1.4 (a), annealed for 1 day (b) and 3 day (c) at 1323 K. (d) Single crystalline hydride particle.

It has been known that in order to exploit the potential of Δ Tad in a finite field for first-order transition materials, a sharp magnetostructural transition is desirable, and more importantly the field dependence of transition temperature (dTc/dH) has to be optimized.[3, 20, 21] The optimal value of dTc/dH for Ni49.8Mn35In15.2 is 7.8 K/T, [3] and 6.5 K/T for La(Fe, Si)13system.[20] From the magnetization versus temperature at different fields (0.2 and 1 T) in Fig. 3, the value of dTc/dH for La1.7Fe11.6Si1.4 is calculated to be 6.3 K/T that is very close to the theoretical optimized one suggested by Sandeman.[20] This ideal case would give rise to the maximum Δ Tad of 13 K.[20] A large Δ Tad of 5.5 K in 2 T has been found by the direct experimental measurement for La1.7Fe11.6Si1.4 (not shown here), but still far from the theoretical value. This discrepancy implies that a higher driven field is required to reach the upper bound of cooling effect. Nevertheless, the large dTc/dH and steep transition in first order transitions are favorable for the table-like temperature dependence of MCE, as reported in antiperovskite Mn3GaC alloys.[22, 23] In comparison, stoichiometric composition La1.0Fe11.6Si1.4 shows a smaller value of dTc/dH about 3.6 K/T and slightly smeared magnetostructural transition at high fields.

Fig. 3. Comparison of magnetization as a function of temperature on applying different field in La1.7Fe11.6Si1.4 and La1.0Fe11.6Si1.4. The insert shows the shift of transition temperature by increasing magnetic field.

Figure 4 shows the isotherms and related magnetic entropy change as a function of temperature and field for La1.7Fe11.6Si1.4 annealed for 3 days and its hydride La1.7Fe11.6Si1.4H2.9. For La1.7Fe11.6Si1.4, the magnetization curves at temperatures well above Tc (173 K) show very small low-field susceptibility due to a little amount of α -Fe phase. A sharp change in magnetization above Tc from 184 K to 175 K takes place at a critical field. Combined the well-defined onset field with the observed large magnetic hysteresis width (∼ 1 T), this compound undergoes a typical first-order magnetostructural transition.[24] The entropy change Δ S is calculated by using the Maxwell relation , where μ 0 and H are permeability and magnetic field, respectively. The maximum Δ S value is 18 J/kg· K when applying a field of 2 T. An interesting feature is that the Δ S peak broadens asymmetrically toward higher temperature with increasing field, which can be ascribed to the field-induced itinerant electron metamagnetic transition.[4] This characteristic causes a peculiar table-like Δ S curve over a wide temperature range of 10 K in 2 T, which is desirable for enhancement of the refrigerant capacity in an ideal Ericsson cycle magnetic refrigeration. In comparison, the temperature range for the Δ S plateau induced by 2 T field change has been reported to be 7 K in stoichiometric La1.0Fe11.9Si1.1, and 2 K in La1.0Fe11.8Si1.2.[25] The full hydrogenated La1.7Fe11.6Si1.4H2.9 sample exhibits much smoother magnetization curves, and its hysteresis width is significantly reduced to 0.1 T. Hysteresis for a system includes two components: intrinsic hysteresis due to lattice incompatibility between the transforming phases and extrinsic hysteresis caused by microstructural features such as grain boundaries and secondary phases.[26] It is assumed that the current hydrogenation brings about the single crystalline 1:13 powders, and very likely removes grain boundaries as well as phase boundaries with the La5Si3 phase. Most importantly, the hydrogenation decrepitation does not impair the MCE. The Δ S still keeps a high level of 17 J/kg· K in 2 T. Moreover, a relatively lower field (e.g., 1 T, which can be more easily generated with Nd– Fe– B permanent magnets) can induce a large Δ S of 14 J/kg· K. The MCE is more pronounced than that observed in stoichiometric La1.0Fe11.74− yMnySi1.26H1.53 (about 12 J/kg· K in 1.6 T), [27] indicating that the H distributes more homogeneously in La1.7Fe11.6Si1.4H2.9. We believe that the present novel microstructure exhibiting the 1:13 grains surrounded by the La5Si3 phase is helpful for hydrogen absorption. The magnetic field dependence of Δ S in samples before and after hydrogenation is also shown in Fig. 4. In the case of non-hydrogenated sample, the Δ S remains constant at low field, rapidly increases in a linear manner at a critical field, then saturates at higher field. This discontinuity becomes smeared in hydrogenated sample, which can be interpreted by the disappearance of Δ S plateau.

Fig. 4. (a), (b) Magnetization versus field at various temperatures, (c), (d) entropy change versus temperature at various fields, and (e), (f) entropy versus field at various temperatures, for non-hydrogenated (left column) and hydrogenated La1.7Fe11.6Si1.4 (right column).

Last, we compare the influence of hydrogenation on the Tc and MCE between La1.7Fe11.6Si1.4 and La1.0Fe11.6Si1.4, as shown in Fig. 5. One aspect is that the decrease of the Δ S upon hydrogenation in La1.0Fe11.6Si1.4 is about 20% (the data taken from Ref. [15]), whereas it keeps constant in La-rich sample. On the other hand, the Tc of La1.7Fe11.6Si1.4 is lower than that of La1.0Fe11.6Si1.4, but hydrogen shifts Tc of La1.7Fe11.6Si1.4 to a higher value exceeding that in La1.0Fe11.6Si1.4. This phenomenon has been explained by the different space for H in the unit cell.[20] From this result, it can be concluded that the 1:13 matrix phase of La1.7Fe11.6Si1.4 should contain more H than the single 1:13 phase in La1.0Fe11.6Si1.4 (about 2041 ppm), although the La5Si3Hy phase also contributes H concentration to the sum H amount.

Fig. 5. Temperature dependence of entropy change for 2 T field change in La1.7Fe11.6Si1.4 and La1.0Fe11.6Si1.4 and their hydrides.

4. Conclusions

We have reported a new off-stoichiometric La1.7Fe11.6Si1.4. The dual-phase structure with 1:13 (80 vol.%) and La5Si3 phases can be obtained by only 1 day annealing. This novel composition exhibits table-like large entropy change. Hydrogenation results in the single crystalline 1:13 particles with size of 60– 80 μ m. The hydride particles contain high H concentration and high Tc, and exhibit large Δ S of 14 J/kg· K in 1 T accompanying with small hysteresis losses. These features suggest that the present La-rich alloy is a highly suitable refrigerant candidate for magnetic refrigerators.

Reference
1 Liu J 2014 Chin. Phys. B 23 047503 DOI:10.1088/1674-1056/23/4/047503 [Cited within:1]
2 Gschneidner K A, Pecharsky V K and Tsokol A O 2005 Rep. Prog. Phys. 68 1479 DOI:10.1088/0034-4885/68/6/R04 [Cited within:1]
3 Liu J, Gottschall T, Skokov K, Moore J and Gutfleisch O 2012 Nat. Mater. 11 620 DOI:10.1038/nmat3334 [Cited within:3]
4 Hu F X, Shen B G, Sun J R, Cheng Z H, Rao G H and Zhang X X 2001 Appl. Phys. Lett. 78 3675 DOI:10.1063/1.1375836 [Cited within:2]
5 Fujieda S, Fujita A and Fukamichi K 2002 Appl. Phys. Lett. 81 1276 DOI:10.1063/1.1498148 [Cited within:1]
6 Shen B G, Sun J R, Hu F X, Zhang H W and Cheng Z H 2009 Adv. Mater. 21 4545 DOI:10.1002/adma.v21:45 [Cited within:1]
7 Lyubina J, Gutfleisch O, Kuz’min M D and Richter M 2009 J. Mag. Mag. Mater. 321 3271 [Cited within:1]
8 Liu J, Moore J, Skokov K, Krautz M, Löwe K, Barcza A, Katter M and Gutfleisch O 2012 Scripta Mater. 67 584 DOI:10.1016/j.scriptamat.2012.05.039 [Cited within:1]
9 Shen B G, Hu F X and Dong Q Y 2013 Chin. Phys. B 22 017502 DOI:10.1088/1674-1056/22/1/017502 [Cited within:1]
10 Shen J, Li Y X, Sun J R and Shen B G 2009 Chin. Phys. B 18 2058 DOI:10.1088/1674-1056/18/5/055 [Cited within:1]
11 Liu J, Krautz M, Skokov K, Woodcock T G and Gutfleisch O 2011 Acta Mater. 59 3602 DOI:10.1016/j.actamat.2011.02.033 [Cited within:1]
12 Yan A, Muller K H and Gutfleisch O 2005 J. Appl. Phys. 97 036102 DOI:10.1063/1.1844605 [Cited within:1]
13 Zhang M X, Liu J, Zhang Y, Dong J D, Yan A R, Skokov K P and Gutfleisch O 2015 J. Mag. Mag. Mater. 377 90 DOI:10.1016/j.jmmm.2014.10.035 [Cited within:1]
14 Katter M, Zellmann V, Reppel G W and Uesuener K 2008 IEEE Trans. Mag. 44 3044 DOI:10.1109/TMAG.2008.2002523 [Cited within:1]
15 Phejar M, Paul-Boncour V and Bessais L 2010 Intermetallics 18 2301 DOI:10.1016/j.intermet.2010.07.022 [Cited within:2]
16 Bao L H, Wei W, Fan W D and Tegus O 2014 J. Alloy. Comp. 589 416 DOI:10.1016/j.jallcom.2013.11.173 [Cited within:1]
17 Fu B, Long Y, Shi P J, Ma T, Bao B, Yan A R and Chen R J 2009 Chin. Phys. B 18 1674 DOI:10.1088/1674-1056/18/4/066 [Cited within:1]
18 Niitsu K and Kainuma R 2012 Intermetallics 20 160 DOI:10.1016/j.intermet.2011.06.005 [Cited within:1]
19 Krautz M, Skokov K, Gottschall T, Teixeira C S, Waske A, Liu J, Schultz L and Gutfleisch O 2014 J. Alloy. Comp. 598 27 DOI:10.1016/j.jallcom.2014.02.015 [Cited within:2]
20 Sand eman K G 2012 Scripta Mater. 67 566 DOI:10.1016/j.scriptamat.2012.02.045 [Cited within:5]
21 Skokove K, Muller K H, Moore J, Liu J, Karpenkov A Y, Krautz M and Gutfleisch O 2013 J. Alloy. Comp. 552 310 DOI:10.1016/j.jallcom.2012.10.008 [Cited within:1]
22 Yu M H, Lewis L H and Moodenbaugh A R 2003 J. Appl. Phys. 93 10128 DOI:10.1063/1.1574591 [Cited within:1]
23 Tohei T, Wada H and Kanomata T 2003 J. Appl. Phys. 94 1800 DOI:10.1063/1.1587265 [Cited within:1]
24 Shen J, Wang F, Zhao J L, Wu J F, Gong M Q, Hu F X, Li Y X, Sun J R and Shen B G 2010 J. Appl. Phys. 10709A909 [Cited within:1]
25 Gutfleisch O, Yan A and Muller K H 2005 J. Appl. Phys. 9710M305 [Cited within:1]
26 Moore J D, Morrison K, Sand eman K G, Katter M and Cohen L F 2009 Appl. Phys. Lett. 95 252504 DOI:10.1063/1.3276565 [Cited within:1]
27 Barcza A, Katter M, Zellmann V, Russek S, Jacobs S and Zimm C 2011 IEEE Trans. Mag. 47 3391 DOI:10.1109/TMAG.2011.2147774 [Cited within:1]