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

Journal logoJOURNAL OF
SYNCHROTRON
RADIATION
ISSN: 1600-5775

Hartree–Fock study of orbital magnetic moments in 3d and 5f magnets and X-ray magnetic circular dichroism

CROSSMARK_Color_square_no_text.svg

aDepartment of Quantum Matter, ADSM, Hiroshima University, Higashi-Hiroshima 739-8526, Japan
*Correspondence e-mail: jo@sci.hiroshima-u.ac.jp

(Received 26 July 2000; accepted 7 November 2000)

The spin (Ms) and orbital magnetic moments (Mo) of the uranium 5f state in the ferromagnetic compound uranium sulfide (US) and of the cobalt 3d state in various transition-metal superlattices are calculated on the basis of a tight-binding model, in which the intra-atomic ff or dd multipole interaction is taken into account using the Hartree–Fock (HF) approximation. The parameters in the model are determined on the basis of available first-principles calculations. For US, the calculated ratio Mo/Ms and magnetic circular dichroism spectrum for U M4,5 absorption are in good agreement with the experimental results. Inclusion of the expectation values of the spin-off-diagonal operators in addition to the number operators in the 5f state is found to be crucially important when describing the 5f magnetic state. A difference in enhancement of Mo of the Co atom between the Co/Pd and Co/Cu superlattices is discussed on the basis of a semi-quantitative calculation, assuming ferromagnetism.

1. Introduction

In magnets, the atomic spin (Ms) and orbital (Mo) magnetic moments are basic quantities and their separate determination is therefore important. In addition to the traditional method to measure Ms and Mo (Bonnenberg et al., 1986[Bonnenberg, D., Hempel, K. A. & Wijn, H. P. (1986). Magnetic Properties of 3d, 4d and 5d Elements, Alloys and Compounds, Landolt-Bornstein New Series, Vol. III/19a, edited by K.-H. Hellwege & O. Madelung. Berlin: Springer.]), the recently developed technique of X-ray magnetic circular dichroism (XMCD) with core-to-valence X-ray absorption (Chen et al., 1995[Chen, C. T., Izderda, Y. U., Lin, H.-J., Smith, N. V., Meig, G., Chaban, E., Ho, G. H., Pellegrin, E. & Sette, F. (1995). Phys. Rev. Lett. 75, 152-155.]) combined with several sum rules (Thole et al., 1992[Thole, B. T., Carra, P., Sette, F. & van der Laan, G. (1992). Phys. Rev. Lett. 68, 1943-1946.]; Carra et al., 1993[Carra, P., Thole, B. T., Altarelli, M. & Wang, X. (1993). Phys. Rev. Lett. 70, 694-697.]) has attracted much attention as a method for site- and symmetry-selective determination of Ms and Mo. From the theoretical point of view, the first-principles local-density approximation (LDA) of density-functional theory is a typical method for calculating magnetic quantities and has been successfully applied to various substances, especially 3d transition-metal systems (Morruzi et al., 1978[Morruzi, V. L., Janak, J. F. & Williams, A. R. (1978). Calculated Electronic Properties of Metals. New York: Pergamon.]).

In the LDA theory, the Kohn–Sham equation is described by the local potential including the spin-dependent electron density. The electric current, which describes Mo, is not, however, included. This means that there is no theoretical framework within which to ­determine Mo self-consistently in the LDA, and Mo is calculated as a quantity induced by Ms through the spin-orbit interaction (SOI). As a result, LDA theory is known to underestimate Mo in some cases. In bulk 3d transition metals, Mo is about one tenth of Ms (Bonnenberg et al., 1986[Bonnenberg, D., Hempel, K. A. & Wijn, H. P. (1986). Magnetic Properties of 3d, 4d and 5d Elements, Alloys and Compounds, Landolt-Bornstein New Series, Vol. III/19a, edited by K.-H. Hellwege & O. Madelung. Berlin: Springer.]) and Mo is sometimes neglected, i.e. the success of the first-­principles LDA approach relies on the experimental fact that Mo ≪ Ms. In 4f rare earths, Mo and Ms are mostly atom-like, i.e. determined by the Hund rule, and a solid-state effect is in many cases irrelevant. In 5f actinide systems, especially in uranium systems, Mo and Ms of the 5f state are determined by solid-state effects, i.e. they are dependent on the atomic environment. LDA theory applied to ferromagnetic U compounds underestimates Mo (Kraft et al., 1995[Kraft, T., Oppeneer, P. M., Antonov, V. N. & Eschrig, H. (1995). Phys. Rev. B, 52, 3561-3570.]).

Many theoretical attempts to improve the underestimation of Mo have been made. They may be roughly classified into two categories. An example of the first category is the so-called current-density-functional theory, which was formulated to extend the density-functional theory to include the orbital current, which describes Mo, as an extra degree of freedom (Vignal & Rasolt, 1987[Vignal, G. & Rasolt, M. (1987). Phys. Rev. Lett. 59, 2360-2363.]; Skudlarski & Vignal, 1993[Skudlarski, P. & Vignal, G. (1993). Phys. Rev. B, 48, 8547-8559.]; Higuchi & Hasegawa, 1997[Higuchi, H. & Hasegawa, A. (1997). J. Phys. Soc. Jpn, 66, 149-158.]). An explicit form of the contribution of the current density is, however, at present unknown and a tentative application of the theory to Co still underestimates Mo of the Co atom (Ebert et al., 1997[Ebert, H., Battocletti, M. & Gross, E. K. U. (1997). Europhys. Lett. 40, 545-550.]). The other category includes the so-­called orbital polarization (OP) (Brooks, 1985[Brooks, M. S. S. (1985). Physica B, 130, 6-12.]; Eriksson et al., 1990[Eriksson, O., Brooks, M. S. S. & Johansson, B. (1990). Phys. Rev. B, 41, 7311-7314.]) and LDA + U (U is an electron–electron interaction parameter) approaches (Solovyev et al., 1998[Solovyev, I. V., Lichtenstein, A. I. & Terakura, K. (1998). Phys. Rev. Lett. 80, 5758-5761.]), intended to calculate Mo ­practically. For problems with the OP approach, the reader is referred to the work of Solovyev et al. (1998[Solovyev, I. V., Lichtenstein, A. I. & Terakura, K. (1998). Phys. Rev. Lett. 80, 5758-5761.]) and Shishidou et al. (1999[Shishidou, T., Oguchi, T. & Jo, T. (1999). Phys. Rev. B, 59, 6813-6823.]). In the LDA + U approach, the wavefunctions are prepared by the LDA calculation, the multipole intra-atomic interaction between electrons is taken into account for the state of interest and the Hartree–Fock (HF) calculation is performed. The essential part of the LDA + U approach seems to be reproduced if an appropriate multi-orbital tight-binding Hamiltonian (Hubbard model) is prepared. In fact, for the antiferromagnet CoO, the HF calculations, with rotational invariance, based on both LDA + U (Solovyev et al., 1998[Solovyev, I. V., Lichtenstein, A. I. & Terakura, K. (1998). Phys. Rev. Lett. 80, 5758-5761.]) and an extended Hubbard model (Shishidou & Jo, 1998[Shishidou, T. & Jo, T. (1998). J. Phys. Soc. Jpn, 67, 2637-2640.]) are found to give similar results for both Mo of Co (∼1 μB) and of the stable magnetic structure.

The purpose of the present work is to review recent HF calculations of Ms and Mo made on the basis of an extended Hubbard model including the intra-atomic multipole interaction, for ferromagnetic US and superlattices of Co atoms in Pd or Cu matrixes. Parameters of the one-electron part of the model are determined on the basis of data available from first-principles LDA calculations. Among the Slater integrals describing the multipole interaction, the HF value of F0 for an atom assuming a suitable electron configuration is, in metals, reduced to a considerable extent by a screening effect, while that of Fk with k ≠ 0 is only reduced to 90∼70% (Norman, 1995[Norman, M. R. (1995). Phys. Rev. B, 52, 1421-1424.]). In the calculations described here, suitable reductions from the HF values for Fk with k ≠ 0 are assumed and values of F0 are determined so that the calculated total moment Ms + Mo agrees with the experimental result; the obtained ratio Mo/Ms is compared with the experimental value.

For US with NaCl structure, in addition to various magnetic measurements (Tillwick & de V. du Plessis, 1976[Tillwick, D. L. & de V. du Plessis, P. (1976). J. Magn. Magn. Mater. 3, 319-328.]; Lander et al., 1991[Lander, G. H., Brooks, M. S. S., Lebech, B., Brown, P. J., Vogt, O. & Mattenberger, K. (1991). J. Appl. Phys. 69, 4803-4806.]), XMCD has recently been observed (Collins et al., 1995[Collins, S. P., Laundy, D., Tang, C. C. & van der Laan, G. (1995). J. Phys. Condens. Matter, 7, 9325-9341.]). Through an HF calculation with rotational invariance, the role of the multipole exchange interaction described by the Gaunt coefficient (Condon & Shortley, 1959[Condon, E. U. & Shortley, G. H. (1959). Theory of Atomic Spectra. Cambridge University Press.]) is clearly seen to be important in the estimation of Mo, which is the case for atoms obeying the Hund rule. Since the establishment of the XMCD sum rules relating Ms and Mo, the enhancement of Mo of Co or Fe compared with the bulk metal has been reported in many XMCD experiments on magnetic multilayers (Wu et al., 1992[Wu, Y., Stöhr, J., Hamsmeiser, B., Samant, M. & Weller, W. (1992). Phys. Rev. Lett. 69, 2307-2310.]; Tischer et al., 1995[Tischer, M., Hjortstan, O., Arvanitis, D., Hunter Dunn, J., May, F., Baberschke, K., Trygg, J., Wills, J. M., Johansson, B. & Ekiksson, O. (1995). Phys. Rev. Lett. 75, 1602-1605.]; Nakajima et al., 1998[Nakajima, N., Koide, T., Shidara, T., Miyauchi, H., Fukutani, H., Fujimori, A., Iio, K., Katayama, T., Nyvlt, M. & Suzuki, Y. (1998). Phys. Rev. Lett. 81, 5229-5232.]). The reported enhancement factor of Mo is at most ∼2. For example, an enhancement of Mo of bulk hexagonal close packed (h.c.p.) Co from ∼0.15 μB to ∼0.25 μB has been reported, which is still much smaller than Ms of 1.6 μB. For Co or Fe impurities in Cs metals, on the other hand, a recent LDA + U calculation (Kwon & Min, 2000[Kwon, S. K. & Min, B. I. (2000). Phys. Rev. Lett. 84, 3970-3973.]) has predicted a value for Mo of ∼3 μB, i.e. almost atom-like, as for rare earths. In the present work, Mo of Co atoms in a Pd or Cu matrix is calculated assuming ferromagnetism. The relation between Mo and the atomic environment is discussed and it is shown that Mo of Co can be comparable to Ms if Co is surrounded by Cu atoms as nearest neighbours.

2. Model

We assume the Hamiltonian

[H = H_1 + H_2,]

where H1 denotes the multipole dd or ff interaction, and the spin-orbit interaction in d or f orbit is given by

[\eqalign{H_1 =\hskip.2em& \textstyle\sum\limits_{i,\nu_1,\nu_2,\nu_3,\nu_4}\langle i\nu_1,i\nu_2| (1 / r_{12})| i\nu_3,i\nu_4\rangle a^+_{i\nu_1}a^+_{i\nu_2}a_{i\nu_4}a_{i\nu_3} \cr &\!+\textstyle\sum\limits_{i,\nu_1,\nu_2}\langle i\nu_1|\zeta l \, s| i\nu_2\rangle a^+_{i\nu_1}a_{i\nu_2} ,}]

with i and ν being the lattice point and the combined index of the magnetic and spin quantum numbers of the d or f orbit, m and σ, respectively; ζ denotes the spin-orbit coupling constant. The matrix element of the dd or ff interaction is expressed in terms of the Slater integral Fk and the Gaunt coefficient ck(lm, lm′), with l = 2 and k = 0, 2 and 4 for the d orbit, and l = 3 and k = 0, 2, 4 and 6 for the f orbit, as follows (Condon & Shortley, 1959[Condon, E. U. & Shortley, G. H. (1959). Theory of Atomic Spectra. Cambridge University Press.]):

[\eqalign{\langle \nu_1\nu_2| (1/r_{12})| \nu_3 \nu_4\rangle = \hskip.2em& \delta_{\sigma_1\sigma_3}\delta_{\sigma_2\sigma_4} \delta_{m_1+m_2,m_3+m_4}\cr &\!\times\textstyle\sum\limits_{k}F^kc^k(l m_1,l m_3)c^k(l m_4,l m_2).}]

H2 denotes the interatomic electron transfer and the atomic level,

[H_2 = \textstyle\sum\limits_{i,j}\textstyle\sum\limits_{\mu_1,\mu_2}t^{\mu_1,\mu_2}_{ij}a^+_{i\mu_1}a_{j\mu_2},]

with μ being the combined index of the magnetic and spin quantum numbers of all the orbits including d and f.

For the multipole dd or ff interaction in H1, we apply the HF approximation: the first term of H1 is replaced by

[\eqalign{&\textstyle\sum\limits_{i,\nu_1,\nu_2,\nu_3,\nu_4}\big[\langle i\nu_1,i\nu_2| (1/r_{12})| i\nu_3,i\nu_4\rangle \cr&\quad - \langle i\nu_1,i\nu_2| (1/r_{12})| i\nu_4,i\nu_3\rangle\big]\langle a^+_{i\nu_2}a_{i\nu_4}\rangle a^+_{i\nu_1}a_{i\nu_3},}]

with [\langle a^+_{i\nu_2}a_{i\nu_4}\rangle] being the expectation value of the operator including the number operator. By diagonalizing the obtained one-body Hamiltonian, we calculate Ms and Mo:

[M_s = \mu_B\textstyle\sum\limits_m(\langle a^+_{m\downarrow}a_{m\downarrow}\rangle - \langle a^+_{m\uparrow}a_{m\uparrow}\rangle)]

and

[M_o = \mu_B\textstyle\sum\limits_{m,\sigma}m\langle a^+_{m\sigma}a_{m\sigma}\rangle .]

The lattice point index i is omitted for simplicity.

3. 5f state and M4,5 XMCD in US

Taking into account the U 5f, 6p, 6d and 7s orbits, and the S 3s, 3p and 3d orbits, the parameters in H2 are determined by a first-principles LDA calculation with the full-potential linear augmented plane wave (FLAPW) method in the paramagnetic state without the 5f spin-orbit interaction. F0 = 0.76, F2 = 5.530, F4 = 4.669 and F6 = 2.881 eV are adopted in the Hamiltonian H1. For details, the reader is referred to the work of Shishidou et al. (1999[Shishidou, T., Oguchi, T. & Jo, T. (1999). Phys. Rev. B, 59, 6813-6823.]), including the F0 dependence of Ms and Mo. As a result of the presence of a large Mo, magnetic anisotropy exits. Values of M5f (= Ms + Mo) of 1.70 for the [111] direction, 1.51 for the [110] direction and 1.28 for the [001] direction are obtained, in units of μB. By comparing the total energies among the three directions, the [111] direction is found to be the easy axis in accordance with the experiment and hereinafter the magnetic moments are presented along the easy axis.

In Table 1[link], the calculated M5f, Ms and Mo values obtained by several methods are compared. The results based on the conventional LDA with SOI reveal an absolute value of M5f that is too small compared with the experimental value of 1.70 μB (Wedgwood, 1972[Wedgwood, F. A. (1972). J. Phys. C, 5, 2427-2444.]) because of an underestimation of Mo. Brooks (1985[Brooks, M. S. S. (1985). Physica B, 130, 6-12.]) applied the OP method and obtained a larger magnitude of Mo and a considerable improvement in M5f. According to Severin et al. (1993[Severin L., Brooks, M. S. S. & Johansson, B. (1993). Phys. Rev. Lett. 71, 3214-3217.]), the individual absolute magnitudes of Ms and Mo of Brooks are too large compared with those obtained from the analysis of the magnetic form factor. Severin et al. (1993[Severin L., Brooks, M. S. S. & Johansson, B. (1993). Phys. Rev. Lett. 71, 3214-3217.]) performed an HF calculation in which the expectation values of only the number operators specified by the spin (σ) and magnetic quantum numbers (m) are taken into account and the exchange integrals are scaled to the LDA (`scaled HF' method). Their result is, as far as Ms and Mo are concerned, similar to ours (`HF TB') with Ms = −1.49 μB and Mo = 3.19 μB, where the off-diagonal operator as well as the number operators are taken into account. In order to estimate Mo well, one should include the orbital-dependent exchange potential, which is taken into account in neither the LDA nor the OP method. The values in parentheses (Table 1[link]) of the neutron measurement by Wedgwood (1972[Wedgwood, F. A. (1972). J. Phys. C, 5, 2427-2444.]) are from an analysis by Severin et al. (1993[Severin L., Brooks, M. S. S. & Johansson, B. (1993). Phys. Rev. Lett. 71, 3214-3217.]). The calculation was also performed neglecting the spin-off-diagonal operators in the present model (`spin-diagonal HF'); the obtained Ms and Mo are apparently similar to those obtained by the HF TB method. Problems with the scaled HF and the spin-diagonal HF method will be discussed below.

Table 1
Comparison of calculated 5f magnetic moments in uranium sulfide

LDA = local-density approximation. SOI = spin-orbit interaction. OP = orbital polarization. HF = Hartree–Fock. FLAPW = full-potential linear augmented plane wave. ASW = augmented sphere wave. LMTO = linearized muffin-tin orbital. TB = tight binding.

Method Reference M5f (μB) Ms (μB) Mo (μB)
LDA + SOI FLAPW Oguchi (1998[Oguchi, T. (1998). Unpublished.]) 0.55 −1.66 2.21
LDA + SOI ASW Kraft et al. (1995[Kraft, T., Oppeneer, P. M., Antonov, V. N. & Eschrig, H. (1995). Phys. Rev. B, 52, 3561-3570.]) 1.1 −1.5 2.6
LDA + SOI LMTO Brooks (1985[Brooks, M. S. S. (1985). Physica B, 130, 6-12.]) 1.1 −2.1 3.2
OP LMTO Brooks (1985[Brooks, M. S. S. (1985). Physica B, 130, 6-12.]) 1.8 −2.2 4.0
OP (scaled HF) Severin et al. (1993[Severin L., Brooks, M. S. S. & Johansson, B. (1993). Phys. Rev. Lett. 71, 3214-3217.]) 1.61 −1.51 3.12
HF TB This work 1.70 −1.49 3.19
Spin-diagonal HF TB This work 1.56 −1.78 3.34
Neutron measurement Wedgwood (1972[Wedgwood, F. A. (1972). J. Phys. C, 5, 2427-2444.]) 1.70 (−1.31) (3.0)
†This value is adjusted to the experimental value.

Fig. 1[link] presents the U 3d → 5f X-ray absorption spectroscopy (XAS) XMCD spectrum ΔF(ω) = F(ω) − F+(ω), calculated with use of the unoccupied Bloch states and energy eigenvalues obtained for HF TB on the basis of Fermi's `Golden rule'; F± represents the XAS spectrum for the incident positive (+) and negative (−) helicities. The Lorentzian convolution with FWHM of 4.0 eV is adopted, which represents the U 3d core-hole lifetime broadening. The calculated XMCD spectral shape is found to be in good agreement with the measurement by Collins et al. (1995[Collins, S. P., Laundy, D., Tang, C. C. & van der Laan, G. (1995). J. Phys. Condens. Matter, 7, 9325-9341.]), with dispersive features in the M5 region. The M5 to M4 intensity ratio RXMCD, determined by combining the so-called Lz (Thole et al., 1992[Thole, B. T., Carra, P., Sette, F. & van der Laan, G. (1992). Phys. Rev. Lett. 68, 1943-1946.]) and Sz (Carra et al., 1993[Carra, P., Thole, B. T., Altarelli, M. & Wang, X. (1993). Phys. Rev. Lett. 70, 694-697.]) XMCD sum rules, is expressed as

[\eqalign{R_{\rm XMCD} = \hskip.2em&\textstyle\int\limits_{M_5}\Delta F(\omega)\,{\rm d}\omega\big / \textstyle\int\limits_{M_4}\Delta F(\omega)\,{\rm d}\omega \cr =\hskip.2em& (5/2)[ \langle L_z\rangle / (\langle L_z\rangle - 2\langle S_z\rangle -6\langle T_z\rangle )]- 1,}]

where 〈Lz〉 = Mo/μB and 〈Sz〉 = Ms/(2μB), and 〈Tz〉 is the expectation value of the z component of the magnetic dipole operator, given by

[{\bf T} = \textstyle\sum\limits_i[{\bf s}_i - 3{\bf r}_i({\bf r}_i\cdot{\bf s}_i)/r_i^2].]

si and ri are the spin and the position vectors of the ith 5f electron, respectively. The HF TB method, taking into account the expectation value of the off-diagonal operators as well as the number operators, gives 〈Tz〉 = −0.36 and RXMCD = 0.169, which is in reasonable agreement with the observed value of 0.13 ± 0.03 (Collins et al., 1995[Collins, S. P., Laundy, D., Tang, C. C. & van der Laan, G. (1995). J. Phys. Condens. Matter, 7, 9325-9341.]). The spin-diagonal HF TB method, on the other hand, gives 〈Tz〉 = −0.22 (∼60% of the HF TB value) and RXMCD = 0.292. The magnitudes of Ms and Mo (or 〈Sz〉 and 〈Lz〉) are rather insensitive to whether the spin-off-diagonal operators are included or not, since Lz and Sz are expressed by the number operators specified by m and σ. Tz is, on the other hand, expressed in terms of spin-off-diagonal operators as well as number operators and its expectation value. Furthermore, RXMCD is more sensitively influenced by the extent of spin-off-diagonal mixing in the Bloch wavefunction. It has been shown that RXMCD can be a severe test of wavefunction and that the HF TB method is a promising method to calculate magnetic quantities including wave functions.

[Figure 1]
Figure 1
Calculated isotropic (thin line) and magnetic circular dichroism (thick line) spectra for U M4,5 absorption in US as a fuction of relative photon energy (Shishidou et al., 1999[Shishidou, T., Oguchi, T. & Jo, T. (1999). Phys. Rev. B, 59, 6813-6823.]).

4. Co/Pd and Co/Cu superlattices

In this section, the relation between Mo or Mo/Ms of transition-metal atoms and various atomic environments will be discussed. For this purpose, superlattices of Co atoms in face-centred cubic (f.c.c.) Pd or Cu hosts have been chosen, as shown in Fig. 2[link], for which Ms and Mo of Co will be calculated. In Fig. 2[link](a), linear chains of Co atoms (solid circles) are periodically embedded in the [110] direction in a three-dimensional Pd or Cu (open circles) matrix. In Fig. 2[link](b), Co chains are in the [100] direction. In Fig. 2[link](c), two-dimensional sheets of Co atoms are stacked in the [001] direction, i.e. forming a kind of multilayer. In Fig. 2[link](d), Co atoms form a three-dimensional superlattice. In Fig. 2[link], the lattice constant of the pure host metal, Pd or Cu, is assumed. The parameters in H2 for pure metals, where s, p and d orbits are taken into account, are determined according to Papaconstantopoulos (1986[Papaconstantopoulos, D. (1986). Handbook of the Band Structure of Elemental Solids. New York: Plenum.]) in the paramagnetic state; these parameters are used in the calculation in the ferromagnetic state. The Co 3d majority spin states are almost filled; thus the neglect of the spin-off-diagonal operators, as discussed in the preceding section, is found to be a good approximation. The calculation for f.c.c. Co gives Ms = 1.52 μB and Mo = 0.11 μB, and that for h.c.p. Co gives Ms = 1.58 μB and Mo = 0.13 μB. The parameters in the superlattices are estimated according to Andersen et al. (1978[Andersen, O. K., Close, W. & Nohl, H. (1978). Phys. Rev. B, 17, 1209-1237.]) and the Co–Pd (Cu) transfer integrals are assumed to be the arithmetic means of the Co–Co and Pd–Pd (Cu–Cu) values. For details of the calculations, the reader is referred to the work of Okutani & Jo (2000[Okutani, M. & Jo, T. (2000). J. Phys. Soc. Jpn, 69, 598-606.]).

[Figure 2]
Figure 2
The superlattices composed of Co atoms (solid circles) and Pd or Cu host atoms (open circles) (Okutani & Jo, 2000[Okutani, M. & Jo, T. (2000). J. Phys. Soc. Jpn, 69, 598-606.]).

In Table 2[link], the calculated Mo and Ms of Co for the superlattices shown in Fig. 2[link] are presented along with the choice of quantization axes. The number of 3d electrons is, compared with the bulk Co, increased by ∼0.2 in the model superlattices and Ms is decreased by ∼0.2 μB. The following observations regarding Ms and Mo of Co atoms can be made.

  • (i) Ms is insensitive to the atomic environment and the quantization axis.

  • (ii) Mo is enhanced compared with bulk Co metal in any case.

  • (iii) The enhancement of Mo is much more remarkable in the Cu matrix.

  • (iv) The extent of enhancement in the Pd matrix is not so sensitive to the atomic environment (0.24 μB < Mo < 0.35 μB).

  • (v) The extent of enhancement in the Cu matrix is strongly dependent on the atomic environment. The enhancement is remarkable in the case where a Co atom is surrounded by Cu atoms as its nearest neighbours.

  • (vi) Mo is dependent on the quantization axis, i.e. anisotropy exists and is more remarkable in the Cu matrix.

Table 2
The calculated atomic spin (Ms) and orbital moments (Mo) of Co in the Pd and Cu matrixes shown in Fig. 2[link]

(a) Co in Pd matrix.

Lattice Ms (μB) Mo (μB) Mo/Ms
Fig. 2[link](a) || 1.37 0.30 0.22
  ⊥1 1.36 0.26 0.19
  ⊥2 1.37 0.27 0.20
Fig. 2[link](b) || 1.37 0.24 0.17
  ⊥1 1.36 0.34 0.25
  ⊥2 1.38 0.35 0.26
Fig. 2[link](c) 1.36 0.25 0.19
  ||1 1.37 0.24 0.18
  ||2 1.37 0.24 0.17
Fig. 2[link](d) d1 1.34 0.33 0.25
  d2 1.35 0.34 0.25

(b) Co in Cu matrix.

Lattice Ms (μB) Mo (μB) Mo/Ms
Fig. 2[link](a) || 1.38 0.38 0.28
  ⊥1 1.38 0.63 0.46
  ⊥2 1.39 0.67 0.48
Fig. 2[link](b) || 1.33 1.43 1.07
  ⊥1 1.35 1.10 0.81
  ⊥2 1.35 1.48 1.09
Fig. 2[link](c) 1.45 0.69 0.47
  ||1 1.45 0.40 0.28
  ||2 1.45 0.32 0.22
Fig. 2[link](d) d1 1.34 1.22 0.91
  d2 1.34 1.45 1.09

Result (i) arises from the fact that the Co majority spin state is almost filled and the number of minority-spin 3d electrons determines Ms. Results (ii), (iii), (iv) and (v) provide a picture of the environment dependence of Mo in the Pd and Cu matrixes. In the case of the Pd matrix, the lattice parameter is considerably larger compared with that of bulk Co, and pure Co with such a lattice constant has Mo of 0.36 μB as a result of a narrowing of the 3d state in the present model. A comparison between this value and those in Table 2[link](a) is suggestive. The effective Co–Co hybridization and the Co–Pd hybridization are similar to each other. The key factor in the enhancement of Mo is, in this case, the lattice parameter, while the atomic environment dependence plays a minor role. The magnitude of 0.36 μB can even be reduced a little by surrounding Pd atoms. In the case of the Cu matrix, on the other hand, Mo of pure Co enlarged to the Cu lattice parameter is only 0.16 μB, which is much smaller than the values given in Table 2[link](b). Since the effective Co–Cu hybridization is much smaller than the Co–Co hybridization in the Cu matrix and also than in pure Co metal, the neighbouring Cu atoms cause the enhancement of Mo of Co through a narrowing of the 3d state. This seems to be sensitively reflected in the strong environment dependence of Mo: in the multilayer (see Fig. 2[link]c), a Co atom is surrounded by four Co atoms and Mo is somewhat smaller (0.7∼0.4 μB) compared with the case in which the Co is surrounded by no Co atoms (Figs. 2[link]b and 2[link]d), where Mo is comparable to Ms.

The contribution to the magnetic anisotropy energy (MAE) (ΔE) is known to be composed of differences in the band energies (ΔEb) and the magnetic dipole–dipole interaction energies (ΔEdd) between two orientations (Szunyogh et al., 1995[Szunyogh, L., Újfalussy, B. & Weinberger, P. (1995). Phys. Rev. B, 51, 9552-9560.]). The band energy is lower for the magnetization direction with larger Mo; Ms is, in the present case, insensitive to the direction. The calculated result for the Co/Cu multilayer [see Table 2[link](b) for the lattice shown in Fig. 2[link](c)] shows that Mo is much larger for the direction perpendicular to the Co layer compared with that for the parallel direction. If it is assumed that Cu atoms in the Co/Cu multilayers play a role similar to that of Au atoms in Co/Au multilayers and that the main contribution to ΔE is ΔEb in Co/Cu, the result is qualitatively consistent with that of the first-principles calculation for Co/Au performed by Újfalussy et al. (1996[Újfalussy, B., Szunyogh, L., Bruno, P. & Weinberger, P. (1996). Phys. Rev. Lett. 77, 1805-1808.]).

Recently, enhancements of Mo, determined by XMCD measurements, have been reported for Co/Pd (Wu et al., 1992[Wu, Y., Stöhr, J., Hamsmeiser, B., Samant, M. & Weller, W. (1992). Phys. Rev. Lett. 69, 2307-2310.]), Co/Cu (Tischer et al., 1995[Tischer, M., Hjortstan, O., Arvanitis, D., Hunter Dunn, J., May, F., Baberschke, K., Trygg, J., Wills, J. M., Johansson, B. & Ekiksson, O. (1995). Phys. Rev. Lett. 75, 1602-1605.]) and Co/Pt multilayer systems (Nakajima et al., 1998[Nakajima, N., Koide, T., Shidara, T., Miyauchi, H., Fukutani, H., Fujimori, A., Iio, K., Katayama, T., Nyvlt, M. & Suzuki, Y. (1998). Phys. Rev. Lett. 81, 5229-5232.]). In these cases, and in a recent study of nanoscale Fe clusters (Edmonds et al., 1999[Edmonds, K. W., Binns, C., Baker, S. H., Thornton, S. C., Norris, C., Goedkoop, J. B., Finazzi M. & Brookes, N. B. (1999). Phys. Rev. B, 60, 472-476.]), the reported enhancement factor, compared with the bulk metals, is at most double. If the monolayer sandwich of the present study is modified to two monolayers in the calculation, Mo of Co is considerably reduced. For a quantitative discussion, a confirmation of the atomic structure, between theory and experiment, will be needed.

5. Concluding remarks

In summary, Ms and Mo of the U 5f state in US and of the Co 3d state in various Co/Pd and Co/Cu superlattices have been calculated with an HF approximation on the basis of a tight-binding model, including the full atomic orbitals in valence states and the multipole interactions between 5f or d electrons. In US, it is stressed that inclusion of the expectation value of the spin-off-diagonal operators in addition to the number operators is crucially important when describing the Bloch state via an analysis of the U M4,5 XMCD spectrum. Calculations assuming ferromagnetism reveal an enhancement of Mo of Co in both the Co/Pd system and the Co/Cu system compared with bulk Co. The enhancement in the Co/Pd system arises from the large lattice constant in the Pd matrix, while that in Co/Cu results from the small hybridization of the Co–Cu pair compared with the Co–Co pair.

The magnitudes of Mo and MAE, i.e. the perpendicular or parallel anisotropy in the magnetic multilayers, are subjects closely related to each other. At present, the first-principles calculation of MAE has been performed only for limited systems, although the 3d-electron number dependence (of magnetic atoms) of MAE has been discussed on the basis of simplified models (see for example Dorantes-Dávia & Pastor, 1998[Dorantes-Dávila, J. & Pastor, G. M. (1998). Phys. Rev. Lett. 81, 208-211.]). It is expected that in the future these first-principles calculations will be developed further and that realistic calculations of MAE in various systems, including magnetic multilayers, will be made.

In itinerant 3d ferromagnets, the thermodynamical properties, e.g. the temperature dependence of magnetization, have been the main subject of theoretical studies (Moriya, 1985[Moriya, T. (1985). Spin Fluctuations in Itinerant Electron Magnetism. Berlin: Springer.]). So far, the temperature dependence of only Ms has been discussed; the temperature dependence of Mo has been neglected. In systems with enhanced Mo, the temperature dependence not only of Ms but also of Mo (or Mo/Ms) should be the subject of interest, especially as the latter may provide new information regarding electronic structure. Recent measurements of Mo in Co clusters on Au(111) (Dürr et al., 1999[Dürr, H. A., Dhesi, S. S., Dudzik, E, Knabben, D, van der Laan, G., Goedkoop, J. B. & Hillebrecht, F. U. (1999). Phys. Rev. B, 59, R701-R704.]) have addressed this subject from an experimental point of view.

Acknowledgements

The author thanks T. Oguchi, A. Tanaka, T. Shishidou and M. Okutani for discussions and collaboration. This work is partially supported by a Grant-in-Aid for Scientific Research from the Ministry of Education, Science, Sports and Culture.

References

First citationAndersen, O. K., Close, W. & Nohl, H. (1978). Phys. Rev. B, 17, 1209–1237. CrossRef CAS Web of Science
First citationBonnenberg, D., Hempel, K. A. & Wijn, H. P. (1986). Magnetic Properties of 3d, 4d and 5d Elements, Alloys and Compounds, Landolt–Bornstein New Series, Vol. III/19a, edited by K.-H. Hellwege & O. Madelung. Berlin: Springer.
First citationBrooks, M. S. S. (1985). Physica B, 130, 6–12. CrossRef CAS
First citationCarra, P., Thole, B. T., Altarelli, M. & Wang, X. (1993). Phys. Rev. Lett. 70, 694–697. CrossRef PubMed CAS Web of Science
First citationChen, C. T., Izderda, Y. U., Lin, H.-J., Smith, N. V., Meig, G., Chaban, E., Ho, G. H., Pellegrin, E. & Sette, F. (1995). Phys. Rev. Lett. 75, 152–155. CrossRef PubMed CAS Web of Science
First citationCollins, S. P., Laundy, D., Tang, C. C. & van der Laan, G. (1995). J. Phys. Condens. Matter, 7, 9325–9341. CrossRef CAS Web of Science
First citationCondon, E. U. & Shortley, G. H. (1959). Theory of Atomic Spectra. Cambridge University Press.
First citationDorantes-Dávila, J. & Pastor, G. M. (1998). Phys. Rev. Lett. 81, 208–211. Web of Science CrossRef
First citationDürr, H. A., Dhesi, S. S., Dudzik, E, Knabben, D, van der Laan, G., Goedkoop, J. B. & Hillebrecht, F. U. (1999). Phys. Rev. B, 59, R701–R704. Web of Science CrossRef
First citationEbert, H., Battocletti, M. & Gross, E. K. U. (1997). Europhys. Lett. 40, 545–550. CrossRef CAS Web of Science
First citationEdmonds, K. W., Binns, C., Baker, S. H., Thornton, S. C., Norris, C., Goedkoop, J. B., Finazzi M. & Brookes, N. B. (1999). Phys. Rev. B, 60, 472–476.
First citationEriksson, O., Brooks, M. S. S. & Johansson, B. (1990). Phys. Rev. B, 41, 7311–7314. CrossRef CAS Web of Science
First citationHiguchi, H. & Hasegawa, A. (1997). J. Phys. Soc. Jpn, 66, 149–158. CrossRef CAS Web of Science
First citationKraft, T., Oppeneer, P. M., Antonov, V. N. & Eschrig, H. (1995). Phys. Rev. B, 52, 3561–3570. CrossRef CAS Web of Science
First citationKwon, S. K. & Min, B. I. (2000). Phys. Rev. Lett. 84, 3970–3973. Web of Science CrossRef PubMed CAS
First citationLander, G. H., Brooks, M. S. S., Lebech, B., Brown, P. J., Vogt, O. & Mattenberger, K. (1991). J. Appl. Phys. 69, 4803–4806. CrossRef CAS Web of Science
First citationMoriya, T. (1985). Spin Fluctuations in Itinerant Electron Magnetism. Berlin: Springer.
First citationMorruzi, V. L., Janak, J. F. & Williams, A. R. (1978). Calculated Electronic Properties of Metals. New York: Pergamon.
First citationNakajima, N., Koide, T., Shidara, T., Miyauchi, H., Fukutani, H., Fujimori, A., Iio, K., Katayama, T., Nyvlt, M. & Suzuki, Y. (1998). Phys. Rev. Lett. 81, 5229–5232. Web of Science CrossRef CAS
First citationNorman, M. R. (1995). Phys. Rev. B, 52, 1421–1424. CrossRef CAS Web of Science
First citationOguchi, T. (1998). Unpublished.
First citationOkutani, M. & Jo, T. (2000). J. Phys. Soc. Jpn, 69, 598–606. Web of Science CrossRef CAS
First citationPapaconstantopoulos, D. (1986). Handbook of the Band Structure of Elemental Solids. New York: Plenum.
First citationSeverin L., Brooks, M. S. S. & Johansson, B. (1993). Phys. Rev. Lett. 71, 3214–3217.
First citationShishidou, T. & Jo, T. (1998). J. Phys. Soc. Jpn, 67, 2637–2640. Web of Science CrossRef CAS
First citationShishidou, T., Oguchi, T. & Jo, T. (1999). Phys. Rev. B, 59, 6813–6823. Web of Science CrossRef CAS
First citationSkudlarski, P. & Vignal, G. (1993). Phys. Rev. B, 48, 8547–8559. CrossRef CAS Web of Science
First citationSolovyev, I. V., Lichtenstein, A. I. & Terakura, K. (1998). Phys. Rev. Lett. 80, 5758–5761. Web of Science CrossRef CAS
First citationSzunyogh, L., Újfalussy, B. & Weinberger, P. (1995). Phys. Rev. B, 51, 9552–9560. CrossRef CAS Web of Science
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
First citationTillwick, D. L. & de V. du Plessis, P. (1976). J. Magn. Magn. Mater. 3, 319–328.
First citationTischer, M., Hjortstan, O., Arvanitis, D., Hunter Dunn, J., May, F., Baberschke, K., Trygg, J., Wills, J. M., Johansson, B. & Ekiksson, O. (1995). Phys. Rev. Lett. 75, 1602–1605. CrossRef PubMed CAS Web of Science
First citationÚjfalussy, B., Szunyogh, L., Bruno, P. & Weinberger, P. (1996). Phys. Rev. Lett. 77, 1805–1808. CrossRef PubMed Web of Science
First citationVignal, G. & Rasolt, M. (1987). Phys. Rev. Lett. 59, 2360–2363. CrossRef PubMed Web of Science
First citationWedgwood, F. A. (1972). J. Phys. C, 5, 2427–2444.  CrossRef Web of Science
First citationWu, Y., Stöhr, J., Hamsmeiser, B., Samant, M. & Weller, W. (1992). Phys. Rev. Lett. 69, 2307–2310. CrossRef PubMed CAS Web of Science

© 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