An efficient composite magnetocaloric material with a tunable temperature transition in K-deficient manganites

R. Skini*a, M. Khlifia and E. K. Hlilb
aLaboratoire de Physique Appliquée, Faculté des Sciences de Sfax, Université de Sfax, B.P. 802, Sfax 3018, Tunisia. E-mail: skini.ridha@ymail.com; Fax: +216 74676609; Tel: +216 25408086
bInstitut Néel, CNRS et Université Joseph Fourier, BP 166, F-38042 Grenoble Cedex 9, France

Received 18th December 2015 , Accepted 29th January 2016

First published on 29th January 2016


Abstract

This paper reports on a magnetic material with high steady relative cooling power (RCP) over a temperature range from 325 to 275 K induced by potassium-deficiency in polycrystalline samples of La0.8K0.2−xxMnO3 (x = 0, 0.1 and 0.2). The possibility to adjust the temperature transition close to room temperature exists by changing the potassium deficiency rate. All the samples were synthesized using the citrate-gel method. X-ray diffraction and magnetization measurements were conducted to examine the crystallographic structure and magnetocaloric properties. All the samples were found to be a single phase and crystallize in rhombohedral symmetry with R[3 with combining macron]c space group. From the magnetic measurements, a second-order magnetic phase transition from the ferromagnetic to paramagnetic state was noticed at the Curie temperature (TC), which was found to decrease from 325 to 275 K when the potassium-deficiency rate increased. Besides, the magnetocaloric effect (MCE) as well as the RCP has been estimated. The obtained results have confirmed that our manganite with potassium-deficiency has significant advantages for magnetic refrigeration. Finally, the critical behavior associated with the magnetic phase transition reveals that the estimated critical exponents are consistent with the prediction of the mean field theory (β = 0.5, γ = 1, and δ = 3) for the x = 0.00 and x = 0.10 samples, while the estimated critical exponents for x = 0.2 were consistent with the prediction of the 3D-Heisenberg model (β = 0.365, γ = 1.336 and δ = 4.8).


1. Introduction

Rare-earth based manganites with the general formula R1−xBxMnO3 (R-rare-earth and B-bivalent ion) has been the focus of many researchers from a fundamental point of view and thanks to their possible potential applications.1–7 The strong compound interactions with remarkable correlation between transport, magnetic and structural properties of such materials are still the interest of fundamental studies.1 Furthermore, manganites reveal several fascinating functional properties, such as colossal magnetoresistance (CMR) and large magnetocaloric effect (MCE).1,7,8 It is recognized that the MCE-based magnetic refrigeration technology (MRT) is a promising substitute for conventional gas compression refrigeration techniques.8–10 In fact, compared to conventional gas compression, magnetic refrigeration offers a greater cooling efficiency. Besides, the use of hazardous elements related to gas compression-based refrigeration is considered as environmentally precarious.9 The selection of a suitable magnetic refrigerant is considered to be the most important factor in governing the performance of a magnetic refrigerator. Examples of the compounds that have been considered as potential magnetic refrigerants (MR) are gadolinium and some other intermetallic alloys.10 Nevertheless, research studies have recently revealed that manganites also have significant qualities that present them as prospective candidates as magnetic refrigerants.7–9 These materials show large MCE and their magnetocaloric properties can be tuned by changing the doping concentration, which offers more flexibility to their applications. Compared to metallic compounds, the resistivity of manganites, which can significantly reduce the eddy current losses, is higher and is an additional advantage of these compounds.9 Furthermore, the latter show great chemical stability and their production cost is relatively lower compared to other potential refrigerants. Bahl et al. recently reported an effective application of manganites in an active magnetic regenerator, demonstrating their feasibility in practical magnetic refrigeration.11 In addition, our original idea was to change the physical properties by creating a deficiency in the materials. However, only a few studies5–7 have been proposed to discuss deficiency in the manganite system. As is well known, deficiency in this system leads to a change in the Mn3+/Mn4+ ratio.

This research work aims to study the effect of potassium deficiency on the structural, magnetic and magnetocaloric properties of polycrystalline La0.8K0.2−xxMnO3 manganite. We report a compound showing a significant magnetocaloric effect with a tunable temperature transition near room temperature.

2. Experimental

The La0.8K0.2−xxMnO3 (x = 0, 0.1 and 0.2) samples were prepared using the citrate-gel method. In fact, stoichiometric amounts of the nitrate precursor reagents La(NO3)3·6H2O, Mn(NO3)2·4H2O and KNO3 were dissolved in water and mixed with ethylene glycol and citric acid, forming a stable solution. The molar ratio of metal[thin space (1/6-em)]:[thin space (1/6-em)]citric acid was 1[thin space (1/6-em)]:[thin space (1/6-em)]1. The solution was then heated on a thermal plate under constant sintering at 80 °C to remove the excess water and obtain a viscous gel. The obtained gel was decomposed at 300 °C and the resulting precursor powder was heated in air at 500 °C, 600 °C and 700 °C for 24 h to ameliorate crystallinity. Subsequently, the powder was pelletized and sintered at 700 °C for 12 h. The samples were quenched in air by removing from the furnace. The phase formation and crystal structure of the powders were confirmed by X-ray diffraction (XRD) using a Cu Kα radiation source. The magnetic measurements were performed using a BS1 and BS2 magnetometer developed at the Néel Institute.

3. Results and discussion

3.1. Structural properties

The X-ray diffraction (XRD) patterns recorded at room temperature for all our synthesized samples La0.8K0.2−xxMnO3 (x = 0, 0.1 and 0.2) are shown in Fig. 1. The superposition of all the peaks shows that all the samples crystallized in the same structure and were efficaciously indexed according to a rhombohedral structure with R[3 with combining macron]c space group, in which the La/K atoms are in the 6a (0,0,1/4) position, Mn is in the 6b (0,0,0) position and O is in the 18e (x, 0,1/4) position. The data were analyzed by the Rietveld method using the Fullprof program.12 Fig. 2 presents the measured and calculated (refined) X-ray diffraction patterns, and the positions of the Bragg reflections for the La0.8K0.2−xxMnO3 samples. The refinement results are listed in Table 1. From these results, it is obvious that the unit cell parameters and volume increase almost linearly with an increase in the K vacancy rate (x). The variation of unit cell volume versus x is plotted in Fig. 3. The introduction of the vacancy (x) in our samples implies a partial conversion of Mn3+ to Mn4+ ions according to the formula La0.83+K0.2−x+0x(Mn0.6−x3+Mn0.4+x4+)O32−. The increase in the vacancy content leads to an increase in the Mn tetravalent ion number, which possesses a smaller ionic radius (rMn4+ = 0.53 Å and rMn3+ Å (ref. 13)). Because several studies have shown that the vacancy has an average radius 〈rV〉 that is not equal to zero,14–17 the decrease in the unit cell volume V with lanthanum vacancies can also be explained by the fact that the average radius of vacancy 〈rV〉 is larger than that of K+.
image file: c5ra27132k-f1.tif
Fig. 1 XRD patterns of the La0.8K0.2−xxMnO3 (x = 0, 0.1 and 0.2) compounds at room temperature.

image file: c5ra27132k-f2.tif
Fig. 2 Observed (open symbols) and calculated (solid lines) X-ray diffraction patterns for La0.8K0.2−xxMnO3 (x = 0, 0.1 and 0.2). The positions of the Bragg reflections are marked with vertical bars. The differences between the observed and the calculated intensities are shown at the bottom of the diagram.
Table 1 Results of Rietveld refinements determined from the XRD patterns recorded at room temperature for the La0.8K0.2−xxMnO3 (x = 0, 0.1 and 0.2) samples
x 0 0.1 0.2
Space group R[3 with combining macron]c R[3 with combining macron]c R[3 with combining macron]c
[thin space (1/6-em)]
Lattice parameter
a = b (Å) 5.50 5.507 5.51
c (Å) 13.36 13.38 13.42
Unit cell volume (Å3) 87.24 87.87 88.24
dMn–O (Å) 1.958 1.962 1.967
θMn–O–Mn (°) 166.3 165.71 165.45
χ2 (%) 1.27 1.51 1.51



image file: c5ra27132k-f3.tif
Fig. 3 Variation of unit cell volume versus the potassium vacancy for the La0.8K0.2−xxMnO3 (x = 0, 0.1 and 0.2) compounds.

In addition, Table 1 proves that while the Mn–O–Mn bond angle decreases, the Mn–O length increases with the doping level x, confirming the increase in the unit cell volume (Fig. 4).


image file: c5ra27132k-f4.tif
Fig. 4 Bond distances (dMn–O) and bond angles 〈Mn–O–Mn〉 as a function of vacancy content for the La0.8K0.2−xxMnO3 (x = 0, 0.1 and 0.2) compounds.

3.2. Magnetic properties

Measurements of the temperature dependence of magnetization M(T) were performed under an applied field of 0.05 T for all the samples and the results are shown in Fig. 5. The M(T) curves reveal that when the temperature increases, all the samples display a ferromagnetic–paramagnetic transition. The paramagnetic to ferromagnetic (PM–FM) transition temperatures (TC) were obtained from the peak of the dM/dT curves (Fig. 6). Thus, it can be clearly noted that the magnetization decreases with an increase in the potassium deficiency rate due to the decrease in the number of Mn3+ ions characterized by their high spin S = 2 and the increase in the number of Mn4+ (S = 3/2) ions according to the La0.83+K0.2−x+0x(Mn0.6−x3+Mn0.4+x4+)O32− equation. Besides, the Curie temperature decreases from 325 to 275 K with an increase in x from 0.00 to 0.2, respectively. The decrease in TC is explained by the reduction of the one-electron bandwidth (W), which is governed by the structural parameters as follows:
image file: c5ra27132k-t1.tif
where γ is the Mn–O–Mn angle, dMn–O is the Mn–O distance and W0 is a positive constant.18 The variation of Curie temperature (TC) as well as the one-electron bandwidth (W) are shown in Fig. 7. Furthermore, the sensitivity of the proposed samples to an applied magnetic field was analyzed. Fig. 8 reveals the variation of the magnetization as a function of the magnetic field at different temperatures on either side of the Curie temperature for the La0.8K0.2−xxMnO3 (x = 0, 0.1 and 0.2) samples. First, the ferromagnetic and paramagnetic properties were confirmed below and above the transition temperature, respectively. The nature of the magnetic transitions was tested using the Banerjee criterion,19 which is based on the Arrott plots (M2 versus H/M). The slope of the H/M versus M2 curve denotes whether a magnetic transition is of a first or second order. In particular, the magnetic transition is of a second order when all the curves have a positive slope. If the curves above the TC show a negative slope in the low M2 region, the transition is of a first order. Fig. 9 shows the H/M versus M2 plots of the isotherms in the vicinity of TC for the samples. Clearly, these isotherms exhibit positive slopes that are indicators of the second-order character of the phase transition.

image file: c5ra27132k-f5.tif
Fig. 5 Variation of magnetization (M) vs. temperature (T) for the La0.8K0.2−xxMnO3 (x = 0, 0.1 and 0.2) compounds measured under an applied magnetic field of 0.05 T.

image file: c5ra27132k-f6.tif
Fig. 6 Variation of dM/dT as a function of temperature (T) for the La0.8K0.2−xxMnO3 (x = 0, 0.1 and 0.2) compounds.

image file: c5ra27132k-f7.tif
Fig. 7 The vacancy content (x) dependence on the Curie temperature (TC) and electron-one bandwidth (W/W0).

image file: c5ra27132k-f8.tif
Fig. 8 Isothermal magnetization of the La0.8K0.2−xxMnO3 (x = 0, 0.1 and 0.2) samples measured at different temperatures around the TC.

image file: c5ra27132k-f9.tif
Fig. 9 Arrott plots (M2 vs. H/M) for the La0.8K0.2−xxMnO3 (x = 0, 0.1 and 0.2) samples around the TC.

3.3. Magnetocaloric effect

The magnetocaloric effect is an inherent property of a magnetic material and its response to the application or removal of a magnetic field. Such a response is maximized when the temperature of the material is near its magnetic ordering temperature (Curie temperature, TC). The total entropy can be defined as the sum of the entropy due to the lattice (SL), the electrons (SE) and the magnetic order (SM). Under a constant pressure, this would be a function of both the temperature (T) and applied magnetic field (H). For an adiabatic process, the application of a magnetic field leads to a decrease in SM, then to an increase in SL, and therefore leads to the heating of the material. By analogy, the suppression of the magnetic field causes the cooling of the material. The magnetocaloric effect is indirectly predicted with Maxwell's relationship20,21 from the magnetization data under different applied magnetic fields from 0 to μ0Hmax.
 
image file: c5ra27132k-t2.tif(1)

Hence, the entropy change ΔSM, which determines the MCE behavior can be numerically calculated from the area enclosed between the two isothermal magnetizations of the M(H,T) curves shown in Fig. 8 and by the use of the following equation:

 
image file: c5ra27132k-t3.tif(2)

Fig. 10 shows the magnetic entropy changes as a function of temperature for the La0.8K0.2−xxMnO3 (x = 0, 0.1 and 0.2) samples under an applied magnetic field up to 5 T. The ΔSmax increases with an increase in the μ0H value for each composition, following the increase in the magnetization and the spin alignment with this magnetic field.


image file: c5ra27132k-f10.tif
Fig. 10 Magnetic entropy change (−ΔSM) as a function of temperature under various magnetic fields between 0.5 and 5 T for the La0.8K0.2−xxMnO3 (x = 0, 0.1 and 0.2) compounds.

As anticipated, the magnetic entropy change (ΔSM) depends on both the applied magnetic field and sample temperature, i.e., ΔSM increases and reaches a maximum value (ΔSMmax) when the temperature approaches the Curie temperature. Interestingly, these curves were found to reveal that all the samples present a large magnetocaloric effect. This ΔSMmax is found to be sensitive to the potassium deficiency. Indeed, under H = 5 T, ΔSMmax is equal to 3.33 J kg−1 K−1 at 325 K, 3.42 J kg−1 K−1 at 300 K and 4.54 J kg−1 K−1 at 275 K for x = 0.00, 0.10 and 0.2, respectively (Fig. 11).


image file: c5ra27132k-f11.tif
Fig. 11 Temperature dependence of the magnetic entropy change corresponding to an applied field of μ0H = 5 T for the La0.8K0.2−xxMnO3 (x = 0, 0.1 and 0.2) compounds.

Moreover, the critical behavior near the paramagnetic to ferromagnetic phase transition temperature has been analyzed based on the field dependence of entropy change. According to Oesterreicher et al.,22 the field dependence of the magnetic entropy change of materials with a second order phase transition can be expressed as ΔSMα(μ0H)n, where n depends on the magnetic state of the sample. The n exponent is found from the fit of the ΔSMmax vs. μ0H curve (Fig. 12(a)), in which n is found to be 0.778; 0.778 and 0.676 in the samples for x = 0.00, 0.10 and 0.2, respectively. It is clear that the value of n is larger than the predicted value of 2/3 (ref. 23 and 24) in the mean field approach for the x = 0.2 sample. This difference can be attributed to the local inhomogeneities or superparamagnetic clusters in the vicinity of a transition temperature.25,26 Then, the other exponents β, γ, and δ, which are associated with the spontaneous magnetization (MS), inverse of initial susceptibility (χ0−1), and magnetization isotherm (MH at TC), respectively,27 are calculated with the help of eqn (3) and (4) using the value of n. The δ exponent is estimated by fitting the M(H) curve at the Curie temperature with the M = DH1/δ equation (D is a critical amplitude),27 this fit is presented in Fig. 12(b), in which the δ exponent is found to be 3.53, 3.11 and 4.44 for x = 0.00, 0.10 and 0.2, respectively.

 
image file: c5ra27132k-t4.tif(3)
 
image file: c5ra27132k-t5.tif(4)


image file: c5ra27132k-f12.tif
Fig. 12 (a) Variation of −ΔSMmax as a function of the applied magnetic field for the La0.8K0.2−xxMnO3 (x = 0, 0.1 and 0.2) samples. The red line indicates the linear fit for n. (b) Magnetization as a function of the magnetic field M(H) at TC (symbols), the solid line shows the fitting with the M = DH1/δ equation.

After a simple calculation using eqn (3) and (4), we find that β = 0.56, γ = 1.42 and δ = 3.53 for the x = 0.00 sample; β = 0.51, γ = 1.69 and δ = 3.11 for the x = 0.1 sample, and β = 0.41, γ = 1.41 and δ = 4.44 for the x = 0.2 sample. These values are consistent with the prediction of the mean field theory (β = 0.5, γ = 1 and δ = 3) for x = 0.00 and x = 0.10, and with the prediction of the 3D-Heisenberg model (β = 0.365, γ = 1.336 and δ = 4.8) for the x = 0.2 sample.

On the other hand, the cooling efficiency of a magnetic refrigerant is evaluated through the so-called relative cooling power (RCP). The latter corresponds to the amount of heat transferred between the cold and hot sinks in the ideal refrigeration cycle, which is defined as follows:28

 
RCP = −ΔSmax × δTFWHM (5)
where ΔSmax is the maximum of magnetic entropy change and δTFWHM is a full width at half maximum. Fig. 13 presents the RCP dependence on the vacancy content x, under different applied magnetic fields.


image file: c5ra27132k-f13.tif
Fig. 13 Variation of the RCP factor as a function of vacancy content (x) under different magnetic fields.

It is clearly seen that for the different x values, the RCP factor remains almost constant under the same magnetic field. Besides, a refrigerator capable of working in a wide temperature range can be achieved with a series of magnetocaloric materials with significant and similar RCP factors. These materials are combined to form a composite refrigerant working in the temperature range limited by their TC. Lastly, from our materials, a magnetocaloric composite that operates with a considerable efficacy in the temperature range between 275 and 325 K can be synthesized.

For an applied magnetic field of 5 T, the RCP values were found to be 238 J kg−1 for x = 0.1. Table 2 exhibits the comparison between our obtained results and those of other magnetocaloric materials. From the comparison with Gd data and from a technological point of view, our data confirms that the proposed material can be considered as a prospective candidate to be used in a cooling system based on magnetic refrigeration. From an industrial point of view, this material presents beneficial parameters such as low cost, lacking rare-earths, low weight, no corrosion, ease of synthesis, and chemical stability. For all these reasons, the suggested material can be considered as a considerable candidate for magnetic refrigeration.

Table 2 Summary of the magnetocaloric properties of the La0.8K0.2−xxMnO3 (x = 0, 0.1 and 0.2) compounds compared to those of other magnetic materials
Materials TC (K) −ΔSMmax (J kg−1 K−1) RCP (J kg−1) ΔCmaxPCminP (J kg−1 K−1) Reference
Gd 299 4.20 196 29
La0.8Ca0.2MnO3 183 2.23 112.36 6
La0.8Na0.2MnO3 335 2.83 76.91 6
La0.8K0.2MnO3 325 1.64 77.16 10.76/−22.19 This work
La0.8K0.10.1MnO3 300 1.65 95.81 10.03/−21.04 This work
La0.80.2MnO3 275 2.53 80.43 30.45/−37.98 This work


The change in specific heat (ΔCP) linked to a magnetic field variation from 0 to μ0H can be calculated as follows:30

 
image file: c5ra27132k-t6.tif(6)

Fig. 14 shows the temperature dependence of the change in specific heat (ΔCP) for different field variations for the La0.8K0.2−xxMnO3 (x = 0, 0.1 and 0.2) samples. ΔCP is calculated from the ΔSM data using eqn (6). Indeed, the value of ΔCP changes abruptly from negative to positive around the Curie temperature and increases quickly with an increase in temperature. The maximum/minimum values of ΔCPCmaxPCminP) increase with the applied magnetic field. They occur at the temperatures of 318.5/339.5, 292.5/313.5 and 273.5/288.5 for the x = 0.0, 0.1 and 0.2 samples, respectively. Moreover, the values of ΔCPCmaxPCminP) under an applied magnetic field of 2 T are listed in Table 2.


image file: c5ra27132k-f14.tif
Fig. 14 Variation of the specific heat (ΔCP) as a function of temperature under different applied magnetic fields for the La0.8K0.2−xxMnO3 (x = 0, 0.1 and 0.2) compounds.

4. Conclusions

Systematic investigations on the magnetic and the magnetocaloric properties of La0.8K0.2−xxMnO3 (x = 0; 0.1 and 0.2) samples were conducted. A second-order transition for magnetic transition from ferromagnetic to paramagnetic phase is observed. The Curie temperature was found to decrease with an increase in the potassium vacancy content. As a main result from an application point of view, for x = 0.1, the TC is equal to 300 K and the RCP values were found to be equal to 238 J kg−1 under an applied magnetic field of 5 T, which is suitable for potential application in magnetic refrigeration at room temperature.

References

  1. Colossal Magnetoresistive Oxides, ed. Y. Tokura, Gordon and Breach Science Publishers, The Netherlands, 2000 Search PubMed.
  2. M. H. Phan, M. B. Morales, N. S. Bingham, H. Srikanth, C. L. Zhang and S. W. Cheong, Phys. Rev. B: Condens. Matter Mater. Phys., 2010, 81, 094413 CrossRef.
  3. A. Biswas and I. Das, Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 74, 172405 CrossRef.
  4. S. Chandra, A. I. Figueroa, B. Ghosh, M. H. Phan, H. Srikanth and A. K. Raychaudhuri, J. Appl. Phys., 2011, 109, 07D720 CrossRef.
  5. R. Skini, M. Khlifi, M. Wali, E. Dhahri and E. K. Hlil, J. Magn. Magn. Mater., 2014, 324, 217–223 CrossRef.
  6. M. Wali, R. Skini, M. Khlifi, E. Dhahri and E. K. Hlil, Dalton Trans., 2015, 44, 12796–12803 RSC.
  7. R. Skini, A. Omri, M. Khlifi, E. Dhahri and E. K. Hlil, J. Magn. Magn. Mater., 2014, 364, 5–10 CrossRef CAS.
  8. A. Biswas, T. Samanta, S. Banerjee and I. Das, J. Appl. Phys., 2008, 103, 013912 CrossRef.
  9. A. Biswas, T. Samanta, S. Banerjee and I. Das, Appl. Phys. Lett., 2008, 92, 012502 CrossRef.
  10. K. A. Gschneidner Jr, V. K. Pecharsky and A. O. Tsokol, Rep. Prog. Phys., 2005, 68, 1479 CrossRef.
  11. C. R. H. Bahl, D. Velazquez, K. Nielsen, K. Engelbrecht, K. B. Andersen, R. Bulatova and N. Pryds, Appl. Phys. Lett., 2012, 100, 121905 CrossRef.
  12. R. A. Young, The Rietveld Method, Oxford University Press, New York, 1993 Search PubMed.
  13. R. D. Shannon, Acta Crystallogr., Sect. A: Found. Crystallogr., 1976, 32, 751 CrossRef.
  14. R. Skini, M. Khlifi, M. Jemmali, E. Dhahri and E. K. Hlil, Physica B, 2015, 457, 314–319 CrossRef CAS.
  15. M. Khlifi, M. Bejar, O. EL Sadek, E. Dhahri, M. A. Ahmed and E. K. Hlil, J. Alloys Compd., 2011, 509, 7410 CrossRef CAS.
  16. W. Boujelben, A. Cheikh-Rouhou, J. Pierre and J. C. Joubert, Physica B, 2002, 321, 37–44 CrossRef CAS.
  17. G. Popov, J. Goldsmith and M. Greenbaltt, J. Solid State Chem., 2003, 175, 52 CrossRef CAS.
  18. R. Skini, M. Khlifi, E. Dhahri and E. K. Hlil, J. Supercond. Novel Magn., 2014, 27, 247 CrossRef CAS.
  19. S. K. Banerjee, Phys. Lett., 1964, 12, 16 CrossRef.
  20. E. Bruck, J. Phys. D: Appl. Phys., 2005, 38, R381 CrossRef.
  21. A. H. Morrish, The Physical Principles of Magnetism, Wiley, New York, 1965, ch. 3 Search PubMed.
  22. H. Oesterreicher and F. T. Paker, J. Appl. Phys., 1984, 55, 4336 CrossRef.
  23. T. L. Phan, T. A. Ho, P. D. Thang, Q. T. Tran, T. D. Thanh, N. X. Phuc, M. H. Phan, B. T. Huy and S. C. Yu, J. Alloys Compd., 2014, 615, 937–945 CrossRef CAS.
  24. V. Franco and A. Conde, Int. J. Refrig., 2010, 33, 465–473 CrossRef CAS.
  25. S. Ghodhbane, A. Dhahri, N. Dhahri, E. K. Hlil and J. Dhahri, J. Alloys Compd., 2013, 550, 358–364 CrossRef CAS.
  26. A. Rostamnejadi, M. Venkatesan, P. Kameli, H. Salamati and J. M. D. Coey, J. Magn. Magn. Mater., 2011, 323, 2214–2218 CrossRef CAS.
  27. M. Khlifi, A. Tozri, E. Dhahri and E. K. Hlil, J. Magn. Magn. Mater., 2012, 324, 2142 CrossRef CAS.
  28. V. K. Pecharsky, K. A. Gschneider and A. O. Tsokol, Rep. Prog. Phys., 2005, 68, 1479 CrossRef.
  29. V. K. Pecharsky and K. A. Gschneidner Jr, Phys. Rev. Lett., 1997, 78, 4494 CrossRef CAS.
  30. W. Zhong, W. Chen, W. P. Ding, N. Zhang, A. Hu, Y. W. Du and Q. J. Yan, Eur. Phys. J. B, 1998, 3, 169 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.