Next Article in Journal
Mechanical Properties of Metallic Glasses
Next Article in Special Issue
Magnetic Moment of Cu-Modified Ni2MnGa Magnetic Shape Memory Alloys
Previous Article in Journal
Synthesis and Characterization of NanocrystallineMg-7.4%Al Powders Produced by Mechanical Alloying
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Exchange Bias and Inverse Magnetocaloric Effect in Co and Mn Co-Doped Ni2MnGa Shape Memory Alloy

School of Mechanical and Aerospace Engineering, Nanyang Technological University, 50 Nanyang Avenue, Singapore 639798, Singapore
*
Author to whom correspondence should be addressed.
Metals 2013, 3(1), 69-76; https://doi.org/10.3390/met3010069
Submission received: 13 December 2012 / Revised: 8 January 2013 / Accepted: 17 January 2013 / Published: 25 January 2013
(This article belongs to the Special Issue Shape Memory Alloys)

Abstract

:
Exchange bias effect observed in the Ni1.68Co0.32Mn1.20Ga0.80 alloy confirms the coexistence of antiferromagnetic and ferromagnetic phases in the martensite phase. A large inverse magnetocaloric effect has been observed within the martensitic transformation temperature range, which is originated from modified magnetic order through magnetic-field-induced phase transformation from partially antiferromagnetic martensite to ferromagnetic austenite. The magnetic entropy change is 16.2 J kg−1 K−1 at 232 K under ΔH = 60 kOe, with the net refrigerant capacity of 68 J kg−1. These properties indicate Co and Mn co-doped Ni2MnGa alloy is a multifunctional material potentially suitable for magnetic refrigeration and spintronics applications.

1. Introduction

Stoichiometric Ni2MnGa is a full Heusler alloy that undergoes two separate transitions on cooling: (i) a magnetic transition (paramagnetic-ferromagnetic (FM) transition in the austenite) at Tc= 376 K, and (ii) a structural transition (martensitic transformation (MT) from austenite to martensite) at TM= 202 K [1]. In both austenitic (below Curie temperature Tc) and martensitic states, the magnetic couplings are FM, giving rise to a small magnetization difference (ΔM) between austenite and martensite. A large magnetocaloric effect (MCE) has been observed when applying a magnetic field at temperatures close to the MT [2,3,4,5,6]. By varying the composition, the transition temperature can be altered within a broad temperature range. The maximum MCE was obtained when structural and magnetic transition temperatures overlap each other (TM = Tc) [3]. Furthermore, an extrinsic inverse MCE (IMCE, namely, positive ΔSM) phenomenon has also been reported in several Ni-Mn-Ga alloys under application of low magnetic field, arising from the coupling between the martensitic and magnetic domains at the mesoscale [4], i.e., magnetization change due to martensite reorientation associated with the high magnetic anisotropy. However, this process cannot result in a true ΔSM since it is originated from the field dependent change of magnetic order parameter [7]. On the other hand, the intrinsic IMCE has been observed in several other systems around the first order magnetic transition with reduction of magnetization during cooling [8,9,10,11,12]. In this paper, we report the IMCE in Ni1.68Co0.32Mn1.20Ga0.80 alloy, in which structural transformation is accompanied with a large metamagnetic transition below Tc (TM < Tc). The observation of exchange bias (EB) effect confirms the coexistence of AFM and FM phases in the martensitic state.

2. Results and Discussion

Figure 1a shows the temperature dependence of magnetization M(T) measured at H = 100 Oe under zero-field-cooled (ZFC), field-cooled cooling (FCC) and field-cooled heating (FCH) modes, respectively. At high temperatures, in the austenitic state, the sample is paramagnetic and orders ferromagnetically below Tc = 429 K (inset of Figure 1a). However, as the temperature further decreases, a sudden drop in magnetization takes place at about 228 K as a result of the MT and entered a steady state with a lower magnetization. The thermal hysteresis observed between the FCC and FCH processes across the MT confirms the first order nature of this transition. The behavior of the magnetization of martensitic state depends strongly on the thermomagnetic history of the sample. The M(T) curves measured under 20 and 60 kOe fields are similar to that of 100 Oe, showing a large ΔM between austenite and martensite. MT temperatures decrease with increasing magnetic field, e.g., the TM is decreased by about 23 K with the application of 60 kOe field (inset of Figure 1b). This shows that the magnetic field can induce reverse MT from martensite to austenite in the Co and Mn co-doped Ni2MnGa alloy which is further confirmed by the resistivity measurement and shows a giant magnetoresistance effect [13].
As we know, the ΔM of Ni2MnGa across the MT is very small due to the fact that both martensitic and austenitic states are FM. However, in the Mn-doped Ni2MnGa alloy, the extra Mn atoms which occupy the Ga sites reduce the distance between Mn atoms, leading to the magnetic moments partially align antiferromagnetically in both martensite and austenite [14]. On the other hand, the partial substitution of Ni by Co in Mn-doped Ni2MnGa can further change the magnetic ordering in the austenitic state from antiferromagnetic (AFM) to FM [15]. As a result, the Co and Mn co-doping produces a competition between AFM and FM, which is similar to the state in the Co and Fe co-doped Ni2MnGa alloys [12]. Extended X-ray Absorption Fine Structure (EXAFS) results show that the distance of Mn atoms decreases across MT from austenite to martensite [16]. The decrease of the Mn-Mn distance between 4(a) and 4(b) sites (4(a) and 4(b) sites are the Mn and Ga sites in the stoichiometric compound, respectively) may introduce AFM exchange, forming partial AFM in the FM matrix, i.e., coexistence of FM and AFM in martensite in the present alloy. One way to clarify this point is to study the EB effect of the sample, which has been used successfully in NiMn-based alloys [17,18,19,20,21].
Figure 1. (a) zero-field-cooled (ZFC), field-cooled cooling (FCC), and field-cooled heating (FCH) M(T) of the sample measured under H = 100 Oe. The inset shows the temperature dependence of the ac susceptibility (χ(T)) at higher temperatures intended to obtain the Tc. (b) M(T) of the sample under high magnetic fields in FCC and FCH conditions respectively. The inset shows the field dependence of MT starting temperature (TM(H)). The TM temperatures under different fields are indicated by arrows.
Figure 1. (a) zero-field-cooled (ZFC), field-cooled cooling (FCC), and field-cooled heating (FCH) M(T) of the sample measured under H = 100 Oe. The inset shows the temperature dependence of the ac susceptibility (χ(T)) at higher temperatures intended to obtain the Tc. (b) M(T) of the sample under high magnetic fields in FCC and FCH conditions respectively. The inset shows the field dependence of MT starting temperature (TM(H)). The TM temperatures under different fields are indicated by arrows.
Metals 03 00069 g001
It is well known that when a sample with FM/AFM interfaces is cooled under magnetic field, the hysteresis loop is shifted along the field axis (called EB) below blocking temperature (TB) [19,22]. In order to confirm the coexistence of AFM and FM, we measured the hysteresis loops of magnetizations from −20 kOe to 20 kOe at 10 K after ZFC and FCC (H = 10 kOe) from 300 K, respectively. For clear visualization of the loops shift, only the loops form −1 to 1 kOe are shown in Figure 2a. The ZFC loop is symmetrical around zero point, whereas the shift of the FCC loops towards negative field provides evidence on the existence of EB. Moreover, the coercivity (Hc) of FCC loop (99.5 Oe) is higher than that of ZFC loop (61.5 Oe), which is ascribed to the development of the exchange anisotropy after FCC [22]. Figure 2b shows the EB field (HEB) and Hc as a function of temperature. The HEB and Hc are defined as HEB = −(H1 + H2)/2 and Hc = −(H1H2)/2, respectively. Here, H1 and H2 are the left and right coercive fields, respectively. The HEB approximately linearly decreases with increasing temperature in the low temperature regions and gradually disappears around TB. The TB can also be identified from ZFC M(T) curve (Figure 1a). The magnetization increases at TB indicating the AFM can no longer pin the FM domain. With the anisotropy of AFM deceasing with increasing temperature, the FM rotation can drag more AFM spins, giving rise to the increase in Hc; whereas the AFM no longer hinders the FM rotation above TB. As a result, the Hc reaches its maximum value and HEB reduces to zero at higher temperatures. This further confirms that some AFM exchange occurs inside the FM matrix in the martensite.
Figure 2. (a) Hysteresis loops of magnetization at 10 K after ZFC and FCC (H = 10 kOe) from 300 K. (b) EB field (HEB) and coercivity (Hc) as a function of temperature.
Figure 2. (a) Hysteresis loops of magnetization at 10 K after ZFC and FCC (H = 10 kOe) from 300 K. (b) EB field (HEB) and coercivity (Hc) as a function of temperature.
Metals 03 00069 g002
The IMCE can be expected within the MT temperature range in Co and Mn co-doped Ni2MnGa alloy due to the reduction of magnetization during the MT resulting from the formation of partially AFM phase. In order to determine the magnetic entropy change (ΔSM) across the MT, we measured the isothermal magnetization M(H) curves of Ni1.68Co0.32Mn1.20Ga0.80. Magnetic field up to 60 kOe was applied isothermally between 210 and 252 K. In Figure 3, only a portion of the curves within the transformation temperature range is shown for clarity. The typical metamagnetic behavior due to the field-induced reverse MT is observed between 222 and 238 K. Above 240 K, the magnetization curves exhibit characteristics of a ferromagnet. Similarly, the metamagnetic behavior disappears below 220 K, giving rise to a FM behavior with lower saturation magnetization than that of high-temperature phase. The ΔSM can be calculated from magnetization isotherms using the Maxwell relation:
Metals 03 00069 i001
For magnetization measurements made at discrete temperatures and field intervals, the ΔSM can be approximated by:
Metals 03 00069 i002
For a truly discontinuous first order phase transition, since the [∂M/∂T]H does not exist theoretically, Equations 1 and 2 cannot be used to calculate the ΔSM. But in reality, the first order phase transition is not ideal and the change of magnetization occurs over a few Kelvin temperature range, as a result, the [∂M/∂T]H can be measured experimentally [23].
Figure 3. Isothermal M(H) of Ni1.68Co0.32Mn1.20Ga0.80 around the MT temperature range. The arrows indicate the direction of change of the applied field, and the shaded area represents the hysteretic loss for T = 228 K isotherm. The inset shows the hysteretic loss as a function of temperature.
Figure 3. Isothermal M(H) of Ni1.68Co0.32Mn1.20Ga0.80 around the MT temperature range. The arrows indicate the direction of change of the applied field, and the shaded area represents the hysteretic loss for T = 228 K isotherm. The inset shows the hysteretic loss as a function of temperature.
Metals 03 00069 g003
The temperature dependencies of the ΔSM under different magnetic fields are shown in Figure 4. The most striking features of the ΔSM are the large and positive entropy change within the MT temperature range. The positive sign of ΔSM exists even under high magnetic field, showing the mechanism involved is different from that of Co-free Ni-Mn-Ga alloys which is related to the coupling between the martensitic and magnetic domains [4]. The modification of magnetic order through magnetic-field-induced reverse MT from partially AFM martensite to FM austenite is responsible for the IMCE in the present alloy. The ΔSMmax increases with ΔH and shows saturation for lager ΔH (inset of Figure 4). The saturation of the ΔSMmax as a function of the field and the broadening of ΔSM (T) with increasing field are characteristic features of first order magnetic transition [7]. The ΔSM is 16.2 J Kg−1 K−1 for ΔH = 60 kOe field. Compared with other IMCE materials, the ΔSM for ΔH = 60 kOe in the present alloy is comparable to the value observed in Ni-Mn-Sn alloys [10], and is two times larger than the value observed in Mn1.95Cr0.05Sb [8]. One important parameter to evaluate the magnetocaloric properties of a given material is the refrigeration capacity (RC), giving the amount of heat that can be transferred in one thermodynamic cycle. The RC is usually calculated by integrating the ΔSM (T) curve over the full width at half maximum (the shaded area in Figure 4), which yields a value of about 125 J kg−1 for ΔH = 60 kOe. For a first order magnetic phase transition, the hysteretic loss is another key parameter for evaluating the MCE around transition, which is equal to the area between the increasing and decreasing magnetic field segments of the magnetization curves (the shaded area in Figure 3). The hysteretic loss as a function of the temperature is shown in the inset of Figure 2, giving a maximum at almost the same temperature as the maximum of the ΔSM (T) curve.The average hysteretic loss over the same temperature range as that of RC is about 57 J kg−1. Thus, the net RC is 68 J kg−1 obtained by subtracting the average hysteretic loss from the RC. This value is comparable to the maximum value obtained by Cu doping in Ni2MnGa alloys, but with opposite sign in ΔSM [6]. Note that the high value of the ΔSM alone is not enough to evaluate the MCE of a material. Another important parameter for MCE is adiabatic temperature change (ΔTad), which can be measured directly in adiabatic system with application of a magnetic field [7,11,24] or be estimated from the magnetization and specific heat measurements [25]. Thus, some further studies, such as the measurement of ΔTad, are needed to further evaluate the IMCE in the present alloy.
Figure 4. Temperature dependence of the ΔSM under different magnetic field variations (ΔH). The shaded area indicates the RC at ΔH = 60 kOe. The inset shows the field dependence of maximum ΔSM within the MT temperature range at T = 232 K.
Figure 4. Temperature dependence of the ΔSM under different magnetic field variations (ΔH). The shaded area indicates the RC at ΔH = 60 kOe. The inset shows the field dependence of maximum ΔSM within the MT temperature range at T = 232 K.
Metals 03 00069 g004

3. Experimental Section

A polycrystalline Ni1.68Co0.32Mn1.20Ga0.80 alloy was prepared with an arc-melting technique under argon atmosphere using Ni, Co, Mn, and Ga of 99.9% purity. The sample was remelted several times and subsequently annealed at 1000 °C for 24 h before being slowly cooled to room temperature to ensure homogeneity. The composition was determined by energy dispersive X-ray analysis (EDX) (Figure 5). The ac susceptibility measurement was performed to determine the Tc. The magnetization measurements were performed on a physical properties measurement system (Quantum Design) platform with a vibrating sample magnetometer module.
Figure 5. EDX spectrum of Ni1.68Co0.32Mn1.20Ga0.80. The inset shows the scanning electron microscopic image. The little rectangle marks the position at which the EDX spectrum was taken.
Figure 5. EDX spectrum of Ni1.68Co0.32Mn1.20Ga0.80. The inset shows the scanning electron microscopic image. The little rectangle marks the position at which the EDX spectrum was taken.
Metals 03 00069 g005

4. Conclusions

In conclusion, a large positive ΔSM has been found in Mn and Co co-doped Ni2MnGa alloy within the MT temperature range. The partial substitution of Ni by Co in Mn-doped Ni2MnGa, a typical partially AFM material, modifies the exchange interaction among Mn atoms and results in a change from AFM to FM in the austenite. The modification of magnetic order from partially AFM martensite to FM austenite through magnetic-field-induced reverse MT is responsible for the IMCE in the present alloy. The observation of EB effect confirms that some AFM exchange exists inside the FM matrix of martensite. The IMCE with a high net RC indicates that the present alloy may be a promising candidate for magnetic refrigeration.

Conflict of Interest

The authors declare no conflict of interest.

References

  1. Vasil’ev, A.N.; Bozhko, A.D.; Khovailo, V.V.; Dikshtein, I.E.; Shavrov, V.G.; Buchelnikov, V.D.; Matsumoto, M.; Suzuki, S.; Takagi, T.; Tani, J. Structural and magnetic phase transitions in shape-memory alloys Ni2 + xMn1 − xGa. Phys. Rev. B 1999, 59, 1113–1120. [Google Scholar]
  2. Hu, F.X.; Shen, B.G.; Sun, J.R. Magnetic entropy change in Ni51.5Mn22.7Ga25.8 alloy. Appl. Phys. Lett. 2000, 76, 3460–3463. [Google Scholar] [CrossRef]
  3. Pareti, L.; Solzi, M.; Albertini, F.; Paoluzi, A. Giant entropy change at the co-occurrence of structural and magnetic transitions in the Ni2.19Mn0.81Ga Heusler alloy. Eur. Phys. J. B 2003, 32, 303–307. [Google Scholar] [CrossRef]
  4. Marcos, J.; Mañosa, L.; Planes, A.; Casanova, F.; Batlle, X.; Labarta, A. Multiscale origin of the magnetocaloric effect in Ni-Mn-Ga shape-memory alloys. Phys. Rev. B 2003, 68, 094401:1–094404:6. [Google Scholar]
  5. Pasquale, M.; Sasso, C.P.; Lewis, L.H.; Giudici, L.; Lograsso, T.; Schlagel, D. Magnetostructural transition and magnetocaloric effect in Ni55Mn20Ga25 single crystals. Phys. Rev. B 2005, 72. [Google Scholar] [CrossRef]
  6. Stadler, S.; Khan, M.; Mitchell, J.; Ali, N.; Gomes, A.M.; Dubenko, I.; Takeuchi, A.Y.; Guimaraes, A.P. Magnetocaloric properties of Ni2Mn1 − xCuxGa. Appl. Phys. Lett. 2006, 88, 192511:1–192511:3. [Google Scholar]
  7. Giguère, A.; Foldeaki, M.; Ravi Gopal, B.; Chahine, R.; Bose, T.K.; Frydman, A.; Barclay, J.A. Direct measurement of the “giant” adiabatic temperature change in Gd5Si2Ge2. Phys. Rev. Lett. 1999, 83, 2262–2265. [Google Scholar] [CrossRef]
  8. Tegus, O.; Bruck, E.; Zhang, L.; Dagula; Buschow, K.H.J.; de Boer, F.R. Magnetic-phase transitions and magnetocaloric effects. Physica B 2002, 319, 174–192. [Google Scholar] [CrossRef]
  9. Han, Z.D.; Wang, D.H.; Zhang, C.L.; Tang, S.L.; Gu, B.X.; Du, Y.W. Large magnetic entropy changes in the Ni45.4Mn41.5In13.1 ferromagnetic shape memory alloy. Appl. Phys. Lett. 2006, 89, 182507:1–182507:3. [Google Scholar]
  10. Krenke, T.; Duman, E.; 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]
  11. Moya, X.; Mañosa, L.; Planes, A.; Aksoy, S.; Acet, M.; Wassermann, E.F.; Krenke, T. Cooling and heating by adiabatic magnetization in the Ni50Mn34In16 magnetic shape-memory alloy. Phys. Rev. B 2007, 75, 184412:1–184412:5. [Google Scholar]
  12. Pathak, A.K.; Dubenko, I.; Karaca, H.E.; Stadler, S.; Ali, N. Large inverse magnetic entropy changes and magnetoresistance in the vicinity of a field-induced martensitic transformation in Ni50 − xCoxMn32 − yFeyGa18. Appl. Phys. Lett. 2010, 97, 062505:1–062505:3. [Google Scholar]
  13. Wang, B.M.; Ren, P.; Liu, Y.; Wang, L. Enhanced magnetoresistance through magnetic-field-induced phase transition in Ni2MnGa co-doped with Co and Mn. J. Magn. Magn. Mater. 2010, 322, 715–717. [Google Scholar] [CrossRef]
  14. Enkovaara, J.; Heczko, O.; Ayuela, A.; Nieminen, R.M. Coexistence of ferromagnetic and antiferromagnetic order in Mn-doped Ni2MnGa. Phys. Rev. B 2003, 67, 212405:1–212405:4. [Google Scholar]
  15. Yu, S.Y.; Cao, Z.X.; Ma, L.; Liu, G.D.; Chen, J.L.; Wu, G.H.; Zhang, B.; Zhang, X.X. Realization of magnetic field-induced reversible martensitic transformation in NiCoMnGa alloys. Appl. Phys. Lett. 2007, 91, 102507:1–102507:3. [Google Scholar]
  16. Bhobe, P.A.; Priolkar, K.R.; Sarode, P.R. Factors influencing the martensitic transformation in Ni50Mn35Sn15: An EXAFS study. J. Phys. Condens. Mater. 2008, 20. [Google Scholar] [CrossRef]
  17. Khan, M.; Dubenko, I.; Stadler, S.; Ali, N. Exchange bias in bulk Mn rich Ni-Mn-Sn Heusler alloys. J. Appl. Phys. 2007, 102, 113914:1–113914:4. [Google Scholar]
  18. Pathak, A.K.; Khan, M.; Gautam, B.R.; Stadler, S.; Dubenko, I.; Ali, N. Exchange bias in bulk Ni-Mn-In-based Heusler alloys. J. Magn. Magn. Mater. 2009, 321, 963–965. [Google Scholar] [CrossRef]
  19. Wang, B.M.; Liu, Y.; Wang, L.; Huang, S.L.; Zhao, Y.; Yang, Y.; Zhang, H. Exchange bias and its training effect in the martensitic state of bulk polycrystalline Ni49.5Mn34.5In16. J. Appl. Phys. 2008, 104, 043916:1–043916:4. [Google Scholar]
  20. Wang, B.M.; Liu, Y.; Xia, B.; Ren, P.; Wang, L. Large exchange bias obtainable through zero-field cooling from an unmagnetized state in Ni-Mn-Sn alloys. J. Appl. Phys. 2012, 111, 043912:1–043912:7. [Google Scholar]
  21. Wang, B.M.; Liu, Y.; Ren, P.; Xia, B.; Ruan, K.B.; Yi, J.B.; Ding, J.; Li, X.G.; Wang, L. Large exchange bias after zero-field cooling from an unmagnetized state. Phys. Rev. Lett. 2011, 106, 077203:1–077203:4. [Google Scholar]
  22. Nogués, J.; Schuller, I.K. Exchange bias. J. Magn. Magn. Mater. 1999, 192, 203–232. [Google Scholar] [CrossRef]
  23. Gschneidner, K.A., Jr.; Pecharsky, V.K.; Tsokol, A.O. Recent developments in magnetocaloric materials. Rep. Prog. Phys. 2005, 68, 10–1088. [Google Scholar]
  24. Khovaylo, V.V.; Skokov, K.P.; Koshkid’ko, Y.S.; Koledov, V.V.; Shavrov, V.G.; Buchelnikov, V.D.; Taskaev, S.V.; Miki, H.; Takagi, T.; Vasiliev, A.N. Adiabatic temperature change at first-order magnetic phase transitions: Ni2.19Mn0.81Ga as a case study. Phys. Rev. B 2008, 78, 060403:1–060403:4. [Google Scholar]
  25. Pecharsky, V.K.; Gschneidner, K.A., Jr. Giant magnetocaloric effect in Gd5(Si2Ge2). Phys. Rev. Lett. 1997, 78, 4494–4497. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Wang, B.; Liu, Y. Exchange Bias and Inverse Magnetocaloric Effect in Co and Mn Co-Doped Ni2MnGa Shape Memory Alloy. Metals 2013, 3, 69-76. https://doi.org/10.3390/met3010069

AMA Style

Wang B, Liu Y. Exchange Bias and Inverse Magnetocaloric Effect in Co and Mn Co-Doped Ni2MnGa Shape Memory Alloy. Metals. 2013; 3(1):69-76. https://doi.org/10.3390/met3010069

Chicago/Turabian Style

Wang, Baomin, and Yong Liu. 2013. "Exchange Bias and Inverse Magnetocaloric Effect in Co and Mn Co-Doped Ni2MnGa Shape Memory Alloy" Metals 3, no. 1: 69-76. https://doi.org/10.3390/met3010069

Article Metrics

Back to TopTop