Synthesis, crystal structures and chemical bonding of

0 downloads 0 Views 575KB Size Report
Note that RE3+ ions are replaced with ... the stabilization of the phase, and the Zintl–Klemm concept is ... Since Li+ is very similar to Mg2+ in terms of ionic size.
inorganic compounds Acta Crystallographica Section C

Crystal Structure Communications ISSN 0108-2701

Synthesis, crystal structures and chemical bonding of RE5xLixGe4 (RE = Nd, Sm and Gd; x ’ 1) with the orthorhombic Gd5Si4 type Nian-Tzu Suen, Tae-Soo You and Svilen Bobev* Department of Chemistry and Biochemistry, 304A Drake Hall, University of Delaware, Newark, DE 19716, USA Correspondence e-mail: [email protected] Received 18 October 2012 Accepted 6 December 2012 Online 13 December 2012

The syntheses and single-crystal and electronic structures of three new ternary lithium rare earth germanides, RE5xLixGe4 (RE = Nd, Sm and Gd; x ’ 1), namely tetrasamarium lithium tetragermanide (Sm3.97Li1.03Ge4), tetraneodymium lithium tetragermanide (Nd3.97Li1.03Ge4) and tetragadolinium lithium tetragermanide (Gd3.96Li1.03Ge4), are reported. All three compounds crystallize in the orthorhombic space group Pnma and adopt the Gd5Si4 structure type (Pearson code oP36). There are six atoms in the asymmetric unit: Li1 in Wyckoff site 4c, RE1 in 8d, RE2 in 8d, Ge1 in 8d, Ge2 in 4c and Ge3 in 4c. One of the RE sites, i.e. RE2, is statistically occupied by RE and Li atoms, accounting for the small deviation from ideal RE4LiGe4 stoichiometry.

Our previous work on Mg substitutions in RE5Ge4 has already demonstrated that the structure is amenable to variations. For brevity, RE will denote a rare earth metal throughout this article. Note that RE3+ ions are replaced with Mg2+, which allows for structural changes that are also coupled with the magnetic properties of the resultant RE5xMgxGe4 materials (Tobash et al., 2009). Importantly, our study has also shown that Mg first replaces the rare earth metal at the 4c site. Mg also replaces some of the RE2 at site 8d, leading to the final formula RE5xMgxGe4 (x ’ 2). The ‘phase width’ can be easily understood from the point of view of the Zintl–Klemm concept (Kauzlarich, 1996). In RE5Ge4 (Sm5Ge4 type), two of the Ge atoms are dimerized and two remain as isolated Ge atoms. The Ge dimer needs six electrons (three on each Ge) to satisfy the octet rule and the isolated Ge atom needs four electrons. Accordingly, the overall electron count for the RE5Ge4 structure will be (RE3+)5(Ge3)2(Ge4)2(e). For RE3Mg2Ge4, the total number of valence electrons contributed by the cations decreases, leading to a structural distortion that brings the isolated Ge atoms together to form an additional dimer. The electronic structure change here can be rationalized as (RE3+)3(Mg2+)2(Ge3)4(e). The ideal electron count will be realised for the formula RE2Mg3Ge4, but our studies have not shown evidence for extending the stoichiometry much beyond RE5xMgxGe4 (x ’ 2) (Tobash et al., 2009). This suggests that metal–metal bonding and efficient packing are also critical for the stabilization of the phase, and the Zintl–Klemm concept is just a good starting point for consideration. Nonetheless, the ideas above can help to explain the fewer substitutions of RE3+ by Li+ cations, as will now be discussed. Since Li+ is very similar to Mg2+ in terms of ionic size (Shannon, 1976), it is not surprising that the crystal chemistry observed when replacing the rare earth metal with Li mirrors

Comment Following the discovery of giant magnetocaloric effects in Gd5Si2Ge2 in 1997 (Pecharsky & Gschneidner, 1997a), great effort has been devoted to the synthesis and study of the structures and physical properties of rare earth silicides and germanides with the Gd5Si4 or Sm5Ge4 structure types, and also of their intermediates (Kyunghan et al., 1997; Misra & Miller, 2008). In this paper, we report the syntheses and singlecrystal structures of three compounds with the Sm5Ge4 structure type, namely Sm3.97Li1.03Ge4, Nd3.97Li1.03Ge4 and Gd3.96Li1.03Ge4. Many articles have already discussed the structural similarities and differences between the Gd5Si4 and Sm5Ge4 structures, which are isopointal but not isostructural (Pecharsky & Gschneidner, 1997b) (Fig. 1). Summarizing these findings briefly, we must point out that there are two pairs of Si—Si bonds in Gd5Si4 but only one type of Ge—Ge bond in Sm5Ge4 (Pecharsky & Gschneidner, 1997b). Furthermore, there are many factors, such as temperature, pressure and chemical make-up, which can cause one of the interslab bonds to form or cleave reversibly (Levin et al., 2002; Magen et al., 2003). Acta Cryst. (2013). C69, 1–4

Figure 1 Side-by-side comparison of the Sm5Ge4 and Gd5Si4 structures (Pecharsky & Gschneidner, 1997b). The rare earth metals are shown as large spheres (green in the electronic version of the paper), Ge/Si atoms as light spheres (orange in the electronic version) and Li atoms as dark spheres (blue). The ‘interslab’ dimers and broken dimers are outlined.

doi:10.1107/S0108270112050032

# 2013 International Union of Crystallography

1

inorganic compounds

Figure 2 (a) A representation of the Sm5xLixGe4 structure, viewed approximately down the [100] direction. The rare earth metals are shown as large spheres (green in the electronic version of the paper), Ge atoms as medium spheres (orange) and Li atoms as small spheres (blue). (b) A representation of the first building unit, comprising a trigonal prism (shaded; orange in the electronic version) and a parallelepiped (hollow framework of rare earth metals). (c) A view of the chain of vertex/edgeshared pentagonal rings built by Li atoms and Ge—Ge dimers.

the results obtained with Mg. The RE5xLixGe4 structure (Gd5Si4 type) is the same as that of RE5xMgxGe4 and has been considered in detail elsewhere (Pavlyuk et al., 1990; Xie et al., 2008; Peter et al., 2012; Fornasini et al., 2012). It can be broken down into two simpler building units (Fig. 2). One is the framework made by the rare earth metals, which is a combination of trigonal prisms and parallelepipeds, with Ge atoms located at the centres of the trigonal prisms. The second building unit is the pentagonal ring (Fig. 2c) made by Li and Ge atoms (paired in the ac plane). The interatomic distances within the Ge1—Ge1(x, y, z + 1) and Ge2—Ge3(x, y, ˚ . There z  1) dumbbells fall in the range 2.633 (2)–2.528 (3) A does not appear to be any correlation between the refined Ge—Ge contacts and the decreasing unit-cell volumes due to the lanthanide contraction. We also note that the Ge—Ge distances are slightly longer than the sum of the corresponding Pauling radii (Pauling, 1960), which suggests the Ge—Ge interactions are best regarded as covalent two-centre–twoelectron bonds. Therefore, the electron count for the RE4LiGe4 structure can be represented as (RE3+)4(Li+)(Ge3)4(e). Such an electronic structure also accounts for the different homogeneity range of RE5xLixGe4 compared with RE5xMgxGe4. For instance, if we apply the same electron count for RE3Li2Ge4, we now have only 11 electrons from the cations, which are not sufficient to supply the electrons needed to form two pairs of Ge—Ge dimers (12 negative charges). Li substitutions in RE5Ge4 were known before our work commenced, the first report being the example of Tm4LiGe4, which was characterized by Pavlyuk et al. (1990) on the basis of single-crystal X-ray diffraction data. The same paper also identifies the isotypic RE4LiGe4 (RE = Y, Gd, Er Lu) from powder X-ray diffraction data without discussing the possible

2

Suen et al.



Sm3.97Li1.03Ge4, Nd3.97Li1.03Ge4 and Gd3.96Li1.03Ge4

homogeneity range. Very recently, Fornasini et al. (2012) reported the Nd4LiGe4 structure, also refined from singlecrystal X-ray diffraction data, but they also did not comment on the small RE/Li disorder we have identified in our sample. The compound Yb2Li0.5Ge2 (Yb4LiGe4) has been studied twice in the last four years, first by Xie et al. (2008) and then by Peter et al. (2012). Yb4LiGe4 was suggested to be a mixedvalent Yb2+/Yb3+ system, similar to Yb5xMgxGe4 (Tobash & Bobev, 2006), but again with no mention of the homogeneity range. It should also be noted that we have chosen to discuss the present structures not as stoichiometric compounds RE4LiGe4, as done previously, but rather as RE5xLixGe4. The deviation from the idealized formula RE4LiGe4 is very small indeed, making it not very conspicuous, but the following two signs of the narrow phase width cannot be ignored. Firstly, powder patterns of the products of reactions with different amounts of lithium metal as a starting material reveal peak shifts, which indicate that the unit-cell volumes change as a function of the nominal composition. Unfortunately, these were not phase-pure materials and further structural analysis was severely hindered. Fornasini et al. (2012) also commented on the heterogeneity of the samples. Some of the common RE–Li–Ge phases identified in this process include RELiGe2 (RE = La–Nd, Sm and Eu; Pavlyuk et al., 1986; Bobev et al., 2012), RE2Li2Ge3 and RE3Li4Ge4 (Pavlyuk et al., 1989; Guo, You & Bobev, 2012), and RE7Li8Ge10 and RE11Li12Ge16 (RE = La–Nd, Sm; Guo, You, Jung & Bobev, 2012). The fact that Pavlyuk et al. (1990) reported a smaller unit cell for their Gd4LiGe4 sample than ours is yet another indicator that the amount of Li in RE5xLixGe4 can be varied within a certain range. The major difficulty in establishing the solubility limits arises from the fact that Li and the rare earth metals have very different melting points, so obtaining equilibrium conditions is nearly impossible. Since the studied crystals were chosen from reactions with an excess of Li, we may speculate that RE5xLixGe4 (x ’ 1) most likely represents the most Li-rich compositions that can be synthesized under the given conditions. Secondly, when the occupancy factors of the Wyckoff sites 8d were freed to vary during trial refinements, for RE1 in all three cases the freed variable was always close to 100%, while the RE2 site consistently refined with 97% occupancy. A very similar observation was made with regard to the RE5xMgxGe4 (x ’ 2) structure (Tobash et al., 2009), so we have assumed that Li can also partially substitute the RE2 atoms just like Mg does. The single-crystal X-ray data for the latter metal showed that the substitution effect was well pronounced and easy to identify, while the very light Li makes it hard to detect or characterize. However, the refinement of a mixed occupied RE2/Li2 site was found to be statistically significant according to the Hamilton test (Hamilton, 1965). Specifically, at a significance level of 0.005, for models A (RE5xLixGe4) and B (RE4LiGe4), the hypothesis that the former is better than the latter can be verified based on R = 1.004 < RB /RA = 1.013, where R is the interpolated value at the given significance level. Acta Cryst. (2013). C69, 1–4

inorganic compounds All of the above confirms that the title compounds are better described as solid solutions RE5xLixGe4 (x ’ 1) with narrow homogeneity ranges, rather than stoichiometric compounds RE4LiGe4.

Data collection Bruker SMART CCD area-detector diffractometer Absorption correction: multi-scan (SADABS; Sheldrick, 2008a) Tmin = 0.410, Tmax = 0.531

10856 measured reflections 1094 independent reflections 823 reflections with I > 2(I) Rint = 0.124

Refinement

Experimental All starting materials were purchased from common chemical vendors [pure elements from Alfa or Aldrich (>99.9 wt%)] and stored and handled in an argon-filled glove-box to prevent their deterioration due to moisture and oxygen. All reactions were carried out by loading stoichiometric amounts of the respective elements into Nb containers, which were then sealed with an arc welder. Subsequently, the Nb containers were flame sealed in evacuated (ca 105 Torr; 1 Torr = 133.322 Pa) fused silica jackets. The synthesis followed the conventional solid-state route through direct fusion, but the process was somewhat complicated due to the very high melting points of the rare earth metals. The general synthetic route was that the reactions were heated in a tube furnace to 1358 K at a rate of 200 K h1, kept there for 5 h and then cooled slowly to 573 K at a rate of 10 K h1. The title compounds were first identified as side products of reactions aimed at synthesizing NdLiGe2 and SmLiGe2. As mentioned in our previous work (Bobev et al., 2012), due to the high melting points of Nd (1300 K) and Sm (1315 K) it is necessary to increase the synthetic temperature. However, this also increases the possibility of Li leaking out of the Nb container, which would destroy the stoichiometric ratio and could lead to other phases. If Li were to leak outside the Nb tube and condense in the silica tube, it could be dangerous and the furnace should be stopped immediately.

R[F 2 > 2(F 2)] = 0.036 wR(F 2) = 0.060 S = 1.05 1094 reflections

45 parameters ˚ 3 max = 2.69 e A ˚ 3 min = 2.16 e A

Compound (III) Crystal data Gd3.96Li1.03Ge4 Mr = 920.66 Orthorhombic, Pnma ˚ a = 7.188 (2) A ˚ b = 14.816 (5) A ˚ c = 7.767 (3) A

˚3 V = 827.2 (5) A Z=4 Mo K radiation  = 45.51 mm1 T = 200 K 0.06  0.05  0.04 mm

Data collection Bruker SMART CCD area-detector diffractometer Absorption correction: multi-scan (SADABS; Sheldrick, 2008a) Tmin = 0.181, Tmax = 0.277

10177 measured reflections 1044 independent reflections 800 reflections with I > 2(I) Rint = 0.141

Refinement R[F 2 > 2(F 2)] = 0.040 wR(F 2) = 0.087 S = 1.01 1044 reflections

45 parameters ˚ 3 max = 2.35 e A ˚ 3 min = 2.66 e A

Compound (I) Crystal data Sm3.97Li1.03Ge4 Mr = 893.68 Orthorhombic, Pnma ˚ a = 7.2846 (16) A ˚ b = 14.938 (3) A ˚ c = 7.8594 (17) A

˚3 V = 855.3 (3) A Z=4 Mo K radiation  = 40.51 mm1 T = 200 K 0.04  0.03  0.03 mm

Data collection Bruker SMART CCD area-detector diffractometer Absorption correction: multi-scan (SADABS; Sheldrick, 2008a) Tmin = 0.316, Tmax = 0.408

11551 measured reflections 1188 independent reflections 968 reflections with I > 2(I) Rint = 0.085

Refinement R[F 2 > 2(F 2)] = 0.028 wR(F 2) = 0.051 S = 1.02 1188 reflections

48 parameters ˚ 3 max = 1.93 e A ˚ 3 min = 2.07 e A

Compound (II) Crystal data Nd3.97Li1.03Ge4 Mr = 870.14 Orthorhombic, Pnma ˚ a = 7.3649 (12) A ˚ b = 15.092 (2) A ˚ c = 7.9406 (13) A Acta Cryst. (2013). C69, 1–4

˚3 V = 882.6 (2) A Z=4 Mo K radiation  = 36.23 mm1 T = 200 K 0.03  0.03  0.02 mm

Crystals were selected under an optical microscope and cut under oil to the desired small dimensions. They were then mounted on glass fibres and quickly placed under a cold nitrogen stream (ca 200 K) on the diffractometer. The crystals diffracted strongly and an exposure time of 8–12 s per frame was sufficient. Refined parameters included the scale factor, extinction coefficients and atomic positions, with the corresponding anisotropic displacement parameters. Li had to be refined with an isotropic displacement parameter in the case of the Nd and Gd compounds, likely because of the inferior crystal quality compared with the Sm analogue. This is not without precedent for elements with such a low Z number in structures dominated by very heavy elements. The atomic positions were standardized using STRUCTURE TIDY (Gelato & Parthe´, 1987) prior to the final refinement step. Structure refinements using the coordinates from the parent Sm5Ge4 type did not converge and it was obvious that the RE3 site (site 4c) was occupied by a much lighter atom. Therefore, in the next least-squares refinement cycles, RE3 was assigned as Li1 and its occupancy factor was allowed to vary. This time the refinements converged, and the site-occupancy factor for Li1 was very close to the full value. In the next round of least-squares refinement, all atoms (except Li in the case of the Nd and Gd compounds) were refined with anisotropic displacement parameters (ADPs) and fixed occupancies. Refinements converged at much lower R values. Subsequent refinement with free RE and Ge site-occupancy factors (done for an individual site, while the remaining site-occupancy factors were kept fixed) proved that the RE2 sites in all three structures are slightly underoccupied (within 6–8). All remaining crystallographic sites Suen et al.



Sm3.97Li1.03Ge4, Nd3.97Li1.03Ge4 and Gd3.96Li1.03Ge4

3

inorganic compounds proved to be fully occupied, with the corresponding deviations from full occupancy less than 2. Since Li is too light when refined together with RE atoms at the RE2 sites, its occupancy factor was very small. The actual solubility limits were not investigated, as obtaining phase-pure samples was hindered by the very different melting points of Li and the rare earth metals. However, we note that such a stoichiometry breadth (and substitution pattern) is not unusual for germanide compounds with this structure type, as evidenced by studies of RE5xMgxGe4 (Tobash et al., 2009). The highest residual density for the Sm compound is located ˚ from Ge1 and the deepest hole is 1.84 A ˚ from Ge2. For the Nd 1.86 A ˚ ) and Gd1 and Gd analogues, the peaks are close to Nd2 (1.9 A ˚ ), while the holes are located near Nd1 (0.84 A ˚ ) and Ge2 (0.85 A ˚ ), respectively. (0.72 A For all compounds, data collection: SMART (Bruker, 2002); cell refinement: SMART; data reduction: SAINT (Bruker, 2002); program(s) used to solve structure: SHELXS97 (Sheldrick, 2008b); program(s) used to refine structure: SHELXL97 (Sheldrick, 2008b); molecular graphics: SHELXTL (Sheldrick, 2008b); software used to prepare material for publication: SHELXTL.

SB acknowledges financial support from the National Science Foundation (NSF) (CAREER grant No. DMR0743916). Supplementary data for this paper are available from the IUCr electronic archives (Reference: FN3120). Services for accessing these data are described at the back of the journal.

References Bobev, S., You, T.-S., Suen, N.-T., Saha, S., Greene, R. & Paglione, J. (2012). Inorg. Chem. 51, 620–628.

4

Suen et al.



Sm3.97Li1.03Ge4, Nd3.97Li1.03Ge4 and Gd3.96Li1.03Ge4

Bruker (2002). SMART and SAINT. Bruker AXS Inc., Madison, Wisconsin, USA. Fornasini, M. L., Palezona, A. & Pani, M. (2012). Intermetallics, 31, 114– 119. Gelato, L. M. & Parthe´, E. (1987). J. Appl. Cryst. 20, 139–143. Guo, S.-P., You, T.-S. & Bobev, S. (2012). Inorg. Chem. 51, 3119–3129. Guo, S.-P., You, T.-S., Jung, Y.-H. & Bobev, S. (2012). Inorg. Chem. 51, 6821– 6829. Hamilton, W. C. (1965). Acta Cryst. 18, 502–510. Kauzlarich, S. M. (1996). Editor. Chemistry, Structure and Bonding of Zintl Phases and Ions. Weinheim: VCH. Kyunghan, A., Pecharsky, V. K. & Gschneidner, K. A. Jr (1997). Phys. Rev. B, 76, 0144151–01441513. Levin, E. M., Gschneidner, K. A. Jr & Pecharsky, V. K. (2002). Phys. Rev. B, 65, 214427–214431. Magen, C., Arnold, Z., Morellon, L., Skorokhod, Y., Algarabel, P. A., Ibarra, M. R. & Kamarad, J. (2003). Phys. Rev. Lett. 91, 207202–207205. Misra, S. & Miller, G. J. (2008). J. Am. Chem. Soc. 130, 13900–13911. Pauling, L. (1960). In The Nature of the Chemical Bond. Ithaca: Cornell University Press. Pavlyuk, V., Bodak, O. & Zavodnik, V. Y. (1990). Dopov. Akad. Nauk Ukr. RSR Ser. B, 12, 29–31. Pavlyuk, V. V., Pecharskii, V. K. & Bodak, O. I. (1986). Dopov. Akad. Nauk Ukr. RSR Ser. A, 7, 78–80. Pavlyuk, V. V., Pecharskii, V. K. & Bodak, O. I. (1989). Dopov. Akad. Nauk Ukr. RSR Ser. B, 2, 51–54. Pecharsky, V. K. & Gschneidner, K. A. Jr (1997a). Phys. Rev. Lett. 78, 4494– 4497. Pecharsky, V. K. & Gschneidner, K. A. Jr (1997b). J. Alloys Compd, 260, 98– 106. Peter, S. C., Disseler, S. M., Svensson, J. N., Carretta, P. & Graf, M. J. (2012). J. Alloys Compd, 516, 126–131. Shannon, R. D. (1976). Acta Cryst. A32, 751–767. Sheldrick, G. M. (2008a). SADABS. University of Go¨ttingen, Germany Sheldrick, G. M. (2008b). Acta Cryst. A64, 112–122. Tobash, P. H. & Bobev, S. (2006). J. Am. Chem. Soc. 128, 3252–3254. Tobash, P. H., Bobev, S., Thompson, J. D. & Sarrao, J. L. (2009). Inorg. Chem. 48, 6641–6651. Xie, Q. X., Kubata, C., Woerle, M. & Nesper, R. (2008). Z. Anorg. Allg. Chem. 634, 2469–2476.

Acta Cryst. (2013). C69, 1–4

supplementary materials

supplementary materials Acta Cryst. (2013). C69, 1-4

[doi:10.1107/S0108270112050032]

Synthesis, crystal structures and chemical bonding of RE5−xLixGe4 (RE = Nd, Sm and Gd; x ≈ 1) with the orthorhombic Gd5Si4 type Nian-Tzu Suen, Tae-Soo You and Svilen Bobev (I) Tetrasamarium lithium tetragermanide Crystal data Sm3.97Li1.03Ge4 Mr = 893.68 Orthorhombic, Pnma Hall symbol: -p 2ac 2n a = 7.2846 (16) Å b = 14.938 (3) Å c = 7.8594 (17) Å V = 855.3 (3) Å3 Z=4

F(000) = 1508 Dx = 6.940 Mg m−3 Mo Kα radiation, λ = 0.71073 Å Cell parameters from 958 reflections θ = 4.1–23.5° µ = 40.51 mm−1 T = 200 K Irregular, grey 0.04 × 0.03 × 0.03 mm

Data collection Bruker SMART CCD area-detector diffractometer Radiation source: fine-focus sealed tube Graphite monochromator φ and ω scans Absorption correction: multi-scan (SADABS; Sheldrick, 2008a) Tmin = 0.316, Tmax = 0.408

11551 measured reflections 1188 independent reflections 968 reflections with I > 2σ(I) Rint = 0.085 θmax = 29.1°, θmin = 2.7° h = −9→9 k = −20→20 l = −10→10

Refinement Refinement on F2 Least-squares matrix: full R[F2 > 2σ(F2)] = 0.028 wR(F2) = 0.051 S = 1.02 1188 reflections 48 parameters 0 restraints Primary atom site location: structure-invariant direct methods

Secondary atom site location: difference Fourier map w = 1/[σ2(Fo2) + (0.015P)2] where P = (Fo2 + 2Fc2)/3 (Δ/σ)max = 0.001 Δρmax = 1.93 e Å−3 Δρmin = −2.07 e Å−3 Extinction correction: SHELXL97 (Sheldrick, 2008b), Fc*=kFc[1+0.001xFc2λ3/sin(2θ)]-1/4 Extinction coefficient: 0.00082 (5)

Special details Experimental. Data collection is performed with two batch runs at φ = 0.00 ° (456 frames), and at φ = 90.00 ° (456 frames). Frame width = 0.40 ° in ω. Data is merged, corrected for decay, and treated with multi-scan absorption corrections.

Acta Cryst. (2013). C69, 1-4

sup-1

supplementary materials Geometry. All esds (except the esd in the dihedral angle between two l.s. planes) are estimated using the full covariance matrix. The cell esds are taken into account individually in the estimation of esds in distances, angles and torsion angles; correlations between esds in cell parameters are only used when they are defined by crystal symmetry. An approximate (isotropic) treatment of cell esds is used for estimating esds involving l.s. planes. Refinement. Refinement of F2 against ALL reflections. The weighted R-factor wR and goodness of fit S are based on F2, conventional R-factors R are based on F, with F set to zero for negative F2. The threshold expression of F2 > 2sigma(F2) is used only for calculating R-factors(gt) etc. and is not relevant to the choice of reflections for refinement. R-factors based on F2 are statistically about twice as large as those based on F, and R-factors based on ALL data will be even larger. Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2)

Sm1 Sm2 Li2 Ge1 Ge2 Ge3 Li1

x

y

z

Uiso*/Ueq

0.01938 (6) 0.32672 (6) 0.32672 (6) 0.16093 (12) 0.01671 (16) 0.27599 (16) 0.158 (3)

0.59890 (3) 0.12815 (3) 0.12815 (3) 0.03620 (6) 0.2500 0.2500 0.2500

0.18707 (5) 0.17864 (5) 0.17864 (5) 0.46727 (11) 0.08118 (16) 0.86645 (15) 0.528 (3)

0.00832 (12) 0.00764 (14) 0.00764 (14) 0.0091 (2) 0.0091 (3) 0.0083 (3) 0.015 (4)

Occ. ( 2σ(I) Rint = 0.124 θmax = 27.9°, θmin = 2.7° h = −9→9 k = −19→19 l = −10→10

Refinement Refinement on F2 Least-squares matrix: full R[F2 > 2σ(F2)] = 0.036 wR(F2) = 0.060 S = 1.05 1094 reflections 45 parameters 0 restraints Primary atom site location: structure-invariant direct methods

Secondary atom site location: difference Fourier map w = 1/[σ2(Fo2) + (0.0075P)2] where P = (Fo2 + 2Fc2)/3 (Δ/σ)max < 0.001 Δρmax = 2.69 e Å−3 Δρmin = −2.16 e Å−3 Extinction correction: SHELXL97 (Sheldrick, 2008b), Fc*=kFc[1+0.001xFc2λ3/sin(2θ)]-1/4 Extinction coefficient: 0.00024 (4)

Special details Experimental. Data collection is performed with two batch runs at φ = 0.00 ° (456 frames), and at φ = 90.00 ° (456 frames). Frame width = 0.40 ° in ω. Data is merged, corrected for decay, and treated with multi-scan absorption corrections. Geometry. All esds (except the esd in the dihedral angle between two l.s. planes) are estimated using the full covariance matrix. The cell esds are taken into account individually in the estimation of esds in distances, angles and torsion angles; correlations between esds in cell parameters are only used when they are defined by crystal symmetry. An approximate (isotropic) treatment of cell esds is used for estimating esds involving l.s. planes.

Acta Cryst. (2013). C69, 1-4

sup-7

supplementary materials Refinement. Refinement of F2 against ALL reflections. The weighted R-factor wR and goodness of fit S are based on F2, conventional R-factors R are based on F, with F set to zero for negative F2. The threshold expression of F2 > 2sigma(F2) is used only for calculating R-factors(gt) etc. and is not relevant to the choice of reflections for refinement. R-factors based on F2 are statistically about twice as large as those based on F, and R-factors based on ALL data will be even larger. Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2)

Nd1 Nd2 Li2 Ge1 Ge2 Ge3 Li1

x

y

z

Uiso*/Ueq

0.02129 (8) 0.32566 (9) 0.32566 (9) 0.15859 (17) 0.0155 (2) 0.2731 (2) 0.163 (4)

0.59950 (4) 0.12811 (4) 0.12811 (4) 0.03601 (8) 0.2500 0.2500 0.2500

0.18745 (8) 0.17943 (8) 0.17943 (8) 0.46828 (15) 0.0794 (2) 0.8672 (2) 0.528 (3)

0.01011 (19) 0.0092 (2) 0.0092 (2) 0.0106 (3) 0.0109 (4) 0.0104 (4) 0.008 (6)*

Occ. ( 2σ(I) Rint = 0.141 θmax = 28.1°, θmin = 2.8° h = −9→9 k = −19→19 l = −10→10

Refinement Refinement on F2 Least-squares matrix: full R[F2 > 2σ(F2)] = 0.040 wR(F2) = 0.087 S = 1.01 1044 reflections 45 parameters 0 restraints Primary atom site location: structure-invariant direct methods

Secondary atom site location: difference Fourier map w = 1/[σ2(Fo2) + (0.030P)2] where P = (Fo2 + 2Fc2)/3 (Δ/σ)max < 0.001 Δρmax = 2.35 e Å−3 Δρmin = −2.66 e Å−3 Extinction correction: SHELXL97 (Sheldrick, 2008b), Fc*=kFc[1+0.001xFc2λ3/sin(2θ)]-1/4 Extinction coefficient: 0.00110 (10)

Special details Experimental. Data collection is performed with two batch runs at φ = 0.00 ° (456 frames), and at φ = 90.00 ° (456 frames). Frame width = 0.40 ° in ω. Data is merged, corrected for decay, and treated with multi-scan absorption corrections. Geometry. All esds (except the esd in the dihedral angle between two l.s. planes) are estimated using the full covariance matrix. The cell esds are taken into account individually in the estimation of esds in distances, angles and torsion angles; correlations between esds in cell parameters are only used when they are defined by crystal symmetry. An approximate (isotropic) treatment of cell esds is used for estimating esds involving l.s. planes. Refinement. Refinement of F2 against ALL reflections. The weighted R-factor wR and goodness of fit S are based on F2, conventional R-factors R are based on F, with F set to zero for negative F2. The threshold expression of F2 > 2sigma(F2) is used only for calculating R-factors(gt) etc. and is not relevant to the choice of reflections for refinement. R-factors based on F2 are statistically about twice as large as those based on F, and R-factors based on ALL data will be even larger. Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2)

Gd1 Gd2 Li2 Ge1 Ge2 Ge3 Li1

x

y

z

Uiso*/Ueq

0.01762 (9) 0.32797 (10) 0.32797 (10) 0.1626 (2) 0.0170 (3) 0.2792 (3) 0.154 (6)

0.59847 (5) 0.12807 (5) 0.12807 (5) 0.03613 (10) 0.2500 0.2500 0.2500

0.18663 (9) 0.17807 (9) 0.17807 (9) 0.46676 (19) 0.0827 (3) 0.8658 (3) 0.531 (6)

0.0099 (2) 0.0091 (3) 0.0091 (3) 0.0108 (4) 0.0105 (5) 0.0103 (5) 0.035 (11)*

Occ. (