Electrical, magnetic, and Hall properties of AlxCoCrFeNi high-entropy alloys

Yih-Farn Kao,Swe-Kai Chen,Ting-Jie Chen,Po-Chou Chu,Jien-Wei Yeh,Su-Jien Lin
DOI: https://doi.org/10.1016/j.jallcom.2010.10.210
IF: 6.2
2011-01-01
Journal of Alloys and Compounds
Abstract:Research highlights ▶ Values of ρ 0 at4.2 K for multi-principal-element alloys C- x , H- x , and D- x are 111.06–196.49, 93.78–162.77, and 120.48–162.05 μΩ cm, respectively. With the same level of carrier density, they are larger than those of conventional alloys, but smaller than those of amorphous ones. RRR values range from 1.03 to 1.27. That is, the electronic transportation in the studied alloy system is dominated by the defects of lattice. These facts imply that the degree of defects in a single lattice for these high-entropy alloys is lower than for amorphous ones, but larger than conventional ones. The TCR value of H-2.00 listed is small (82.5 ppm/K). From the ρ 0 ( T ) fitting, it is demonstrated that the Kondo-like behavior exists for some compositions. ▶ BCC phase has high M s value at low temperature in this alloy system. At 300 K, alloys H-0.50, H-1.25, and H-2.00 are ferromagnetic, while alloys H-0.00, H-0.25, and H-0.75 are paramagnetic. Al and AlNi-rich phases decreases ferromagnetism for single FCC and single BCC alloys H- x , respectively. The spin glass behavior is observed for some compositions of this study. ▶ Alloys H- x are of the hole-like carrier type. The stronger the ferromagnetic alloys the higher the % AHE , i.e., the anomalous Hall effect (AHE) is magnetic dependent. Carrier density for H- x (10 22–23 cm −3 ) is similar to that for conventional alloys. Alloys H- x have small carrier mobility (0.40–2.61 cm 2 V −1 s −1 ). It is associated with the heavier lattice defects of HEAs as compared to the conventional alloys with the same carrier density. Abstract This investigation explores the electrical and magnetic properties of as-cast, -homogenized, and -deformed Al x CoCrFeNi (C- x , H- x , and D- x , respectively) alloys at various temperatures from 4.2 to 300 K. Experimental results reveal that carrier density of the alloys is of 10 22–23 cm −3 . H- x has a carrier mobility of 0.40–2.61 cm 2 V −1 s −1 . The residual electrical resistivity of the alloys varies from 100 to 220 μΩ cm. The temperature coefficient of resistivity (TCR) of H-2.00 is small (82.5 ppm/K). Therefore, defects in the lattice dominate electrical transportation. Some compositions exhibit Kondo-like behavior. At 300 K, H-0.50, H-1.25, and H-2.00 are ferromagnetic, while H-0.00, H-0.25, and H-0.75 are paramagnetic. Al and AlNi-rich phases reduce the ferromagnetism of single FCC and single BCC H- x , respectively. Spin glass behavior of some compositions is also observed. Alloys H- x are of the hole-like carrier type, and ferromagnetic H- x exhibits an anomalous Hall effect (AHE). Keywords Bulk Al x CoCrFeNi alloys Cast Homogenization Plastic deformation Melt-spinning Electrical resistivity Hall effect Carrier density Carrier mobility Magnetic property Kondo-like effect Lattice defects 1 Introduction The first study on high-entropy alloys (HEAs) was published in 1996 [1] . Today, research into HEAs addresses their mechanical, anticorrosion, hydrogen storage [2] , and thermophysical [3] properties, among others. Relevant results demonstrate that HEAs have simple microstructures in the form of a solid solution of multiple elements [4–6] , a favorable capacity to form nano-scale precipitates [5,7] , high thermal stability [8] , superior extensive or compressive properties [9] , extremely high hardness [10] , excellent anticorrosive properties [11,12] , and special thermophysical and magnetic properties [3] . The crystallinity of HEAs is commonly simple, even though they are comprised more than five elements. The simple crystal lattices exhibit both the individual characteristics of their constituents and collective characteristics. For example, they exhibit the collective mechanical and thermal properties of a solid solution, but the anticorrosion performance of its individual constituent elements. Understanding the physical properties, including electrical, magnetic, and thermal properties, of HEAs can help in understanding their lattice. The magnetic property of CoCrCuFeNiTi x alloys [13] has been studied. The FCC solid solutions that comprise in CoCrCuFeNi and CoCrCuFeNiTi 0.5 alloys exhibit typical paramagnetism, whereas CoCrCuFeNiTi 0.8 and CoCrCuFeNiTi alloys exhibit superparamagnetism, which is attributable to the embedding of nanoparticle assemblies in the amorphous phase with the addition of Ti. Al x CoCrFeNi alloys have been widely studied to elucidate their microstructural and mechanical [4] , anticorrosive [11] , and thermally expansive [3] properties. However, their electrical and magnetic properties are still not fully understood. This investigation, which extends another, [4] , aims to study the electrical and magnetic properties of Al x CoCrFeNi alloys by measuring resistivity, magnetization, and the Hall effect. The values of the relevant parameters obtained using these three methods will be compared with each other, to provide insight into the physical properties of Al x CoCrFeNi alloys. 2 Experimental details A total of 40–50 g of Al, Co, Cr, Fe, and Ni with purities of greater than 99.5% was used to prepare Al x CoCrFeNi (0 ≤ x ≤ 2) alloys using a vacuum arc-remelter. After the alloys were re-melted, they were turned over and the process was repeated at least three times to ensure that the alloys were completely mixed. As-cast alloys were heated to 1100 °C at 20 °C per min and held at that temperature for 24 h. The samples were then quenched in water. These heated and quenched samples are called as-homogenized ones herein, except where otherwise specified. The as-homogenized samples were cold-deformed by a DBR250 two-high rolling mill to reduce their thickness by 75%. Cracks appeared in the samples with molar ratios of Al of more than 0.875 ( x > 0.875) because they had high hardness. Therefore, the as-deformed samples were those with 0 ≤ x ≤ 0.875 in this investigation. Here, C- x , H- x , and D- x represent as-cast, -homogenized, and -deformed Al x CoCrFeNi (0 ≤ x ≤ 2) alloys, respectively. Melt-spun samples with x = 0.25, M-0.25, were also prepared and their resistivity was measured and compared. The samples were cut using a diamond cutter into pieces with a thickness of 2 mm, and then ground to a thickness of less than 500 μm, using smaller-number sandpaper, to enable their resistivity to be measured. An EG & G Model 5210 Dual Phase Lock-in Amplifier was used to measure the resistance by the four-point probe method. To measure the resistivity of the samples at low temperatures, the samples were gradually sunk into liquid helium in a Dewar flask, and their resistance was continuously measured between 300 and 4.2 K. A superconducting quantum interference device magnetometer MPMS5 (SQUID), from American Quantum Design, was used to measure the magnetic properties of the samples, including the hysteresis loop ( M – H curve) and the magnetization vs. temperature curve ( M – T curve). The two states in which the M – T curves were measured were the zero-field cooling (ZFC) and field cooling (FC) ones. The hysteresis loops were measured at 300, 50, and 5 K. The M – T curves of 5–150 K were measured in a magnetic field of 100 Oe. The manipulating current that was used to measure the Hall effect was 10 mA; the applied magnetic field was up to ±9 T, and the temperature range was 5–300 K. 3 Results and discussion 3.1 Microstructure of alloys C- x , H- x , D- x , 900 °C-homogenized H-0.25, and M-0.25 As shown in Fig. 1 , the structures of the alloys C- x , H- x , and D- x , (with increasing amount of Al, given by x ) are single FCC, duplex FCC–BCC, and single BCC, respectively. Al is a BCC stabilizer in Al x CoCrFeNi alloys. The ranges of x in which the alloys C- x , H- x , and D- x are in a duplex phase are 0.45 ≤ x ≤ 0.88, 0.30 ≤ x ≤ 1.17, and 0.30 ≤ x ≤ 0.875, respectively [4] . Alloys 900 °C-homogenized-H-0.25 and M-0.25 comprise duplex FCC-AlNi-rich precipitate BCC and single FCC, respectively. 3.2 Electrical properties 3.2.1 Highly concentrated point defects in lattice of HEAs Fig. 2 (a)–(c) presents resistivity data for alloys C- x , H- x ( x = 0–2.00), and D- x ′ ( x ′ = 0–0.875), respectively, from 4.2 to 300 K. Table 1 lists the resistivity at 4.2 K ( ρ 0 ), resistivity at 300 K ( ρ 300 ), residual resistivity ratio (RRR), defined as ρ 0 / ρ 300 , and temperature coefficient of resistivity (TCR) defined as ( ρ 300 − ρ 150 )/[ ρ 150 (300 − 150)]. Studies [4,14] have shown that Al is a strong BCC former and can promote atomic bonding in high-entropy alloys (HEAs). Therefore, the mechanical property [15] , wear behavior [16] , and anti-corrosion ability [17] are closely related to the Al content. However, no linear relationship between resistivity and x is observed in Fig. 2 (a)–(c), indicating that the microstructure is not the only factor that affects the electrical resistivity of this alloy system. The ranges of the ρ 0 values of alloys C- x , H- x , and D- x are 111.06–196.49, 93.78–162.77, and 120.48–162.05 μΩ cm, respectively. The ρ 0 values of amorphous AlCoCrCu 0.5 FeNi [18] and Nd–Fe–B [19] alloys are 289 and 250 μΩ cm, respectively; the ρ 0 values of crystallized ternary Ni 75 Fe 17 Cr 8 , Ni 75 Fe 13 Cr 12 , and Ni 68 Fe 17.5 Cr 14.5 alloys are 58.1, 89.6, and 92.4 μΩ cm, respectively [20] . The ρ 0 values of the quinary alloys of interest herein exceed those of the aforementioned traditional crystalline alloys, but are less than those of amorphous alloys. Electron transportation is usually associated with the ‘orderliness’ of a lattice. The above discussion suggests that the orderliness of the lattices of the studied alloys is between that of an ordered structure (traditional crystalline alloys) and that of a disordered structure (amorphous alloy). HEAs exhibit long-range order but seemingly less degree short-range order in the sense of a single lattice. With respect to electronic conduction, quinary HEAs are metallic (see below) but have highly-concentrated point defects, resulting from multi-principal elements in a pseudo-unitary lattice (PUL) cell. The RRR values range from 1.03 to 1.27, as indicated in Table 1 . Accordingly, electronic transportation in the alloys of interest is dominated by the defects from 4.2 to 300 K, the defects in the lattice of alloys dominate electronic transportation therein. Additionally, the TCR value of H- x in Table 1 is low (82.5 ppm/K), because of the lesser orderliness of each lattice. 3.2.2 Resistivity of alloys FCC C-0.25, H-0.25, and D-0.25 Fig. 2 (d) plots data on the resistivity of alloys C-0.25, H-0.25 (900 and 1100 °C), D-0.25 (50% and 75% reductions) and M-0.25 from 4.2 to 300 K. Fig. 3 reveals that the ρ 0 values increase in the order 1100 °C-homogenized H-0.25, C-0.25, 50% reduction-D-0.25, 75% reduction-D-0.25, M-0.25, and 900 °C-homogenized H-0.25. Greater deformation or a higher quenching/cooling rate corresponds to more introduced defects, and higher ρ 0 values. The 900 °C-homogenized H-0.25 has a relatively high ρ 0 . This is associated with a small amount of BCC AlNi-rich precipitate [21] . Additionally, the ρ 0 of D-0.25 after 75% cold deformation slightly exceeds that of D-0.25 after 50% cold deformation. 3.2.3 Kondo-like behavior The dependence of resistivity of temperature is not simple, because it usually involves several effects over several temperature ranges. Conduction electrons usually interact with phonons at high temperatures, such as 300–1000 K, and the dependence in this temperature range is ρ ∝ T [20,22,23] . Conduction electrons usually interact with magnetic atoms in the alloys at intermediate temperatures, such as 100–300 K, and ρ ∝ T 2 . When the temperature declines to 0–50 K, phonon ( T 3 ), magnetic ( T 2 ), and electron–electron interaction ( T 1/2 ) effects occur simultaneously [20] . Hence, suitable temperature ranges must be carefully selected to study the influence of temperature on the effects of all of the above phenomena. Experiments have shown two typical features of electrical conductivity at low temperatures. They are residual resistivity (equal to ρ 0, as described in Section 3.2 .) and the Kondo effect [24] . Residual resistivity is the resistivity that is caused by the scattering of electron waves by static defects that disturb the periodicity of the lattice [22] . According to Matthiessen's rule, the net resistivity is given by ρ = ρ L + ρ i , where ρ L represents the resistivity that is caused by the thermal phonons, and ρ i is the residual resistivity. Therefore, ρ i (0) is the extrapolated resistivity to 0 K because ρ L is commonly independent of the number of defects when the defect concentration is weak, and is usually proportional to the temperature. Therefore, ρ L vanishes as T approaches 0 K. The Kondo effect, however, vanishes at high temperatures, but localized spin fluctuation and RKKY interaction [22] increase resistivity at low temperatures. The Kondo effect commonly arises in dilute alloy systems. Although HEAs are not dilute, they exhibit a Kondo-like behavior, which will be discussed in the following section. In Fig. 2 (a)–(c), two formulae are applied to fit ρ ( T ). At low temperatures (4.2–66 K), ρ = ρ 0 + A ln( T ) + BT 2 + CT 3 is used to model Kondo-like ( A ), magnetic ( B ), and low-temperature phonon ( C ) effects. At intermediate temperatures (100–300 K), the fitting formula, ρ = ρ 0 + BT 2 + DT , where B and D are the coefficients of magnetic and phonon effects, respectively, is utilized. The Kondo-like effect is usually considered [25] to be proportional to lnT at low temperatures, such that Δ ρ / ρ min ∝ ln T . Tables 2 and 3 display the results of fitting. Alloys C- x ( x = 0.25–0.50, and 2.00), H- x ( x = 0.25-2), and D-0.375 exhibit Kondo-like behavior. At low temperatures, B is clearly one to two orders of magnitude higher than at intermediate temperatures. A is always negative in Kondo-like samples, and larger absolute values of A are associated with a stronger Kondo-like effect, especially in H-2.00. The magnetic contribution, T 2 , competes with the Kondo-like effect. When the magnetic contribution greatly exceeds the Kondo-like effect, resistivity remains constant with a change in temperature and levels off at low temperatures. The Kondo-like and the magnetic effects in H- x slightly increase with temperature at low temperatures, but the magnetic and phonon effects decrease with increasing temperature at intermediate temperatures . Evaluating the phonon effect at low temperatures, T 3 , is difficult, because it is extremely small and frequently mixed with a magnetic effect, T 2 [20] . 3.3 Magnetic properties of alloys H- x Fig. 4 (a)–(f) plot SQUID data for alloys H- x , respectively, at 5, 50, and 300 K. Table 4 presents the related magnetic parameters of alloys H- x . In the next two sections, most of the factors that are related to the magnetic property of alloys H- x , are associated with (i) the crystal structure and (ii) the Al or AlNi-rich phase. First, the results that are explained by the crystal structure are discussed. Table 4 reveals that all H- x alloys are ferromagnetic at low temperatures (5 and 50 K). Additionally, the saturation magnetization ( M s ) of alloys H-1.25 and H-2.00 exceeds that of alloys H-0 and H-0.25, indicating that at low temperatures, the BCC phase has a higher M s value than the FCC phase. At room temperature (300 K), alloys H-0.50, H-1.25, and H-2.00 remain ferromagnetic, while alloys H-0, H-0.25, and H-0.75 are paramagnetic. The estimated Curie temperatures ( T C ) are taken from the curves of the reciprocal of magnetic susceptibility ( χ −1 ) against temperature. Table 4 shows no regular relationship between T C and x for alloys H-0, H-0.25, and H-0.50. Next, M s , to which both BCC and FCC phases contribute, is estimated when 0.5 ≤ x ≤ 0.75. The M s of the two phases at 5 K can be calculated numerically as follows: (1) M s = X FCC M s , FCC + X BCC M s , BCC where X FCC , M s,FCC , X BCC , and M s,BCC are the volume fraction of FCC, the saturation magnetization of FCC, the volume fraction of BCC, and the saturation magnetization of BCC, respectively. However, X FCC and X BCC can be obtained from the fitting of hardness [4] . Therefore, the above equation can be rewritten as, (2) 140.11 = 0.7551 M s , FCC + 0.2449 M s , BCC   ( for   H- 0.50 ) (3) 119.93 = 0.5765 M s , FCC + 0.4235 M s , BCC   ( for   H- 0.75 ) Accordingly, M s,FCC = 167.78 (emu cm −3 ) and M s,BCC = 54.79 (emu cm −3 ). The reason for the result M s,BCC < M s,FCC is ascribed to the existence of Al or AlNi-rich phase as discussed further in the following paragraph. The results for which the Al or AlNi-rich phase are responsible is discussed. The M s value declines as x increases in the ranges 0 ≤ x ≤ 0.25 and 1.25 ≤ x ≤ 2.00, as indicated in Table 4 . The M s and magnetic susceptibility ( χ −1 ) values of alloy H-0 are smaller than those of alloy H-0.25, implying that Al reduces the ferromagnetism of the FCC phase alloys (0 ≤ x ≤ 0.25). According to a study of Alnico magnets [26] , the AlNi-rich phase exhibits weak ferromagnetism. In one study [4] , H-2.00 contained an AlNi-rich ordered BCC phase. Hence, the fact that the M s value of H-2.00 is smaller than that of H-1.25, which has no AlNi-rich ordered BCC phase, is unsurprising. In summary, Al reduces the ferromagnetism of single-FCC and single-BCC alloys H- x , by forming Al and AlNi-rich phases, respectively. In Fig. 5 (a)–(b) and (d)–(f), the maximum magnetization can be observed at 5–40 K. This reentrant spin glass behavior occurs at the freezing transition point ( T f ) at which electronic spins are frozen. 3.4 Hall effect of alloys H- x The Hall effect measurements are made in alloys H-0.25 (FCC), H-0.50 (duplex), H-0.75 (duplex), H-1.00 (duplex), and H-1.25 (BCC). Fig. 6 (a)–(b) present data on the Hall resistivity ( ρ H ) of the H- x samples at 5 and 300 K, respectively. Table 5 presents the related parameters. Both the ordinary Hall effect (OHE) and the anomalous Hall effect (AHE) can be observed because of the ferromagnetism of the samples. Fig. 6 plots Hall resistivity ( ρ H ) against magnetic field at 5 and 300 K, and reveals an obvious covariation of AHE and OHE. For alloys H-0.25 and H-0.75, at temperatures above the Curie temperature (300 K), only OHE occurs, while at temperatures below the Curie temperature (5 K), AHE clearly occurs. For H-1.25, the Curie temperature exceeds 300 K, and the curve at 300 K is shifted from that of 5 K since the magnetization is lower at 300 K. For AHE, the Hall resistivity ρ H can be expressed as, (4) ρ H = R 0 B + 4 π M s R s where R 0 is the ordinary Hall coefficient; B is the magnetic induction; M s is the saturation magnetization, and R s is the anomalous Hall coefficient. R 0 B is the Lorentz force that acts on the conduction electrons, and the second term is attributable to the spin-orbit interaction [27–29] in a ferromagnet. The values of R s M s and R 0 are obtained from the intercept and the slope, respectively, and the related fitting results are listed in Table 6 . Notably, the M s values of the samples are obtained from the SQUID data in Section 3.3 . Only the ferromagnetic samples yield R s values in Table 6 , revealing that AHE dominates in ferromagnetic materials. Additionally, values of R s at 5 K are close to those at 300 K, and this result shows that R s is almost independent of temperature. Graphical analysis yields % of OHE and % of AHE, % OHE and % AHE , can be obtained. The relevant equations are as follows: (5) % OHE = R 0 B / ( R 0 B + 4 π M s R s ) (6) % AHE = 1 − % OHE Table 6 presents % OHE and % AHE at 9 T. Clearly, % OHE exceeds % AHE at 300 K because H- x exhibits weaker ferromagnetism at 300 K than at 5 K. The % AHE at 5 K is approximately equal to the saturation magnetization, M s , at 5 K in Table 6 . Samples with higher M s , especially H-1.25, have higher % AHE , indicating that more strongly ferromagnetic samples have higher % AHE , and so % AHE depends on the magnetism. The results in Fig. 6 support this conclusion. The Hall resistivity of alloys H- x is positive, as shown in Fig. 6 . These alloys are similar to Glassy TE-TL alloys [30] . The carriers in these samples are hole-like, although in some cases, the carriers are electron-like [31] . Based on ρ H and ρ (the resistivity data in Section 3.2 .), the related parameters of the Hall effect, presented in Table 5 , can be obtained. The carrier density in a high-entropy alloy system H- x (10 22–23 cm −3 ) is similar to that in conventional alloys. However, the H- x high-entropy alloy system has lower carrier mobility (0.40–2.61 cm 2 V −1 s −1 ) than conventional alloys with the same carrier density, because HEAs have heavier lattice defects than conventional alloys. This explanation warrants further study. 4 Conclusions The values of ρ 0 at 4⋅2 K for alloys C- x , H- x , and D- x are 111.06–196.49, 93.78–162.77, and 120.48–162.05 μΩ cm, respectively. They are larger than those of conventional alloys with the same carrier density, but smaller than those of amorphous alloys. The RRR values range from 1.03 to 1.27. Therefore, the defects in the lattice dominate the electronic transportation in the studied alloy system. These facts imply that the density of defects in a single lattice in these high-entropy alloys is less than in amorphous alloys, but greater than in conventional alloys. The presented TCR value of H-2.00 is small (82.5 ppm/K). The fitted ρ ( T ) reveals that some compositions exhibit Kondo-like behavior. The single BCC structure in the alloy system has a high M s value at low temperatures. At 300 K, alloys H-0.50, H-1.25, and H-2.00 are ferromagnetic, while alloys H-0.00, H-0.25, and H-0.75 are paramagnetic. Al and AlNi-rich phases reduce the ferromagnetism of single FCC and single BCC H- x alloys, respectively. Some of the compositions of interest in this investigation exhibit spin glass behavior. Alloys H- x are of the hole-like carrier type. More strongly ferromagnetic alloys have higher % AHE , and an anomalous Hall effect (AHE) that depends more strongly on magnetism. The carrier density of H- x (10 22–23 cm −3 ) is similar to that of conventional alloys. Alloys H- x have a low carrier mobility (0.40–2.61 cm 2 V −1 s −1 ), which is associated with the heavier lattice defects of HEAs than of the conventional alloys with the same carrier density. Acknowledgement SKC would like to thank the financial support of this research from the National Science Council of the Republic of China, Taiwan , under the Contract No. NSC96-2221-E007-066-MY3. Ted Knoy is appreciated for his editorial assistance. References [1] K.H. Huang, A Study on the Multicomponent Alloy Systems Containing Equal-mole Elements, M.S. thesis, Department of Material Science and Engineering, NTHU, Taiwan, 1996. [2] Y.F. Kao S.K. Chen J.H. Sheu J.T. Lin W.E. Lin J.W. Yeh S.J. Lin T.H. Liou C.W. Wang Int. J. Hydrogen Energy 35 2010 9046 9059 [3] H.P. Chou Y.S. Chang S.K. Chen J.W. Yeh Mater. Sci. Eng. B163 2009 184 189 [4] Y.F. Kao T.J. Chen S.K. Chen J.W. Yeh J. Alloys Compd. 52 2009 1026 1034 [5] C.J. Tong Y.L. Chen S.K. Chen J.W. Yeh T.T. Shun C.H. Tsai S.J. Lin S.Y. Chang Metal. Mater. Trans. A36 2005 881 893 [6] J.W. Yeh Ann. Chim. Sci. Mater. 31 2006 633 648 [7] J.W. Yeh S.K. Chen S.J. Lin J.Y. Gan T.S. Chin T.T. Shun C.H. Tsau S.Y. Chang Adv. Eng. Mater. 6 2004 299 303 [8] S. Varalakshmi M. Kamaraj B.S. Murty Mater. Sci. Eng. A527 2010 1027 1030 [9] Y.J. Zhou Y. Zhang Y.L. Wang G.L. Chen Appl. Phys. Lett. 90 2007 181904 181906 [10] K.C. Hsieh C.F. Yu W.T. Hsieh W.R. Chiang J.S. Ku J.H. Lai C.P. Tu C.C. Yang J. Alloys Compd. 483 2009 209 212 [11] Y.F. Kao T.D. Lee S.K. Chen Y.S. Chang Corros. Sci. 52 2010 1026 1034 [12] Y.Y. Chen U.T. Hong H.C. Shih J.W. Yeh T. Duval Corros. Sci. 47 2005 2679 2699 [13] X.F. Wang Y. Zhang Y. Qiao G.L. Chen Intermetallics 15 2007 357 362 [14] W.Y. Tang J.W. Yeh Metall. Mater. Trans. A 40 2009 1479 1486 [15] S.T. Chen W.Y. Tang Y.F. Kuo S.Y. Chen C.H. Tsau T.T. Shun J.W. Yeh Mater. Sci. Eng. A 527 2010 5818 5825 [16] J.M. Wu S.J. Lin J.W. Yeh S.K. Chen Y.S. Huang H.C. Chen Wear 261 2006 513 519 [17] C.P. Lee C.C. Chang Y.Y. Chen J.W. Yeh H.C. Shih Corrs. Sci. 50 2008 2053 2060 [18] M.H. Cai, Study on the Microstructure and Electrical Properties Evolution of High-Entropy Alloy Thin Films, M.S. thesis, Department of Material Science and Engineering, NTHU, Taiwan, 2003. [19] T. Saito J. Alloys Compd. 505 2009 23 28 [20] S. Chakraborty A.K. Majumdar J. Magn. Magn. Mater. 186 1998 357 372 [21] A.A. Al-Aql Mater. Des. 24 2003 547 550 [22] C. Kittel Introduction to Solid State Physics 8th ed. 2005 Wiley New York [23] P.J. Cote L.V. Meisel Phys. Rev. Lett. 39 1977 102 105 [24] J. Kondo Prog. Theo. Phys. 32 1964 37 49 [25] B. Shen Q. Guo Y. Gong W.S. Zhan J.G. Zhao J. Appl. Phys. 81 1997 4661 4663 [26] B.D. Cullity Introduction to Magnetic Materials 1972 Addison-Wesley Reading, Massachusetts [27] L. Berger G. Bergmann C.L. Chien C.R. Westgate The Hall Effect and Its Applications 1980 Plenum Press New York and London [28] J. Smit Physica 16 1951 612 627 [29] L. Berger Phys. Rev. B 2 1970 4559 4566 [30] H. Beck H.-J. Gűntherodt Glassy Metals III, Topics in Applied Physics vol. 72 1994 Springer Berlin [31] P.C. Chu, Hall effect and related electrical and magnetic properties in various quinary and senary high-entropy alloys selected from Al, Co, Cr, Fe, Ni, and Ti, M.S. thesis, Department of Material Science and Engineering, NTHU, Taiwan, 2009.
What problem does this paper attempt to address?