research papers\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logoJOURNAL OF
SYNCHROTRON
RADIATION
ISSN: 1600-5775

Origin of the X-ray magnetic circular dichroism at the L-edges of the rare-earths in RxR1−x′Al2 systems

CROSSMARK_Color_square_no_text.svg

aInstituto de Ciencia de Materiales de Aragón, CSIC-Universidad de Zaragoza, 50009 Zaragoza, Spain, bDepartamento de Fisica de la Materia Condensada, Universidad de Zaragoza, 50009 Zaragoza, Spain, cAdvanced Photon Source, Argonne National Laboratory, Argonne, IL 60439, USA, dDepartamento de Ciencia y Tecnología de Materiales y Fluidos, Universidad de Zaragoza, 50009 Zaragoza, Spain, eNational Institute of Materials Physics, PO Box MG-07, Bucharest, Romania, fGraduate School of Science, Hiroshima University, 1-3-1 Kagamiyama, Higashi-Hiroshima 739-8526, Japan, and gJapan Synchrotron Radiation Research Institute/SPring-8, 1-1-1 Kouto, Sayo, Hyogo 679-5198, Japan
*Correspondence e-mail: jchaboy@unizar.es

(Received 9 January 2009; accepted 17 March 2009; online 1 April 2009)

An X-ray magnetic circular dichroism (XMCD) study performed at the rare-earth L2,3-edges in the RxR1−x′Al2 compounds is presented. It is shown that both R and R′ atoms contribute to the XMCD recorded at the L-edges of the selected rare-earth, either R or R′. The amplitude of the XMCD signal is not directly correlated to the magnetization or to the value of the individual (R, R′) magnetic moments, but it is related to the molecular field acting on the rare-earth tuned in the photoabsorption process. This result closes a longstanding study of the origin of the XMCD at the L-edge of the rare-earths in multi-component systems, allowing a full understanding of the exact nature of these signals.

1. Introduction

In recent years the advent of new synchrotron radiation sources has led to the development of magnetic studies on the microscopic scale by using X-ray core-level spectroscopies such as X-ray circular magnetic dichroism (XMCD) (Lovesey & Collins, 1996[Lovesey, S. W. & Collins, S. P. (1996). Editors. X-ray Scattering and Absorption by Magnetic Materials. London: Clarendon.]; Stöhr, 1999[Stöhr, J. (1999). J. Magn. Magn. Mater. 200, 470-497.]; Kortright et al., 1999[Kortright, J. B., Awschalom, D. D., Stöhr, J., Bader, S. D., Idzerda, Y. U., Parkin, S. S. P., Schuller, I. K. & Siegmann, H.-C. (1999). J. Magn. Magn. Mater. 207, 7-44.]; Funk et al., 2005[Funk, T., Deb, A., George, S. J., Wang, H. & Cramer, S. P. (2005). Coord. Chem. Rev. 249, 3-30.]). The great advantage of using X-rays lies in the fact that each element can be probed separately since the energy of the X-ray absorption edges is characteristic for each element. XMCD combined with the so-called sum rules may provide element-specific information about the orbital (Thole et al., 1992[Thole, B. T., Carra, P., Sette, F. & van der Laan, G. (1992). Phys. Rev. Lett. 68, 1943-1946.]) and spin (Carra et al., 1993[Carra, P., Thole, B. T., Altarelli, M. & Wang, X. (1993). Phys. Rev. Lett. 70, 694-697.]) magnetic moments of the states of the absorbing atom probed in the absorption process.

In this way, XMCD has been mostly applied to those cases in which the final states are localized, such as the d-states (L2,3-edges) of the late transition metals like Co and Ni (Lovesey & Collins, 1996[Lovesey, S. W. & Collins, S. P. (1996). Editors. X-ray Scattering and Absorption by Magnetic Materials. London: Clarendon.]; Stöhr, 1999[Stöhr, J. (1999). J. Magn. Magn. Mater. 200, 470-497.]; Kortright et al., 1999[Kortright, J. B., Awschalom, D. D., Stöhr, J., Bader, S. D., Idzerda, Y. U., Parkin, S. S. P., Schuller, I. K. & Siegmann, H.-C. (1999). J. Magn. Magn. Mater. 207, 7-44.]; Chen et al., 1993[Chen, C. T., Idzerda, Y. U., Lin, H.-J., Meigs, G., Chaiken, A., Prinz, G. A. & Ho, G. H. (1993). Phys. Rev. B, 48, 642-645.]). However, the application of the spin sum rule (Carra et al., 1993[Carra, P., Thole, B. T., Altarelli, M. & Wang, X. (1993). Phys. Rev. Lett. 70, 694-697.]) worsens in other cases because of two main problems: the unknown value of the magnetic dipole operator Tz and the need for a large L2,3-edge spin–orbit splitting. This 2p spin–orbit splitting is strongly reduced for light transition metals, and quantum mechanical mixing of j3/2 and j1/2 excitations is present. This mixing reduces the observed XMCD related spin and magnetic dipole term contributions and prevents the direct application of XMCD spin sum rules (Goering, 2005[Goering, E. (2005). Philos. Mag. B, 85, 2895-2911.]; Crocombettey et al., 1996[Crocombettey, J. P., Thole, B. T. & Jollet, F. (1996). J. Phys. Condens. Matter, 8, 4095-4105.]). Moreover, the strong electron core-hole correlations in the early 3d transition metals lead to dramatic changes in the spectral intensities and a severe deviation of the spin sum rule (Scherz et al., 2004[Scherz, A., Wende, H. & Baberschke, K. (2004). Appl. Phys. A, 78, 843-846.]).

The problem of extracting quantitative magnetic information from the XMCD spectra increases in those cases in which the final states are delocalized, such as the 4p states (K-edge) of the 3d transition metals (T) and the 5d states (L2,3-edges) of the rare-earths (R). Initially, the Fe K-edge XMCD was thought to be proportional to the p-projected spin density of states (Schütz et al., 1987[Schütz, G., Wagner, W., Wilhelm, W., Kienle, P., Zeller, R., Frahm, R. & Materlik, G. (1987). Phys. Rev. Lett. 58, 737-740.]). However, this interpretation failed to account for the cases of Co and Ni (Schütz & Wienke, 1989[Schütz, G. & Wienke, H. (1989). Hyperfine Interact. 50, 457-476.]; Schütz et al., 1989[Schütz, G., Frahm, R., Wienke, R., Wilhelm, W., Wagner, W. & Kienle, P. (1989). Rev. Sci. Instrum. 60, 1661-1665.]; Stähler et al., 1993[Stähler, S., Schütz, G. & Ebert, H. (1993). Phys. Rev. B, 47, 818-826.]). According to Igarashi & Hirai, who pointed out the critical role of the hybridization of the 4p states with the 3d states at neighbouring sites (Igarashi & Hirai, 1994[Igarashi, J. & Hirai, K. (1994). Phys. Rev. B, 50, 17820-17829.], 1996[Igarashi, J. & Hirai, K. (1996). Phys. Rev. B, 53, 6442-6450.]), the shape of the XMCD spectrum near the K-edge of the ferromagnetic metals such as Fe, Co and Ni is determined by the 3d-projected orbital magnetization density of states. On the other hand, two main problems limit the use of XMCD for studying the magnetic properties of the 5d states of the rare-earths by using XMCD at the R L2,3-edges. Both the contribution of quadrupolar transitions (Carra et al., 1991[Carra, P., Harmon, B. N., Thole, B. T., Altarelli, M. & Sawatzky, G. A. (1991). Phys. Rev. Lett. 66, 2495-2498.]; Lang et al., 1992[Lang, J. C., Kycia, S. W., Wang, X. D., Harmon, B. N., Goldman, A. I., Branagan, D. J., McCallum, R. W. & Finkelstein, K. D. (1992). Phys. Rev. B, 46, 5298-5302.], 1994[Lang, J. C., Wang, X., Harmon, B. N., Goldman, A. I., Dennis, K. W., McCallum, R. W. & Finkelstein, K. D. (1994). Phys. Rev. B, 50, 13805-13808.], 1995[Lang, J. C., Srajer, G., Detlefs, C., Goldman, A. I., König, H., Wang, X., Harmon, B. N. & McCallum, R. W. (1995). Phys. Rev. Lett. 74, 4935-4938.]; Chaboy et al., 1998a[Chaboy, J., Bartolomé, F., García, L. M. & Cibin, G. (1998a). Phys. Rev. B, 57, R5598-R5601.]) and the spin dependence of the radial matrix elements of the dipolar transitions (Wang et al., 1993[Wang, X., Leung, T. C., Harmon, B. N. & Carra, P. (1993). Phys. Rev. B, 47, 9087-9090.]) affect both the shape and the sign of the XMCD signals. The appearance of these effects prevents the determination of the 5d magnetic moments from the simple application of the sum rules and, moreover, the correct determination of the sign of the magnetic coupling (Schütz et al., 1988[Schütz, G., Wienke, R., Knülle, M., Wilhelm, W., Wagner, W., Kienle, P. & Frahm, R. (1988). Z. Phys. B, 73, 67-75.]).

These problems become more acute when both localized and delocalized moments are present in the material, as in the case of RT intermetallics (Laguna-Marco, 2007[Laguna-Marco, M. A. (2007). Editor. A New Insight into the Interpretation of the T K-edge and R L2,3-Edges XMCD Spectra in R-T Intermetallics. Zaragoza: Prensas Universitarias de Zaragoza.]). Several works have shown that the rare-earth contributes to the XMCD spectrum recorded at the T K-edge (Chaboy et al., 1996[Chaboy, J., Maruyama, H., García, L. M., Bartolomé, J., Kobayashi, K., Kawamura, N., Marcelli, A. & Bozukov, L. (1996). Phys. Rev. B, 54, R15637-R15640.], 1998b[Chaboy, J., García, L. M., Bartolomé, F., Maruyama, H., Marcelli, A. & Bozukov, L. (1998b). Phys. Rev. B, 57, 13386-13389.], 2003[Chaboy, J., Piquer, C., Plugaru, N., Artigas, M., Maruyama, H., Kawamura, N. & Suzuki, M. (2003). J. Appl. Phys. 93, 475-478.], 2004[Chaboy, J., Laguna-Marco, M. A., Sánchez, M. C., Maruyama, H., Kawamura, N. & Suzuki, M. (2004). Phys. Rev. B, 69, 134421.], 2007a[Chaboy, J., Laguna-Marco, M. A., Maruyama, H., Ishimatsu, N., Isohama, Y. & Kawamura, N. (2007a). Phys. Rev. B, 75, 144405.],d[Chaboy, J., Piquer, C., Plugaru, N., Bartolomé, F., Laguna-Marco, M. A. & Plazaola, F. (2007d). Phys. Rev. B, 76, 134408.]; Laguna-Marco et al., 2005a[Laguna-Marco, M. A., Chaboy, J. & Maruyama, H. (2005a). Phys. Rev. B, 72, 094408.], 2007[Laguna-Marco, M. A. (2007). Editor. A New Insight into the Interpretation of the T K-edge and R L2,3-Edges XMCD Spectra in R-T Intermetallics. Zaragoza: Prensas Universitarias de Zaragoza.]; Ishimatsu et al., 2007[Ishimatsu, N., Miyamoto, S., Maruyama, H., Chaboy, J., Laguna-Marco, M. A. & Kawamura, N. (2007). Phys. Rev. B, 75, 180402.]; Rueff et al., 1998[Rueff, J. P., Galéra, R. M., Giorgetti, C., Dartyge, E., Brouder, C. & Alouani, M. (1998). Phys. Rev. B, 58, 12271-12281.]) and, conversely, the transition metal contributes to the XMCD signals recorded at the rare-earth L2,3-edges (Laguna-Marco et al., 2005b[Laguna-Marco, M. A., Chaboy, J., Piquer, C., Maruyama, H., Ishimatsu, N., Kawamura, N., Takagaki, M. & Suzuki, M. (2005b). Phys. Rev. B, 72, 052412.], 2008a[Laguna-Marco, M. A., Chaboy, J. & Piquer, C. (2008a). Phys. Rev. B, 77, 125132.],b[Laguna-Marco, M. A., Chaboy, J. & Piquer, C. (2008b). J. Appl. Phys. 103, 07E141.]; Giorgetti et al., 2004[Giorgetti, C., Dartyge, E., Baudelet, F. & Galéra, R.-M. (2004). Phys. Rev. B, 70, 035105.]; Chaboy et al., 2008[Chaboy, J., Laguna-Marco, M. Á., Piquer, C., Boada, R., Maruyama, H. & Kawamura, N. (2008). J. Synchrotron Rad. 15, 440-448.]; Boada et al., 2009a[Boada, R., Laguna-Marco, M. Á. & Chaboy, J. (2009). J. Synchrotron Rad. 16, 38-42.]). The systematic work performed to date has demonstrated the universality of this behaviour within the RT intermetallics. However, while the shape of these contributions does not depend significantly on the particular R/T stoichiometry, especially at the R L-edges, their amplitude shows a strong variation through different series. Indeed, in several cases of pure RT2 compounds (T = Fe, Co) the rare-earth contribution to the T K-edge XMCD spectra is so large as to hinder that of the T 4p states that, however, are still present as demonstrated in the (RxLu1−x)Fe2 series by diluting the magnetic rare-earth with a non-magnetic one (Boada et al., 2009b[Boada, R., Piquer, C., Laguna-Marco, M. A., Kawamura, N., Suzuki, M. & Chaboy, J. (2009). Phys. Rev. B. Submitted.]). This contribution is associated with the exchange interaction with the rare-earth, being mainly determined by the magnitude of the molecular field that the rare-earth magnetic moments cause at the T sites (Laguna-Marco et al., 2009[Laguna-Marco, M. A., Piquer, C. & Chaboy, J. (2009). Phys. Rev. B. Submitted.]). A similar scheme was applied to the case of the T contribution to the rare-earth L2,3-edges XMCD. Our results have shown that it is possible to establish a relationship between the T contribution to the rare-earth L2,3-edges XMCD and the molecular field acting on the absorbing sites (Chaboy et al., 2007b[Chaboy, J., Laguna-Marco, M. A., Piquer, C., Maruyama, H. & Kawamura, N. (2007b). J. Phys. Condens. Matter, 19, 436225.]). These results suggest the possibility of extracting quantitative magnetic information from the analysis of the rare-earth L2,3-edges XMCD, especially regarding both the exchange interaction between the 5d moments and those of the matrix, and also on the R(5d)–T(d) hybridization.

The experiments above have been performed in systems in which there is competition between the localized 4f magnetism of the rare-earth and the itinerant magnetism of the transition metal, and in which the R(5d)–T(d) hybridization plays a fundamental role into governing the magnetic properties of the compounds (Campbell, 1972[Campbell, I. A. (1972). J. Phys. F, 2, L47-L50.]; Yamada et al., 1984[Yamada, H., Inoue, J., Terao, K., Kanda, S. & Shimizu, M. (1984). J. Phys. F, 14, 1943-1960.]; Yamada & Shimizu, 1985[Yamada, H. & Shimizu, M. (1985). J. Phys. F, 15, L175-L180.]). However, similar behaviour has been observed in the R(Al1−xTx)2 Laves phases (Laguna-Marco, 2007[Laguna-Marco, M. A. (2007). Editor. A New Insight into the Interpretation of the T K-edge and R L2,3-Edges XMCD Spectra in R-T Intermetallics. Zaragoza: Prensas Universitarias de Zaragoza.]; Laguna-Marco et al., 2007a[Laguna-Marco, M. A., Chaboy, J., Piquer, C., Maruyama, H., Ishimatsu, N. & Kawamura, N. (2007a). AIP Conf. Proc. 882, 484-486.],b[Laguna-Marco, M. A., Chaboy, J., Piquer, C., Maruyama, H., Kawamura, N. & Takagaki, M. (2007b). J. Magn. Magn. Mater. 316, e425-e427.], 2009[Laguna-Marco, M. A., Piquer, C. & Chaboy, J. (2009). Phys. Rev. B. Submitted.]) in which the magnetism shows an evolution from RKKY-like to itinerant magnetism. Therefore, we have tailored the study of the relationship between the XMCD signals recorded at the L2,3-edges of the rare-earths and the magnetic properties of systems in which no T-3d magnetic moment and, consequently, no R(5d)–Fe(3d) hybridization are present. In this way, the systematic XMCD study performed on the RAl2 series has shown that there is a clear relationship between the rare-earth L2-edge XMCD and the molecular field coefficients nRR (Laguna-Marco et al., 2008b[Laguna-Marco, M. A., Chaboy, J. & Piquer, C. (2008b). J. Appl. Phys. 103, 07E141.]). More specifically, our results showed that the intensity of the R L2 XMCD spectra remarkably mimics the modification of the nRR coefficient through the RAl2 series, which suggests that analysis of the XMCD signal can be used to obtain valuable information about the R(4f )–R(5d) exchange.

In order to go further in this subject we have extended our previous studies to the case of the substituted R1−x[R_{x}^{\,\prime}]Al2 compounds. Here we present an XMCD study performed at the rare-earth L2,3-edges in the Rx[R_{1-x}^{\,\,\prime}]Al2 compounds. The magnetic behaviour of these systems is accounted for in terms of the RKKY interaction and, consequently, it is not expected that the hybridization of the 5d states plays a significant role. However, our results show that in the case of the substituted Gd1−x[R_{x}^{\,\prime}]Al2 compounds (with R′ = Dy, Ho, Er) the amplitude of the Gd L2,3-edge XMCD spectra decreases upon doping despite the magnetization of the system increasing. In contrast, the substitution of Gd by R′ entails a strong modification of the magnetic ordering temperature, which indicates that the exchange interaction is strongly modified with respect to that of the undoped RAl2 compounds. In this work it is shown that the observed decrease in the amplitude of the Gd L2,3-edge XMCD is due to an R′ contribution, whose intensity is correlated with the molecular field that the R′ atoms exert on the Gd sublattice. Hence, our experimental findings provide a deeper insight into the interpretation of the L2,3-edge XMCD spectra of the rare-earths by showing that the amplitude of the XMCD signals can be quantitatively related to the magnetic properties of the systems under study. In this way we have proven that these signals are determined not only by the magnetization of the probed sublattice but also by the other magnetic moments present in the material which should be considered in terms of the molecular field acting on the absorbing sites. These results open the possibility of using XMCD at the rare-earth L2,3-edges to extract experimentally quantitative magnetic information.

2. Experimental

RAl2 compounds (R = Pr, Nd, Sm, Gd, Tb, Dy, Ho and Er) were prepared by melting the pure elements in a high-frequency induction furnace, under Ar protective atmosphere. The as-cast alloys were wrapped in Ta foil and enclosed in silica tubes, under Ar gas. All the alloys were annealed at 1073 K for 100 h and then quenched to room temperature. A similar procedure was applied to the synthesis of the Gd1−xRxAl2 compounds (R = Dy, Ho, Er and Lu) that were annealed at 1023 K for one week. Structural characterization was performed at room temperature by means of powder X-ray diffraction, using a rotating-anode Rigaku diffractometer in the Bragg–Brentano geometry, with Cu Kα radiation. The magnetic measurements were performed by using a commercial SQUID magnetometer (Quantum Design MPMS-5s). Magnetic isotherms were measured on loose powders in applied magnetic fields H ≤ 5 T.

XMCD experiments were performed at the beamline BL39XU of the SPring-8 facility (Maruyama, 2001[Maruyama, H. (2001). J. Synchrotron Rad. 8, 125-128.]). For the measurements, homogeneous layers of the powdered samples were made by spreading fine powders of the material onto adhesive tape. The thickness and homogeneity of the samples were optimized to obtain the best signal-to-noise ratio, giving a total absorption jump of ∼1 at about 150 eV above the edge. In all of the cases, the origin of the energy scale, E0, was chosen at the inflection point of the absorption edge, and the XAS spectra were normalized to the averaged absorption coefficient at high energy. The XMCD spectra were recorded at different fixed temperatures in the transmission mode using the helicity-modulation technique (Suzuki et al., 1998[Suzuki, M., Kawamura, N., Mizumaki, M., Urata, A., Maruyama, H., Goto, S. & Ishikawa, T. (1998). Jpn. J. Appl. Phys. 37, L1488-L1490.]). The sample was magnetized by an external magnetic field, H = 2 T, applied in the direction of the incident beam, and the helicity was changed from positive to negative at each energy point. In this way the spin-dependent absorption coefficient was obtained as the difference of the absorption coefficient, μc = (μμ+), for antiparallel, μ, and parallel, μ+, orientation of the photon helicity and sample magnetization. For the sake of accuracy, the direction of the applied magnetic field was reversed to obtain the XMCD, now corresponding to μc = (μ+μ), by switching the helicity. Subtraction of the XMCD spectra recorded for both field orientations cancels, if present, any spurious signal.

3. Results and discussion

Both RAl2 and R1−x[R_{x}^{\,\prime}]Al2 series of compounds crystallize in the cubic Laves C15 MgCu2 structure. The diffraction patterns were Rietveld refined using the FULLPROF code (Rodriguez-Carvajal, 1993[Rodriguez-Carvajal, J. (1993). Physica B, 192, 55-69.]). All the samples were found to be single phase and the cell parameters were determined from the X-ray diffraction patterns (see Fig. 1[link]). The cell parameters of both the RAl2 and the Gd1−xRxAl2 compounds are summarized in Table 1[link]. As shown in Fig. 1[link], the cell volume of the pure RAl2 compounds decreases as the atomic number increases, as expected because of the lanthanide contraction. In addition, the cell parameters of the Gd1−xRxAl2 compounds follow a linear variation with the composition (x) between the end-members, in agreement with Vegard's law.

Table 1
Main structural and magnetic parameters of the RAl2 and R1−x[R_{x}^{\,\prime}]Al2 compounds: lattice constant (a), unit-cell volume (V), Curie temperature (TC) and the magnetization measured at T = 4.2 K and H = 5 T (M5T)

Compound a (Å) V3) TC (K) M5T (μB/f.u.)
PrAl2 8.028 517.3 33 2.94
NdAl2 8.001 512.1 79 2.39
SmAl2 7.942 501.0 64 0.24
GdAl2 7.898 492.6 164 6.74
TbAl2 7.863 486.1 109 8.90
DyAl2 7.837 481.4 58 9.64
HoAl2 7.812 476.8 29 9.20
ErAl2 7.791 472.9 13 7.63
Gd0.75Dy0.25Al2 7.880 489.3 145 8.37
Gd0.50Dy0.50Al2 7.864 486.2 115 8.76
Gd0.75Ho0.25Al2 7.874 488.2 141 8.46
Gd0.50Ho0.50Al2 7.855 484.7 101 8.64
Gd0.75Er0.25Al2 7.875 488.4 135 7.80
Gd0.50Er0.50Al2 7.839 481.7 94 7.73
Gd0.75Lu0.25Al2 7.856 484.8 128 5.26
[Figure 1]
Figure 1
(a) X-ray powder diffraction patterns of the RAl2 series. (b) Variation of the lattice parameter through the RAl2 series. (c) Dependence of the lattice parameter on the rare-earth substitution in the Gd1−x[R_{x}^{\,\prime}]Al2 [R′ = Dy (filled circles), Ho (red, open circles), Er (blue, filled squares) and Lu (green, open squares)] series. This figure is in colour in the electronic version of this paper.

The temperature dependence of the magnetization through the RAl2 series is shown in Fig. 2[link] together with the magnetization versus applied magnetic field curves recorded at 4.2 K. The RAl2 compounds order ferromagnetically. The magnetic order temperature, TC, of the series scales well with the de Gennes factor (g − 1)2J(J + 1) (de Gennes, 1966[Gennes, P. G. de (1966). C. R. Acad. Sci. (Paris), 247, 1836-1838.]), in agreement with the predictions of the simple RKKY model. This result indicates that the magnetic order is mainly driven by the R sublattice, which differs from the isomorphous RFe2 compounds where magnetic order is governed by the itinerant subsystem. The magnetic moments of these compounds are slightly smaller than the corresponding free rare-earth ion values, gJ, indicating some crystal field quenching (Purwins & Leson, 1990[Purwins, H. G. & Leson, A. (1990). Adv. Phys. 39, 309-405.]). The results summarized in Fig. 2[link] and Table 1[link] are in agreement with the numerous investigations that have been performed on RAl2 compounds (Buschow, 1979[Buschow, K. H. J. (1979). Rep. Prog. Phys. 42, 1374-1477.]; Purwins & Leson, 1990[Purwins, H. G. & Leson, A. (1990). Adv. Phys. 39, 309-405.]).

[Figure 2]
Figure 2
(a) Temperature dependence of the magnetization of the RAl2 compounds measured under an applied magnetic field H = 0.1 T. (b) Magnetization versus applied magnetic field curves of the RAl2 compounds recorded at T = 4.2 K. (c) Comparison of the magnetization of the RAl2 at T = 4.2 K and H = 5 T (black, filled circles) and the free-ion μ4f magnetic moments (red, open circles). (d) Comparison of the Curie temperatures of the RAl2 compounds (red, open circles) and the J(J + 1)(g − 1)2 dependence (black, filled circles) through the series. This figure is in colour in the electronic version of this paper.

Within the series, GdAl2 exhibits the highest Curie temperature (TC = 164 K). This fact, and the absence of any significant effect of crystal fields on the S state of the Gd3+ ion (Burd & Lee, 1977[Burd, J. & Lee, E. W. (1977). J. Phys. C, 10, 4581-4586.]), makes GdAl2 a good candidate for studying systematically the effect of replacing Gd by a different rare-earth. Hence, we have synthesized the Gd1−xRxAl2 compounds in which Gd is replaced by Dy, Ho, Er and non-magnetic Lu. The dilution of Gd by Lu is expected to act as a simple magnetic dilution effect because Lu carries no magnetic moment and the magnetic state of Gd is thought to remain unaffected by the substitution. For a 25% Lu substitution, the magnetization decreases from 6.74 μB/f.u. in GdAl2 to 5.26 μB/f.u. Hence, it is 78% of the magnetization of the parent compound, in agreement with the expected magnetic dilution effect. A similar reduction is found for the magnetic order temperature that decreases from 164 K in GdAl2 to 128 K in Gd0.75Lu0.25Al2 (see Fig. 3[link]).

[Figure 3]
Figure 3
(a) Temperature dependence of the magnetization of the Gd1−x[R_{x}^{\,\prime}]Al2 compounds measured under an applied magnetic field H = 0.1 T: R′ = Er (blue, filled down triangles), Ho (green, filled circles) and Dy (red, filled triangles); and x = 0.25 (solid symbols); x = 0.5 (open symbols). (b) Comparison of the magnetization versus applied magnetic field curves of the pure RAl2 (solid symbols) and Gd0.5[R_{0.5}^{\,\prime}]Al2 (open symbols) compounds for R′ = Er (blue, open down triangles), Ho (green, open circles) and Dy (red, open triangles). This figure is in colour in the electronic version of this paper.

In the case of the substitution of Gd by a magnetic rare-earth, the ferromagnetic order is preserved although the substitution has different consequences regarding both the magnetization and the magnetic ordering temperature of the systems. The magnetization of Gd1−xRxAl2 is reinforced with respect to what is expected from the addition of the magnet­ization of both RAl2 and R′Al2 pure compounds according to a two-sublattice model. Within this framework, it is assumed that the total magnetization of the Gd0.5[R_{0.5}^{\,\prime}]Al2 compounds corresponds to the weighted addition of the magnetization of the parent single compounds, MTot = 0.5M(GdAl2) + 0.5M(R′Al2). In the case of Gd0.5Dy0.5Al2, this model yields MCalc = 8.19 μB/f.u., while the measured magnetization at T = 4.2 K and H = 5 T is Mexp = 8.76 μB/f.u. The same trend is found for both Gd0.5Ho0.5Al2, Mexp = 8.64 versus MCalc 7.97 μB/f.u., and Gd0.5Er0.5Al2, Mexp = 7.73 versus MCalc 7.19 μB/f.u. Interestingly, a better agreement is found if the magnetization is calculated by using the 4f free-ion magnetic moments values: MCalc = 8.5 μB/f.u. for Dy and Ho, and 8 μB/f.u. for Er. These results indicate that the rare-earth ions in the Gd0.5[R_{0.5}^{\,\prime}]Al2 compounds retain their magnetic properties upon dilution. However, the modification of the magnetic ordering temperature indicates that the exchange interaction is strongly modified with respect to that of the RAl2 compounds.

This situation represents the best starting point to study the relationship between the rare-earth L2,3-edge XMCD signals and the magnetic properties of the materials. In principle, if the amplitude of the XMCD signal was proportional to the magnetic moment of the absorbing atom or to the magnetization of the compound, no significant variation of the amplitude of the XMCD signal recorded at the Gd L2,3-edges through the Gd0.5[R_{0.5}^{\,\prime}]Al2 series would be expected. The total magnetization of the systems increases after the Gd substitution. Therefore, the amplitude of the XMCD signals should be increased through the Gd0.5[R_{0.5}^{\,\prime}]Al2 series. By contrast, if the amplitude of the XMCD signals was determined by the exchange interaction, a strong modification of the XMCD amplitude should be expected. In this respect, we show in Fig. 4[link] the XMCD spectra recorded at the Gd L2-edge in the investigated Gd0.5[R_{0.5}^{\,\prime}]Al2 compounds. The Gd L2-edge XMCD spectrum shows in all cases a main negative peak at the edge. This spectral profile is similar for all the compounds and only the amplitude of the signal is modified upon substitution of Gd by another lanthanide. Similar results are found at the Gd L3-edge. In this case, the XMCD is positive and its intensity is one half of the L2 one. Moreover, the intensity of the main peak of the XMCD spectrum and the integral of the signal over all the experimental range scale perfectly at both the L2- and L3-edges. Therefore, the modification of the amplitude of the XMCD spectra will be referred hereafter to the intensity values.

[Figure 4]
Figure 4
(a) Comparison of the XMCD spectra recorded at T = 50 K at the Gd L2-edge in the case of GdAl2 (black, filled circles), the Gd0.5[R_{0.5}^{\,\prime}]Al2 compounds with R′ = Er (blue, open squares), Ho (green, filled squares), Dy (red, open circles), and in Gd0.75Lu0.25Al2 (dark green, open triangles). (b) Comparison of the XMCD spectra recorded at T = 50 K at the Gd L2-edge in the Gd0.5[R_{0.5}^{\,\prime}]Al2 compounds after subtracting the XMCD of the GdAl2 one: R′ = Er (blue, open squares), Ho (green, filled squares), Dy (red, open circles), and in Gd0.75Lu0.25Al2 (black, filled circles). This figure is in colour in the electronic version of this paper.

As shown in Fig. 4[link], the intensity of the Gd L2-edge XMCD in Gd0.75Lu0.25Al2 is ∼90% of that in GdAl2. However, the intensity of the XMCD signal markedly decreases through the Gd0.5[R_{0.5}^{\,\prime}]Al2 series. Compared with that of GdAl2, its amplitude is ∼80%, 50% and 25% for Dy, Ho and Er, respectively. Such a variation is not expected in terms of the diminution of the Gd magnetic moment. Similarly, this behaviour is not expected on the basis of the modification of the total magnetization, which increases after substitution. It might be argued that the magnetization at the working temperatures departs from the relation found at T = 4.2 K and H = 5 T. However, this is not the case. We have recorded the magnetization under the same experimental conditions as for the XMCD measurements, i.e. T = 50 K and H = 2 T, verifying that a similar relation holds. Under these conditions the magnetization of GdAl2, Gd0.5Dy0.5Al2, Gd0.5Ho0.5Al2 and Gd0.75Lu0.25Al2 is ∼86% of that at T = 4.2 K and H = 5 T, while it is only ∼60% in the case of Gd0.5Er0.5Al2. Finally, it should also be noted that the variation of the XMCD intensity through the Gd0.5[R_{0.5}^{\,\prime}]Al2 series does not follow that of the magnetic ordering temperature. Indeed, as shown in Table 1[link], the maximum variation should be expected for the Gd0.5Er0.5Al2 compound for which TC/TC(GdAl2) ≃ 0.6. By contrast, the ratio of the XMCD amplitude for this compound, XMCD/XMCD(GdAl2) = 0.25, is significantly smaller than the ratio of the magnetic ordering temperatures.

Atempting to gain a deeper insight into this peculiar behaviour, we have considered our previous findings in the case of RT2 compounds in which T is a 3d transition metal such as Fe or Co. In these cases we have demonstrated that the transition metal contributes to the XMCD signals recorded at the rare-earth L2,3-edges (Laguna-Marco et al., 2005b[Laguna-Marco, M. A., Chaboy, J., Piquer, C., Maruyama, H., Ishimatsu, N., Kawamura, N., Takagaki, M. & Suzuki, M. (2005b). Phys. Rev. B, 72, 052412.], 2007a[Laguna-Marco, M. A., Chaboy, J., Piquer, C., Maruyama, H., Ishimatsu, N. & Kawamura, N. (2007a). AIP Conf. Proc. 882, 484-486.], 2008a[Laguna-Marco, M. A., Chaboy, J. & Piquer, C. (2008a). Phys. Rev. B, 77, 125132.],b[Laguna-Marco, M. A., Chaboy, J. & Piquer, C. (2008b). J. Appl. Phys. 103, 07E141.]), and, moreover, that this contribution is related to the molecular field owing to the transition metal acting at the rare-earth sites (Chaboy et al., 2007b[Chaboy, J., Laguna-Marco, M. A., Piquer, C., Maruyama, H. & Kawamura, N. (2007b). J. Phys. Condens. Matter, 19, 436225.]). Then, we have checked the possibility that the rare-earth substituting Gd in the Gd0.5[R_{0.5}^{\,\prime}]Al2 compounds could also contribute to the XMCD of the Gd L-edges. The existence of such a contribution can be highlighted by subtracting the XMCD spectra of GdAl2 from those of the Gd0.5[R_{0.5}^{\,\prime}]Al2 compounds. The result of the substraction is reported in Fig. 4[link] for the case of the L2-edge (similar results are found at the L3-edge). In all cases, with the exception of Lu, the extracted signal shows a sharp positive peak at the edge. While its shape is the same for all of the compounds, the intensity shows marked differences as a function of the rare-earth (Dy, Ho, Er) that substitutes for Gd. While in the case of Lu this difference can be associated with a simple dilution effect, the magnitude of the differences found for the cases in which R is a magnetic rare-earth does not show a direct relationship with the magnetic moments. The maximum difference signal is found for the case of Er, its intensity being ∼1.5 and 4 times greater than those of Ho and Dy, respectively. As commented above, these differences cannot be easily accounted for in terms of the different magnetic moments of the rare-earths substituting Gd, i.e. μR = 10 μB for Dy and Ho, and 9 μB for Er.

We have studied whether this contribution can be regarded as being due to the hybridization of the 5d states of both Gd and R′ ions. In this way, the Gd 5d states will be polarized by the interatomic 4f–5d exchange but, in addition, the R′ ions should also contribute to the polarization of the 5d band through the action of the interatomic molecular field. It should be noted that the sign of the XMCD spectra indicates the orientation of the magnetic moment of the states probed in the photoabsorption process with respect to the magnetization of the system. However, the XMCD signals recorded at the L2,3-edges of the rare-earths yield an erroneous sign (Schütz et al., 1988[Schütz, G., Wienke, R., Knülle, M., Wilhelm, W., Wagner, W., Kienle, P. & Frahm, R. (1988). Z. Phys. B, 73, 67-75.]). Indeed, if one considers that the XMCD spectrum reflects the difference in the density of empty states with different spin moment, this model yields that the 5d spin would be antiparallel to the 4f spin, which is in contradiction with the current knowledge (Campbell, 1972[Campbell, I. A. (1972). J. Phys. F, 2, L47-L50.]; Yamada et al., 1984[Yamada, H., Inoue, J., Terao, K., Kanda, S. & Shimizu, M. (1984). J. Phys. F, 14, 1943-1960.]; Yamada & Shimizu, 1985[Yamada, H. & Shimizu, M. (1985). J. Phys. F, 15, L175-L180.]). This paradox has been solved by taking into account the spin dependence of the matrix elements previously neglected (Wang et al., 1993[Wang, X., Leung, T. C., Harmon, B. N. & Carra, P. (1993). Phys. Rev. B, 47, 9087-9090.]). It should be noted that the L2,3 XMCD returns the correct sign when the rare-earth does not exhibit a 4f magnetic moment and the 5d spin is induced by a transition metal or by another rare-earth showing a localized 4f moment. This is the case for the Lu L2,3 XMCD in LuFe2 and Gd0.75Lu0.25Al2 compounds (Chaboy et al., 2007c[Chaboy, J., Laguna-Marco, M. A., Piquer, C., Maruyama, H., Kawamura, N., Ishimatsu, N., Suzuki, M. & Takagaki, M. (2007c). Phys. Rev. B, 75, 064410.]). In a similar way, the results reported in Fig. 4[link] indicate that the polarization of the 5d states associated with the R′ atoms agrees with the ferromagnetic coupling of the Gd and R′ magnetic moments in the Gd0.5[R_{0.5}^{\,\prime}]Al2 compounds. In order to explore the relationship between the R′ contribution to the Gd L2,3 XMCD signals and the molecular field owing to the R′ ions acting on the Gd sites, we have calculated Bmol(R′ → Gd) within a mean-field approach (Buschow, 1988[Buschow, K. H. J. (1988). In Handbook of Magnetic Materials, Vol. 4, edited by K. H. J. Buschow, ch. 1. Amsterdam: Elsevier Science.]). In this model, the intersublattice molecular field coefficient describing the interaction between the two rare-earth sublattices is given by

[N_{RR'}=\left({|T_C-T_R|\,|T_C-T_{R'}|}\over{C_R\,C_{R'}}\right)^{1/2},\eqno(1)]

where CR = NRgJ2J(J + 1)μB2/3kBT, and TR and TR are the Curie temperatures of the RAl2 and R′Al2 compounds, respectively. The corresponding values of NRR and Bmol, obtained as Bmol(R′ → Gd) = NGdRμR, are listed in Table 2[link].

Table 2
Intersublattice molecular coefficient and molecular field that the R′ atoms exerts on the Gd sites

Compound NRR (10−5 T m A−1) Bmol (R′ → Gd) (T)
Gd0.5Dy0.5Al2 3.7 7.1
Gd0.5Ho0.5Al2 4.7 8.9
Gd0.5Er0.5Al2 5.8 10.0

As we can see in Table 2[link], the molecular field coefficient NRR increases from Dy to Er as the maximum intensity of the subtracted Gd L2 XMCD signal does. The NRR increase reflects the variation of TC observed along the Gd1−x[R_{x}^{\,\prime}]Al2 series. In the case of Er, TC increases from 13 K in ErAl2 to 94 K in Gd0.5Er0.5Al2, that is ΔTC = 81 K, which corresponds to a ∼623% increase. In the Ho series, TC increases from 29 K in HoAl2 to 101 K in Gd0.5Ho0.5Al2, or equivalently, ΔTC = 72 K, which corresponds to an increase of ∼248%, whereas in the case of Dy this increase is only ∼98% as TC increases from 58 K in DyAl2 to 115 K in Gd0.5Dy0.5Al2. That is, the values of TC indicate that the exchange interaction between Gd and R′ increases from Dy to Er. Finally, the obtained values of Bmol(R′ → Gd) have been compared in Fig. 5[link] with the intensity of the subtracted Gd L2 XMCD signals. The good agreement shown in this figure gives support to our hypothesis regarding the relationship between the XMCD and the molecular field acting on the absorbing sites.

[Figure 5]
Figure 5
Top: comparison of the maximum intensity of the difference XMCD spectra (black, filled circles) and the intersublattice molecular field (blue, open triangles) through the Gd0.5[R_{0.5}^{\,\prime}]Al2 series (see text for details). Bottom: comparison of the relative maximum intensity of the difference XMCD spectra prior (black, filled circles) and after applying the magnetization correction (red, open circles) (see text for details) and the relative variation of the intersublattice molecular field (blue, open triangles) through the Gd0.5[R_{0.5}^{\,\prime}]Al2 series. This figure is in colour in the electronic version of this paper.

The magnitude of the subtracted Gd L2 XMCD signals deserves a final comment, being of the same order of magnitude as those previously found in the RFe2 systems (Laguna-Marco et al., 2008a[Laguna-Marco, M. A., Chaboy, J. & Piquer, C. (2008a). Phys. Rev. B, 77, 125132.]). In this respect, it should be noted that when applying the procedure above we have subtracted the same Gd L2 XMCD signal of GdAl2 recorded at T = 50 K as those of the Gd0.5[R_{0.5}^{\,\prime}]Al2 compounds recorded under the same conditions. However, this temperature (T = 50 K) is not so far from the magnetic ordering temperatures of the substituted Gd0.5[R_{0.5}^{\,\prime}]Al2 compounds and, consequently, this approach might no longer be valid. Indeed, the magnetization measured at T = 50 K and H = 5 T is 6.09 μB/f.u. and 7.95 μB/f.u. for GdAl2 and Gd0.5Dy0.5Al2, respectively, yielding a M5T(50 K)/M5T(4.2 K) ratio of 0.9. This ratio decreases to 0.8 for Gd0.5Ho0.5Al2 [M5T(50 K) = 6.83 μB/f.u.] and to 0.7 in the case of Gd0.5Er0.5Al2 [M5T(50 K) = 5.32 μB/f.u.]. These results suggest that subtracting the XMCD signal of GdAl2 from the XMCD signals of the Gd0.5[R_{0.5}^{\,\prime}]Al2 compounds is appropriate for the Dy case, but probably not for the Ho and Er cases. Consequently, we have modified the subtraction procedure by taking into account these factors. In this way, the GdAl2 signal has been factorized by 0.8/0.9 and 0.7/0.9 prior to subtracting it from the XMCD spectra of the Ho and Er compounds, respectively. The result of such a procedure is incorporated in the bottom panel of Fig. 5[link]. As shown in the figure, while the magnitude of the extracted signal is reduced, the trend of the intensity variation through the substituted series does not vary. As a consequence, a comparison of the intensity of the extracted XMCD signals and the value of the intersublattice molecular field also shows a remarkable agreement.

After applying this correction we found that the intensity of the extracted XMCD signals in the Gd0.5[R_{0.5}^{\,\prime}]Al2 compounds is about a half of that found in the RFe2 series, while the intersublattice molecular field is about five times smaller (applying the same procedure as discussed above yields Bmol = 45.4 T for GdFe2). These differences are presumably due to the different hybridization mechanism occurring in both R0.5[R_{0.5}^{\,\prime}]Al2, R(5d)–R′(5d), and RFe2, R(5d)–Fe(3d), systems. The present results suggest that, while the magnitude of the discussed effect on the XMCD spectra is determined by the details of the interplay of the hybridization and the spin polarization of the R(5d) states, the relative variation through the same series is mainly determined by the molecular field acting on the rare-earth absorbing sites.

4. Summary and conclusions

We have presented here an XMCD study performed at the rare-earth L2,3-edges in the case of Rx[R_{1-x}^{\,\,\prime}]Al2 compounds. It is shown that both R and R′ atoms contribute to the XMCD recorded at the L-edges of the selected rare-earth. In the case of the Gdx[R_{1-x}^{\,\,\prime}]Al2 compounds, we have extracted the contribution of R′ to the Gd L2,3-edges XMCD spectra. We have shown that the variation of the intensity of this contribution through the series is determined by the molecular field owing to the R′ atoms acting on the Gd sites. These results close a longstanding study on the origin and the interpretation of the L2,3-edge XMCD spectra of the rare-earths in the case of multi-component magnetic systems.

Acknowledgements

This work was partially supported by a Spanish CICYT-MAT2008-06542-C04-01 grant. MALM and RB acknowledge the Ministerio de Eduación y Ciencia of Spain for a Postdoctoral and a PhD grant, respectively. This study was performed with the approval of Japan Synchrotron Radiation Research Institute (JASRI) (Proposals No. 1999A0388 and 2001A0062).

References

First citationBoada, R., Laguna-Marco, M. Á. & Chaboy, J. (2009). J. Synchrotron Rad. 16, 38–42.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationBoada, R., Piquer, C., Laguna-Marco, M. A., Kawamura, N., Suzuki, M. & Chaboy, J. (2009). Phys. Rev. B. Submitted.  Google Scholar
First citationBurd, J. & Lee, E. W. (1977). J. Phys. C, 10, 4581–4586.  CrossRef CAS Web of Science Google Scholar
First citationBuschow, K. H. J. (1979). Rep. Prog. Phys. 42, 1374–1477.  CrossRef Web of Science Google Scholar
First citationBuschow, K. H. J. (1988). In Handbook of Magnetic Materials, Vol. 4, edited by K. H. J. Buschow, ch. 1. Amsterdam: Elsevier Science.  Google Scholar
First citationCampbell, I. A. (1972). J. Phys. F, 2, L47–L50.  CrossRef CAS Web of Science Google Scholar
First citationCarra, P., Harmon, B. N., Thole, B. T., Altarelli, M. & Sawatzky, G. A. (1991). Phys. Rev. Lett. 66, 2495–2498.  CrossRef PubMed CAS Web of Science Google Scholar
First citationCarra, P., Thole, B. T., Altarelli, M. & Wang, X. (1993). Phys. Rev. Lett. 70, 694–697.  CrossRef PubMed CAS Web of Science Google Scholar
First citationChaboy, J., Bartolomé, F., García, L. M. & Cibin, G. (1998a). Phys. Rev. B, 57, R5598–R5601.  Web of Science CrossRef CAS Google Scholar
First citationChaboy, J., García, L. M., Bartolomé, F., Maruyama, H., Marcelli, A. & Bozukov, L. (1998b). Phys. Rev. B, 57, 13386–13389.  Web of Science CrossRef CAS Google Scholar
First citationChaboy, J., Laguna-Marco, M. A., Maruyama, H., Ishimatsu, N., Isohama, Y. & Kawamura, N. (2007a). Phys. Rev. B, 75, 144405.  Web of Science CrossRef Google Scholar
First citationChaboy, J., Laguna-Marco, M. Á., Piquer, C., Boada, R., Maruyama, H. & Kawamura, N. (2008). J. Synchrotron Rad. 15, 440–448.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationChaboy, J., Laguna-Marco, M. A., Piquer, C., Maruyama, H. & Kawamura, N. (2007b). J. Phys. Condens. Matter, 19, 436225.  Web of Science CrossRef Google Scholar
First citationChaboy, J., Laguna-Marco, M. A., Piquer, C., Maruyama, H., Kawamura, N., Ishimatsu, N., Suzuki, M. & Takagaki, M. (2007c). Phys. Rev. B, 75, 064410.  Web of Science CrossRef Google Scholar
First citationChaboy, J., Laguna-Marco, M. A., Sánchez, M. C., Maruyama, H., Kawamura, N. & Suzuki, M. (2004). Phys. Rev. B, 69, 134421.  Web of Science CrossRef Google Scholar
First citationChaboy, J., Maruyama, H., García, L. M., Bartolomé, J., Kobayashi, K., Kawamura, N., Marcelli, A. & Bozukov, L. (1996). Phys. Rev. B, 54, R15637–R15640.  CrossRef CAS Google Scholar
First citationChaboy, J., Piquer, C., Plugaru, N., Artigas, M., Maruyama, H., Kawamura, N. & Suzuki, M. (2003). J. Appl. Phys. 93, 475–478.  Web of Science CrossRef CAS Google Scholar
First citationChaboy, J., Piquer, C., Plugaru, N., Bartolomé, F., Laguna-Marco, M. A. & Plazaola, F. (2007d). Phys. Rev. B, 76, 134408.  Web of Science CrossRef Google Scholar
First citationChen, C. T., Idzerda, Y. U., Lin, H.-J., Meigs, G., Chaiken, A., Prinz, G. A. & Ho, G. H. (1993). Phys. Rev. B, 48, 642–645.  CrossRef CAS Web of Science Google Scholar
First citationCrocombettey, J. P., Thole, B. T. & Jollet, F. (1996). J. Phys. Condens. Matter, 8, 4095–4105.  Google Scholar
First citationFunk, T., Deb, A., George, S. J., Wang, H. & Cramer, S. P. (2005). Coord. Chem. Rev. 249, 3–30.  Web of Science CrossRef CAS Google Scholar
First citationGennes, P. G. de (1966). C. R. Acad. Sci. (Paris), 247, 1836–1838.  Google Scholar
First citationGiorgetti, C., Dartyge, E., Baudelet, F. & Galéra, R.-M. (2004). Phys. Rev. B, 70, 035105.  Web of Science CrossRef Google Scholar
First citationGoering, E. (2005). Philos. Mag. B, 85, 2895–2911.  CAS Google Scholar
First citationIgarashi, J. & Hirai, K. (1994). Phys. Rev. B, 50, 17820–17829.  CrossRef CAS Web of Science Google Scholar
First citationIgarashi, J. & Hirai, K. (1996). Phys. Rev. B, 53, 6442–6450.  CrossRef CAS Web of Science Google Scholar
First citationIshimatsu, N., Miyamoto, S., Maruyama, H., Chaboy, J., Laguna-Marco, M. A. & Kawamura, N. (2007). Phys. Rev. B, 75, 180402.  Web of Science CrossRef Google Scholar
First citationKortright, J. B., Awschalom, D. D., Stöhr, J., Bader, S. D., Idzerda, Y. U., Parkin, S. S. P., Schuller, I. K. & Siegmann, H.-C. (1999). J. Magn. Magn. Mater. 207, 7–44.  Web of Science CrossRef CAS Google Scholar
First citationLaguna-Marco, M. A. (2007). Editor. A New Insight into the Interpretation of the T K-edge and R L2,3-Edges XMCD Spectra in R–T Intermetallics. Zaragoza: Prensas Universitarias de Zaragoza.  Google Scholar
First citationLaguna-Marco, M. A., Chaboy, J. & Maruyama, H. (2005a). Phys. Rev. B, 72, 094408.  Web of Science CrossRef Google Scholar
First citationLaguna-Marco, M. A., Chaboy, J. & Piquer, C. (2008a). Phys. Rev. B, 77, 125132.  Web of Science CrossRef Google Scholar
First citationLaguna-Marco, M. A., Chaboy, J. & Piquer, C. (2008b). J. Appl. Phys. 103, 07E141.  Web of Science CrossRef Google Scholar
First citationLaguna-Marco, M. A., Chaboy, J., Piquer, C., Maruyama, H., Ishimatsu, N. & Kawamura, N. (2007a). AIP Conf. Proc. 882, 484–486.  CrossRef CAS Google Scholar
First citationLaguna-Marco, M. A., Chaboy, J., Piquer, C., Maruyama, H., Ishimatsu, N., Kawamura, N., Takagaki, M. & Suzuki, M. (2005b). Phys. Rev. B, 72, 052412.  Web of Science CrossRef Google Scholar
First citationLaguna-Marco, M. A., Chaboy, J., Piquer, C., Maruyama, H., Kawamura, N. & Takagaki, M. (2007b). J. Magn. Magn. Mater. 316, e425–e427.  Web of Science CrossRef CAS Google Scholar
First citationLaguna-Marco, M. A., Piquer, C. & Chaboy, J. (2009). Phys. Rev. B. Submitted.  Google Scholar
First citationLang, J. C., Kycia, S. W., Wang, X. D., Harmon, B. N., Goldman, A. I., Branagan, D. J., McCallum, R. W. & Finkelstein, K. D. (1992). Phys. Rev. B, 46, 5298–5302.  CrossRef CAS Web of Science Google Scholar
First citationLang, J. C., Srajer, G., Detlefs, C., Goldman, A. I., König, H., Wang, X., Harmon, B. N. & McCallum, R. W. (1995). Phys. Rev. Lett. 74, 4935–4938.  CrossRef PubMed CAS Web of Science Google Scholar
First citationLang, J. C., Wang, X., Harmon, B. N., Goldman, A. I., Dennis, K. W., McCallum, R. W. & Finkelstein, K. D. (1994). Phys. Rev. B, 50, 13805–13808.  CrossRef CAS Web of Science Google Scholar
First citationLovesey, S. W. & Collins, S. P. (1996). Editors. X-ray Scattering and Absorption by Magnetic Materials. London: Clarendon.  Google Scholar
First citationMaruyama, H. (2001). J. Synchrotron Rad. 8, 125–128.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationPurwins, H. G. & Leson, A. (1990). Adv. Phys. 39, 309–405.  CrossRef CAS Web of Science Google Scholar
First citationRodriguez-Carvajal, J. (1993). Physica B, 192, 55–69.  CrossRef CAS Web of Science Google Scholar
First citationRueff, J. P., Galéra, R. M., Giorgetti, C., Dartyge, E., Brouder, C. & Alouani, M. (1998). Phys. Rev. B, 58, 12271–12281.  Web of Science CrossRef CAS Google Scholar
First citationScherz, A., Wende, H. & Baberschke, K. (2004). Appl. Phys. A, 78, 843–846.  Web of Science CrossRef CAS Google Scholar
First citationSchütz, G., Frahm, R., Wienke, R., Wilhelm, W., Wagner, W. & Kienle, P. (1989). Rev. Sci. Instrum. 60, 1661–1665.  CrossRef Google Scholar
First citationSchütz, G., Wagner, W., Wilhelm, W., Kienle, P., Zeller, R., Frahm, R. & Materlik, G. (1987). Phys. Rev. Lett. 58, 737–740.  CrossRef PubMed Web of Science Google Scholar
First citationSchütz, G. & Wienke, H. (1989). Hyperfine Interact. 50, 457–476.  Google Scholar
First citationSchütz, G., Wienke, R., Knülle, M., Wilhelm, W., Wagner, W., Kienle, P. & Frahm, R. (1988). Z. Phys. B, 73, 67–75.  Google Scholar
First citationStähler, S., Schütz, G. & Ebert, H. (1993). Phys. Rev. B, 47, 818–826.  CrossRef Web of Science Google Scholar
First citationStöhr, J. (1999). J. Magn. Magn. Mater. 200, 470–497.  Web of Science CrossRef CAS Google Scholar
First citationSuzuki, M., Kawamura, N., Mizumaki, M., Urata, A., Maruyama, H., Goto, S. & Ishikawa, T. (1998). Jpn. J. Appl. Phys. 37, L1488–L1490.  Web of Science CrossRef CAS Google Scholar
First citationThole, B. T., Carra, P., Sette, F. & van der Laan, G. (1992). Phys. Rev. Lett. 68, 1943–1946.  CrossRef PubMed CAS Web of Science Google Scholar
First citationWang, X., Leung, T. C., Harmon, B. N. & Carra, P. (1993). Phys. Rev. B, 47, 9087–9090.  CrossRef CAS Web of Science Google Scholar
First citationYamada, H., Inoue, J., Terao, K., Kanda, S. & Shimizu, M. (1984). J. Phys. F, 14, 1943–1960.  CrossRef CAS Web of Science Google Scholar
First citationYamada, H. & Shimizu, M. (1985). J. Phys. F, 15, L175–L180.  CrossRef CAS Web of Science Google Scholar

© International Union of Crystallography. Prior permission is not required to reproduce short quotations, tables and figures from this article, provided the original authors and source are cited. For more information, click here.

Journal logoJOURNAL OF
SYNCHROTRON
RADIATION
ISSN: 1600-5775
Follow J. Synchrotron Rad.
Sign up for e-alerts
Follow J. Synchrotron Rad. on Twitter
Follow us on facebook
Sign up for RSS feeds