arXiv:cond-mat/0209326v1 [cond-mat.mtrl-sci] 13 Sep 2002

Electron Momentum Density in Cu0.9Al0.1

M. Samsel-Czeka³a1, G. Kontrym-Sznajd1, G. Döring2, W. Schülke2, J. Kwiatkowska3, F. Maniawski3, S. Kaprzyk4,5 and A. Bansil5

1 Institute of Low Temperature and Structure Research, Polish Academy of Sciences, .O.Box 1410, 50-950 Wroc³aw 2, Poland

2 Institute of Physics, University of Dortmund, D-44221 Dortmund, Germany

3 The Henryk Niewodniczañski Institute of Nuclear Physics, Radzikowskiego 152, 31-342 Kraków, Poland

4 Faculty of Physics and Nuclear Techniques, Academy of Mining and Metallurgy, al. Mickiewicza 30, 30-073 Kraków, Poland

5 Department of Physics, Northeastern University, Boston, Massachusetts 02115, USA

Published in: Appl. Phys. A (2002)

Abstract

A reconstruction technique based on the solution of the Radon transform in terms of Jacobi polynomials is used to obtain the 3D electron momentum density r(p) from nine high-resolution Compton profiles (CPs) for a Cu0.9Al0.1 disordered alloy single crystal. The method was also applied to theoretical CPs computed within the Korringa-Kohn-Rostoker coherent potential approximation (KKR-CPA) first-principles scheme for the same nine orientations of the crystal. The experimental density r(p) is in satisfactory agreement with the theoretical density and shows most details of the Fermi surface (FS) and exhibits electron correlation effects. We comment on the map of the FS obtained by folding the reconstructed r(p) into the first Brillouin zone which yields the occupation number density, r(k). A test of the validity of data via a consistency condition (within our reconstruction algorithm) as well as the propagation of experimental noise in the reconstruction of both r(p) and r(k) are investigated.

PACS: 74.25.Jb, 71.23.-k, 13.60.Fz, 71.18.+y, 42.30.Wb, 61.66.Dk

1. Introduction

The electron momentum density, r(p), can be probed via Compton scattering experiments in which measured Compton profiles (CPs)

(1)

provide plane projections of r(p) along the direction of the scattering vector assumed to be parallel to the pz-axis.

The 3D momentum density r(p) may then be reconstructed by measuring CPs along various directions of the scattering vector. Although this problem has a long history, the recent possibility of obtaining high resolution CPs in wide classes of materials has rejuvenated interest in reconstruction techniques as a way of getting a handle on FS signatures and electron correlation effects in the underlying momentum density [1]. With this motivation, we have carried out extensive CP measurements on a Cu0.9Al0.1 alloy single crystal (fcc solid solution phase), and analyzed the results in terms of parallel KKR-CPA computations [2,3].

Furthermore, we have reconstructed the 3D momentum density using the technique of Ref. [4] which is based on the inversion of the Radon transforms in terms of spherical harmonics and Jacobi polynomials. r(p) and J(pz) are first expanded into lattice harmonics Fln(Q,j):

, (2)

(3)

where the index n distinguishes harmonics of the same order l and (b ,a ) describe the polar and azimuthal angles of the pz-axis with respect to the reciprocal lattice. The radial components of the measured profiles are now expanded in terms of Jacobi polynomials Pk(a,b):

. (4)

The radial parts of the momentum density are then given by

. (5)

Our technique has the advantage that the Fourier transforms involved in the standard approaches [5] are circumvented and that the experimental errors (due to statistics) are taken into account more naturally due to the expansion of the measured profiles into orthogonal polynomials (4) which have

least-squares approximation properties [6]. Moreover, such an expansion allows us to check the validity of the data via a consistency condition.

In order to visualize the shape of the FS we have performed an LCW-folding [7] of r(p) to obtain r(k) in the reduced momentum space k. We also consider the propagation of simulated noise in the reconstructed densities r(p) and r(k) within the applied reconstruction algorithm.

2. Compton profiles

The samples of Cu0.9Al0.1 were prepared as follows. A single crystal of the alloy was grown from the melt by a modified Bridgman method. A split graphite crucible containing an oriented crystalline seed at its bottom was used, allowing the growth of a single crystal of the required orientation. The orientation was checked by the Laue X-ray diffraction method and the crystal was cut into slices with surface normals along the [100], [110] and [111] directions, using a diamond saw. The samples were then ground with fine abrasive powders and chemically polished to remove the damaged surface.

For the samples thus prepared, high-resolution CPs were measured at the beamline ID 15b of the ESRF (Grenoble, France) for 9 directions of the momentum transfer vector pz, shown in Fig. 1 and given by the angles (Q,j): (900,00)º [100]; (900,100); (900,200); (900,450)º [110]; (800,450); (54.740,450)º [111]; (63.530,39.430); (72.500,34.670); (81.580,30.360).

Fig. 1 Directions pz along which CPs in Cu0.9Al0.1 were measured and computed are marked in the non-equivalent part of the cubic Brillouin zone.

Utilizing a bent Si(311) single crystal monochromator, a scanning type Rowland spectrometer with a Ge(440) analyzer crystal and a NaI detector, we collected 2.5´ 107 counts in each CP corresponding to 2.5´ 105 counts at the maximum of the peak. The momentum space resolution was 0.2 a.u. at pz=0, which was determined by measuring the width of the quasi-elastic line and correcting for the known energy dependence of the spectrometer resolution [8] and the distribution of the scattering angles due to the finite horizontal angular acceptance of the analyzer. To extract the valence CPs, the measured profiles were corrected for the following energy dependent effects: air absorption along the flight path of the scattered photons, self-absorption in the scattering sample, reflectivity and vertical acceptance of the analyzer crystal, a scale correction and the energy dependent relation between the measured cross section and the CP [9,10]. Furthermore, the contribution of the core electrons to the CP, calculated utilizing Hartree-Fock wave functions [11], was subtracted. Multiple scattering amounting to about 10 percent of the total profile was taken into account by a Monte-Carlo simulation [12] and also subtracted. Since this simulation makes some simplifying assumptions with respect to the shape of the sample and the propagation of polarization, it cannot be completely ruled out that, at least for certain sample orientations, small systematic errors are introduced. Finally, the valence electron CPs were normalized to the number of valence electrons Nval utilizing

(6)

Theoretical CPs (for the pz directions marked in Fig. 1) were computed within the fully selfconsistent KKR-CPA framework [2,3] for the same crystal orientations. The alloy electronic structure problem was solved to a high degree of self-consistency for the fcc lattice, using a lattice constant a=6.860 a.u., obtained by minimizing the total energy. The KKR-CPA self-consistency cycles were carried out in the complex energy plane using elliptic contours with 48 energy points. Using the converged Cu and Al crystal potentials the momentum density <r(p)>, averaged over configurations of the disordered alloy, was calculated at 1291 p points obtained by adding the same number of reciprocal lattice vectors to each of the 1183 k points in the irreducible 1/48th of the Brillouin zone [13]. Thus, the total number of p points defining r(p) was 48´ 1291´ 1183. Subsequently, the required CPs were computed following (1).

As pointed out in Ref. [15], all measured CPs must be interdependent as they are plane projections of the same density. This leads to the so-called consistency condition (CC), which the data have to satisfy and which was studied in [16-19] for line projections. Experimental profiles must be consistent, i.e. the inconsistent part of each profile should not exceed the experimental noise, or else the profile might be contaminated by an extra systematic error [20]. The CC is automatically imposed on the experimental data via our reconstruction technique [4] by the condition that the lowest polynomial for each gln(p) (see (4)) is of the order l.

The CC can be used for checking the quality of the data before reconstructing r(p) [20]. The procedure to do this is the following. Each experimental profile J(pz) º J(b ,a )(p) is expanded into a series of even Jacobi polynomials:

. (7)

Next, we create functions gln(p) given by

, (8)

where the ãlnk are obtained by expanding the ak(b ,a ) from (7) into cubic harmonics:

. (9)

In order to obtain equation (4), the first l/2 coefficients ãlnk must be equal to zero (consistency condition).

Some results of applying this procedure are displayed in Fig. 2, where two different experimental profiles J(pz) are approximated by corresponding J* (pz) which are constructed from six gln(p), each consisting of a series of 60 Jacobi polynomials:

. (10)

Fig. 2 Results of applying the consistency condition (CC) and the Jacobi polynomial expansion to a correct profile (left) and to a somewhat incorrect [100] profile (right) in terms of the following differences discussed in the text:

J(pz) - JCC* (pz) (solid line);

J* (pz) - JCC* (pz) (black dots);

J(pz) - JCC* (pz) (squares).

The experimental noise (Ö N) is marked by dashed lines where N denotes the total number of counts for each sampling point.

The inconsistent part of the data is defined as the difference J* (pz) - JCC* (pz), where JCC* (pz) denotes J* (pz) after applying the CC, i.e., after the first l/2 coefficients in (10) have been set equal to zero. Figure 2 shows first of all how an expression in the form of (10) describes two different experimental profiles, on the left one with a small inconsistent part, on the right a different one (measured with pz along [100]) with a significant inconsistent part. The difference between J(pz) and J* (pz) is given by the solid curve which in both cases fluctuates between the dashed curves which represent the noise, defined as the square root of the total number of counts (valence and core contributions) for each sampling point of J(pz). The inconsistent part of each of the curves, given by J* (pz) - JCC* (pz), is denoted by the black dots. For the profile on the left of Fig. 2 this part is comparable to the noise; for the profile on the right it is much larger than the noise (note the difference between the vertical scales). Finally, the difference J(pz) - JCC* (pz), represented by the squares, shows how the profiles change by the combined application of the CC and the expansion into Jacobi polynomials if the data are described by six gln(p) functions (3) obtained from 9 CPs. In the [100] profile on the right of Fig. 2 there is a sizable residual systematic error. Here we point out that among 9 measured CPs only one (J[100]) has its inconsistent part larger than the noise possibly due to an improper multiple scattering correction. The same procedure applied to the theoretical CPs does not introduce any changes. It shows that they were calculated with high accuracy.

In Fig. 3 it is shown how application of the CC changes the experimental CPs and improves the agreement between theory and experiment. For the theoretical CPs all calculations of the anisotropy D J º J[100] - J[110] [cases (a-c) in the left half of Fig. 3] give almost the same result, irrespective of whether one considers the calculated profiles J(pz), the approximate profiles, J* (pz), or the approximate profiles JCC* (pz) corrected by the consistency condition. This shows that the theoretical profiles were evaluated with high accuracy. In contrast, the experimental anisotropy shown on the right is substantially changed by the consistency condition [case (c)] and in this case begins to look similar to the corresponding theoretical results on the left hand side of Fig. 3.

Fig. 3 Anisotropy D J º J[100] - J[110] for theoretical (left) and experimental (right) CPs when three different choices of profiles discussed in the text are used: (a) D J(pz) (solid line); (b) D J* (pz) (filled circles); (c) D JCC* (pz) (squares). The experimental noise (Ö N) is marked by dashed lines as in Fig. 2.

 

 

 

3 Reconstruction procedure

Turning to the reconstruction, we emphasize that only for special sets of directions one can use as many terms as the number of profiles in the expansions (2) and (3); in general the number of terms should be smaller [21-25]. In the present case of 9 profiles these expansions had to be restricted to a smaller number of lattice harmonics by making a least-squares fit of the truncated series (3) [26]. Thus, before interpreting the experimental data one should consider how many density components can be used and how accurately one is able to reproduce the densities from the available projections for the directions pz displayed in Fig. 1.

For this purpose we checked how many gl(p) can be used for interpreting the experimental profiles, taking into account the magnitude of the statistical error in the experiment. To this end, in Fig. 4 both the theoretical and the experimental anisotropic components gl(p) are depicted together with the experimental error. In the figure the experimental gl(p) are presented after their inconsistent parts defined (see (8) as

(11)

are subtracted. It is clear that g10(p) and g12(p) are smaller than the statistical error; therefore these two components should not be taken into account in the reconstruction procedure.

 

Fig. 4 Functions gl(p) for l¹ 0 obtained from 9 theoretical (thick solid lines) and 9 experimental (open circles connected by thin solid lines) CPs, the latter after removal of their inconsistent parts. The experimental noise (Ö N) is marked by dashed lines as in Fig. 2.

We have examined how well we are able to reproduce the theoretical density in Cu0.9Al0.1, computed directly within the KKR-CPA scheme, by density components reconstructed from 9 theoretical profiles. The reconstructed theoretical r(p) consisting of only the first four radial components reproduces such details as the lack of a Fermi break along [111] due to the presence of the neck and the small anisotropy between the [100] and [110] directions. However, it does not reproduce the prominent Umklapp component along [100], suggesting that one would be unable to observe this Umklapp in the densities reconstructed from the experimental profiles as well (this will be illustrated in Fig. 7 below). In Fig. 5 we show that in order to reproduce this Umklapp component along [100], one should use six functions rn(p). However, even then there are discrepancies in reproducing some details of the Umklapp component and furthermore the break at the Fermi momentum along [100] is broadened.

Fig. 5. Theoretical KKR-CPA 3D momentum density r(p) in Cu0.9Al0.1 along three high symmetry directions is compared with corresponding r(p) reconstructed from 9 theoretical CPs described by 6 lattice harmonics (normalized to r(0) = 1).

 

 

 

 

4. Momentum densities

Our presentation of the reconstructed densities is divided into two subsections, Sects. 4.1 and 4.2, which discuss the analysis of the densities in extended p-space and reduced k-space, respectively.

4.1 p-space

As in many other CP studies, electron-electron (e-e) correlation effects produce differences between experimental and LDA-calculated Compton profiles. These differences are very similar for all profiles considered - we present two examples in Fig. 6. This type of discrepancy can be partly removed by means of the Lam-Platzman correction [1,27,28]. Since the e-e correlations manifest themselves mainly in the isotropic part of the momentum density, r0(p), we focus below on the anisotropic part of r(p), i.e. r a(p) = r(p) - r0(p).

Fig. 6 Differences between theoretical and experimental Compton profiles before (dots joined by thin solid lines) and after (solid line) applying the consistency condition. In comparison, the noise is shown by dashed lines.

In accord with the considerations outlined in Section III above, we restrict the analysis of the reconstructed densities to a description by three anisotropic radial density components r4, r6 and r8. The anisotropic parts, r a(p), along three high and three low symmetry directions, consisting of three anisotropic components, are displayed in Fig. 7 (results for two anisotropic components are very similar). For the high symmetry directions we present also the anisotropic parts of the theoretical densities computed directly, i.e. r(p) - r0(p), where r0(p) denotes the isotropic average reconstructed from the theoretical CPs. The anisotropic part r a(p) in the symmetry plane (110) is displayed in Fig. 8 showing that the general features of r a(p) are reproduced reasonably well.

 

 

 

Fig. 7 Anisotropic part, r a(p), of r(p), using r4, r6 and r8, along three high and three low symmetry directions in units of r(0)=1. Densities reconstructed from theoretical and experimental CPs are marked by circles and dots, respectively. Theoretically computed KKR-CPA densities r(p) - r0(p) are marked by solid lines for the three high symmetry directions.

 

Our routine studies show that when we use more than 40 expansion coefficients ak (see (5) and (8)), for describing the experimental rl(p) (with l¹ 0) we do not further obtain any essential detail but only undesired fluctuations coming from the experimental noise. This is not the case for theoretical densities, which neither contain experimental noise nor are smeared by the resolution function of the equipment. However, since for comparing theory and experiment it is reasonable to use the same number of ak, all results presented below are for 40 ak coefficients.

 

 

 

 

 

Fig. 8 Anisotropic part of r(p) consisting of three components rn in the (110) plane as discussed in the text, reconstructed from theoretical (a) and experimental (b) CPs. Each half of the figure extends over the range 1.83´ 1.83 (a.u.)2. The density ranges from -0.08 to 0.13 in units of r(0)=1. A step size between two of 12 contour lines equals 0.02. Lighter grey shades denote higher values.

 

In order to investigate the propagation of statistical noise into the final reconstructed results we have artificially superimposed noise on the theoretical CPs consistent with the statistics of the experiment [5´ 105 counts per sampling point at the CP peak (valence + core contribution)]. The noise is simulated using a gaussian random number generator with standard deviation s =Ö N, where N denotes the total number of counts per sampling point. Theoretical profiles are taken in the range 0 to 5 a.u. for the same sampling momenta pz as measured in the experiment.

Having performed simulations for M=20 sets of 9 profiles we were able to estimate the noise propagation in terms of standard deviations s [r(p)] using estimators

, (12)

where . The same analysis was also performed for r(k). Our results for s [ra(p)] are presented in Fig. 9 and show that the error is the highest along the high symmetry directions, particularly along [001] and [111].

The propagation of noise s [r(p)] in Compton scattering experiments was studied in [1,5,28] within the Fourier-Bessel reconstruction technique and in [27] for the direct Fourier transform. All authors

found that s [r(p)] has strong maxima along the high symmetry directions, in particular along [001] and [111]. Our understanding of this behavior of s [r(p)] is the following.

Each profile contains noise which, of course, has no symmetry. However, the measurements are performed only in the non-equivalent part of the Brillouin zone and the data (and therefore also the noise) are expanded into lattice harmonics. So, one will always find a higher error s [r(p)] along the high symmetry directions [001] and [111] because of the extreme values of the lattice harmonics in those directions.

 

 

Fig. 9 Error map, s [ra(p)], in the (110) plane based on the first three anisotropic components in units of r(0)=1. The figure covers the range 0.92´ 0.92$ (a.u.)2 and contains a total of 11 contours with a step size of D [ra]= 0.0018.

 

 

 

4.2 k-space

In order to obtain information about the FS, r(p) should be folded back into the first Brillouin zone via LCW-folding [7] to obtain r(k) as

, (13)

where k denotes a vector in the first Brillouin zone and the summation is performed over all reciprocal lattice vectors G. As a result of this procedure the lattice effects on r(p) are emphasized and r(k) shows the sum of the occupation numbers nl (unity for filled states and zero otherwise within the independent particle model) over the bands l. The results of the folding are shown in Fig. 10. In our case the summation in (13) has been carried out over a cube in momentum space of sides 5.0 a.u. and involves 65 vectors G. We divided these vectors, according to their length, into 7 groups: 0 refers only to the vector [000]; 1 refers to vector [000] and eight vectors of type <111>; 2 refers to the previous set plus six vectors of type <200>, etc. The summation in (13) was performed in steps using the above groups of reciprocal vectors G. In the case of the theoretical densities the LCW procedure converges with increasing number of vectors G, although most of the detail in r(k) is determined by the first few reciprocal vectors. In the case of experimental data (due to experimental errors and limited resolution) it does not make sense to use a higher number of vectors G than presented in Fig. 10.

Fig. 10 Folded momentum density r(k) in Cu0.9Al0.1 in the repeated zone scheme, obtained from theoretical (b) and experimental (c and d) CPs where r(p) contains four components (r0,r4, r6,r8). Panel (c) shows the experimental densities after subtraction of e-e correlation effects, i.e. by replacing the experimental r0 by the theoretical one as discussed in the text. The corresponding error, s [r(k)], is shown in panel (a). Different rows of figures show r(k) obtained by using successively larger groups of vectors G in the folding process as indicated by the corresponding numbers 0 to 2 (see text for details). In panels (b)-(d) the step sizes D r for a total of 31 contour lines for rows marked 0, 1 and 2 equal 0.027, 0.02, 0.017 in units of r(0)=1, respectively. In panel (a), there is a total of 6 contours and the step size D s [r(k)] is 0.005.

 

The folded r(k) derived from the theoretical CPs, shown in Fig. 10 (b), as well as that employing the experimental CPs [Fig. 10 (d)], displays clear signatures of the well-known FS of Cu. The most important discrepancy between the theoretical and the experimental r(k) is observed along the [111] direction around the neck. This discrepancy is much larger than the corresponding error s [r(k)] depicted in Fig. 10 (a) and is ascribed to e-e correlations. This is consistent with the results of Fig. 10 (c) where the experimental r0(p) (which contains correlation) is replaced by the theoretical r0(p) which does not contain Lam-Platzman correlation correction [14].

The error, s [r(k)], around the G point [Fig. 10(a)] is relatively small due to the application of an averaging procedure, also applied to the reconstructed densities. Since rl(p) (for small momenta) strongly depends on the number of coefficients alnk, one can eliminate some of the noise by considering rl(p) as a function of the number of alnk and taking its average.

5. Conclusions

For the first time our new reconstruction procedure has been applied to real experimental data. A comparison of the 3D momentum density r(p) reconstructed from theoretical profiles with r(p) computed directly (without reconstruction) shows that this procedure is successful.

In order to obtain a map of the FS we applied the LCW pocedure, i.e. folding reconstructed densities r(p) from the extended zone p into the first Brillouin zone (reduced zone k). These results indicate that the folded density r(k) yields a reasonable description of the FS obtained within the framework of the fully selfconsistent KKR-CPA scheme.

The principal discrepancy observed along the [111] direction around the neck is attributed to e-e correlation effects which are also responsible for much of the difference between theory and experiment in the extended p-space.

Our results with respect to the propagation of noise into the reconstructed densities show that since each density component, rl(p), contains noise and since the lattice harmonics have extremes along the high symmetry directions, there will always be an anisotropy in the reconstructed noise which has a tendency to accumulate along the high symmetry directions (in particular along [001] and [111]).

Acknowledgements

We are grateful to Peer Mijnarends and Bernardo Barbiellini for important discussions. This work was supported by the Polish State Committee for Scientific Research (Grant 2 P03B 083 16), the German Federal Ministry of Education and Research Contract 05 ST 8 HRA, the US Department of Energy contract W-31-109-ENG-38, and benefited from the allocation of supercomputer time at the NERSC and the Northeastern University Advanced Scientific Computation Center (NU-ASCC).

References

[1] See, e.g., G. Stutz, F. Wohlert, A. Kaprolat, W. Schülke, Y. Sakurai, Y. Tanaka, M. Ito, H. Kawata, N. Shiotani, S. Kaprzyk and A. Bansil, Phys. Rev. B 60, 7099 (1999), and references therein.

[2] A. Bansil, Z. Naturforsch. A 48a, 165 (1993)

[3] A. Bansil, S. Kaprzyk, P.E. Mijnarends and J. Tobo³a, Phys. Rev. B 60, 13396 (1999)

[4] G. Kontrym-Sznajd, M. Samsel-Czeka³a, Appl. Phys. A 70, 89 (2000); G. Kontrym-Sznajd, M. Samsel, R.N. West, Acta Phys. Polon. A 95, 591 (1999)

[5] N.K. Hansen, P. Pattison, J.R. Schneider, Z. Phys. B 66, 305 (1987)

[6] A.M. Cormack, Phys. Med. Biol. 18, 195 (1973)

[7] D.G. Lock, V.H.C. Crisp, R.N. West, J. Phys. F3, 561 (1973)

[8] P. Suortti, P. Pattison and W. Weyrich, J. Appl. Cryst 19, 343 (1986)

[9] R. Ribberfors, Phys. Rev B 12, 2067 (1975)

[10] R. Ribberfors, Phys. Rev B 12, 3136 (1975)

[11] F. Biggs, L.B. Mendelsohn and J.B. Mann, At. Data and Nucl. Data Tables 16, 201 (1795)

[12] J. Felsteiner, P. Pattison and M.J. Cooper, Phil. Mag. 30, 537 (1974)

[13] A. Bansil, B. Barbiellini, S. Kaprzyk and P.E. Mijnarends, J. Phys. Chem. Solids 62, 2191 (2001)

[14] L. Lam and P.M. Platzman, Phys. Rev. B 9, 5122 (1974)

[15] G. Kontrym-Sznajd, Appl. Phys. A 70, 97 (2000)

[16] A.M. Cormack, J. Appl. Phys. A 35, 2908 (1964)

[17] R.M. Lewitt, Proc. IEEE 71, 390 (1983)

[18] R.M. Lewitt and R.H.T. Bates, Optik 50, 189 (1978)

[19] H. Kudo and T. Saito, J. Opt. Soc. Am. A 8, 1148 (1991)

[20] R.M. Lewitt, private communication

[21] P.E. Mijnarends, Phys. Rev. 160, 512 (1967)

[22] F.M. Mueller and M.G. Priestley, Phys. Rev. 148, 638 (1966)

[23] A. Bansil, Solid State Commun. 16, 885 (1975)

[24] W.R. Fehlner, S.B. Nickerson, S.H. Vosko, Solid State Commun. 19, 83 (1976); W.R. Fehlner, S.H. Vosko, Can. J. Phys. 54, 2159 (1976).

[25] G. Kontrym-Sznajd, A. Jura and M. Samsel-Czeka³a, Appl. Phys. A 74, 605-612 (2002)

[26] C. Lanczos, Applied Analysis (Pitman and Sons, Ltd., London 1964), Chap. II

[27] Y. Tanaka, Y. Sakurai, A.T. Stewart, N. Shiotani, P.E. Mijnarends, S. Kaprzyk, A. Bansil, Phys. Rev. B 63, 045120 (2001)

[28] W. Schülke, G. Stutz, F. Wohlert, and A. Kaprolat, Phys. Rev. B 54, 14381 (1996)