Home Synthesis, crystal and electronic structure of the new sodium chain sulfido cobaltates(II), Na3CoS3 and Na5[CoS2]2(Br)
Article Publicly Available

Synthesis, crystal and electronic structure of the new sodium chain sulfido cobaltates(II), Na3CoS3 and Na5[CoS2]2(Br)

  • Pirmin Stüble , Jan P. Kägi and Caroline Röhr EMAIL logo
Published/Copyright: November 24, 2016
Become an author with De Gruyter Brill

Abstract

The sulfido cobaltate(II) Na3CoS3 was synthesized from stoichiometric quantities of Na2S, elemental cobalt and sulfur at a maximum temperature of 1100°C. According to Na3CoS3=Na12[Co2S5(S2)][Co2S3(S2)] the orthorhombic structure of a new type (space group Cmc21, a=884.24(2), b=2177.38(5), c=1193.20(3) pm, Z=12, R1=0.0205) contains two different anions: i. dimers [Co2S5(S2)]8− of two edge-sharing [CoS4] tetrahedra with five sulfido and one η1-disulfido ligand; ii. chains 1[Co2S3(S2)]4 of [CoS4] tetrahedra connected via μ-sulfido (3×) besides μ-1,2-disulfido (1×) ligands. The second title compound, Na5[CoS2]2(Br), which has been likewise synthesized as a pure phase from stoichiometric quantities of Na2S, Co, S and NaBr, is isotypic to Na5[CoS2]2(S) (Na6PbO5 type; tetragonal, space group I4mm, a=914.58(7), c=625.59(5) pm; R1=0.0412). The structure contains linear chains 1[CoS4/2]2 running along the tetragonal c axis. In between, Br ions are interspersed, which are coordinated by square pyramids of Na+ ions. The isotypic hydrogensulfide Na5[CoS2]2(SH) was obtained via the addition of NaSH. Several synthetic and structural arguments suggest that Na5[CoS2]2(SH) and the previously described pure sulfide are the same compound. The results of DFT band structure calculations (GGA+U, AFM spin ordering) are used to discuss and compare the chemical bonding in the new sulfido cobaltates(II) with that of the reference compound Na2[CoS2]. They also allow for obtaining insight into the superexchange path, which is responsible for the strong antiferromagnetic Co spin ordering along the chains.

1 Introduction

In contrast to the rich crystal chemistry of alkali sulfido ferrates, where Fe(II), Fe(III) and mixed valence metalates with a very diverse crystal chemistry occur, the chemistry of the respective cobaltates is dominated by the presence of Co(II) and is limited to the ortho-cobaltates A6[CoS4] (known for A=Na and K exclusively [1], [2], [3], [4]), the compounds A2[CoS2] with linear [CoS4/2]2− chains observed for all alkali cations [5] and the defect layers of the A2Co3S4 structure family, which are known for the heavy alkali cations K+ to Cs+ [6], [7], [8]. With these cations, ThCr2Si2-like layer structures have been described in addition [6], [9]. The only sulfido cobaltates containing Co(III) are the mixed-valent cubic phases A9Co2S7 (A=K, Rb, Cs; [10]), which exhibit distinct trigonal planar anions [CoIIS3]4− and [CoIIIS3]3− besides isolated sulfide ions, and the likewise mixed-valent chain compound Na5[CoS2]2(S) [11]. Most of these sulfido cobaltates have been synthesized in the 1980s by the Bronger group, using an H2S gas stream as sulfur source. The crystal chemistry of the corresponding alkali sulfido ferrates could be further extended in recent years [6], [12], [13], [14], [15], [16], [17] using alternative direct synthetic routes, i.e. starting from elemental sodium, metal M and sulfur (or in the case of the heavier alkali elements from the previously prepared alkali sulfide A2Sx, M and S). The ternary systems A-Co-S have now been similarly reexamined.

Herein we report on the new sodium sulfido cobaltate Na3CoS3, which – in contrast to its simple formula – is neither of a common structure nor is it a Co(III) compound. In addition, the unsuccessful attempts to reproduce the mixed-valent chain compound Na5[CoS2]2(S) [11] via direct syntheses suggested to prepare the actual isotypic pure Co(II) compound Na5[Co2]2(Br), in which bromide imitates the SH ions suspected in ‘Na5[CoS2]2(S)’.

2 Experimental

2.1 Preparation and phase analysis

The title compounds were synthesized in corundum crucibles sealed in steel autoclaves under an argon atmosphere starting from cobalt powder (99%, Schuchardt München) and elemental sulfur (99%, Merck) with anhydrous Na2S (95%, ABCR GmbH Karlsruhe ) and (for the initial sample) pure sodium (Merck) as Na sources. For the synthesis of the ‘double salts’, NaSH (ABCR) and NaBr (Merck) were used as additional starting materials. The sealed steel autoclaves were heated up to maximum temperatures (Tmax) between 400 and 1100°C, held there for 4 h and subsequently cooled to r. t. with rates of 20 to 30 Kh−1. After the preparations, representative parts of the reguli were ground and sealed in capillaries with a diameter of 0.3 mm. X-ray powder diagrams were collected on a transmission powder diffraction system (STADI-P, Dectris Mythen 1 K detector, Stoe & Cie, Darmstadt, Mo Kα radiation, graphite monochromator). For the phase analysis, the measured powder diagrams were compared to the calculated (program Lazy-Pulverix [18]) reflections of the title compounds and other known phases in the systems Na–Co–S–(Br/H). Rietveld refinements were performed using the programs Gsas [19] and Expgui [20].

Black-metallic shiny crystals of the new phase Na3CoS3 were first obtained from a sulfur-rich sample of overall composition Na2.4CoS6.6, which consisted of 86.8 mg (2.78 mmol) of elemental sodium, 92.6 mg (1.57 mmol) of cobalt and 332.9 (10.38 mmol) of sulfur. Due to the exothermic reaction, the sample container was slowly heated up to 200°C followed by rapid heating to 760°C. It was held at this temperature for 4 h and cooled to r. t. with a rate of 10 Kh−1. Most reflections of the complex powder diagram could be indexed with the calculated data of Na3CoS3 (Tables 1 and 2). In addition, reflections of CoS2 (which could be optically identified as octahedral crystals with a golden metallic lustre) and Na6[CoS4] (light-orange transparent crystals) are also found in the diagram. A further minor phase, which was identified using single crystal methods, are the metallic-greenish needles of the known compound Na2[CoS2]. Attempts to synthesize Na3CoS3 from samples of stoichiometric composition of the elements failed due to breaking of the corundum sample tubes, even with smaller amounts of the starting materials. The choice of Na2S as an alternative sodium source, i.e. starting from Na2S, Co and S in a 3:2:3 ratio, resulted only in inhomogeneous products when applying the previously used maximum temperature of 760°C. At an elevated Tmax of 1100°C this protocol turned out to be very successful for the synthesis of Na3CoS3. Samples with an overall weight of approx. 1 g (e.g. 522.4 mg/6.69 mmol Na2S, 263.0 mg/4.46 mmol Co and 214.6 mg/6.69 mmol S) yielded homogeneous reguli of the target compound, consisting of xenomorphic aggregates with dark-gray matt metallic gloss. Very small traces of Na6[CoS4] can be identified in the powder pattern as well as from an optical inspection of the products.

Table 1:

Crystallographic data and details of the data collection and structure determination for Na3CoS3 and Na5 [CoS2]2 (Br).

CompoundNa3CoS3Na5[CoS2]2(Br)
Crystal systemorthorhombictetragonal
Space groupCmc21, no. 36I4mm, no. 107
Pearson symboloC84tI24
Structure typenewNa6PbO5
Temperature, °C20
Lattice parameters, pm
a884.24(2)914.58(7)
b2177.38(5)
c1193.20(3)625.59(5)
Volume of the unit cell, 106 pm32297.30(9)523.28(9)
Z122
Density (X-ray), g cm−32.592.80
DiffractometerBruker AXS CCDStoe IPDS-2
Mo–Kα radiation
Absorption coefficient μMo−, mm−14.27.9
θ range, deg1.9–32.53.2–29.1
No. of reflections collected141962726
No. of independent reflections3223410
Rint0.02940.0755
CorrectionsLorentz, polarization, absorption
(Multi-Scan [21])(X shape [22])
Structure solutionShelxs-2013 [23]
Structure refinementShelxl-2013 [23]
No. of free parameters14924
Goodness-of-fit on F21.0211.141
Flack- x parameter/twin ratio0.092(10)53(5):47(5)a
R Values [for refl. with I≥2σ (I)]
R10.02050.0412
wR20.03890.1150
R Values (all data)
R10.02480.0437
wR20.03970.1168
Residual elect. density, e×10−6pm−3+0.6/−0.7+0.9/−0.8

aRefined as an inversion twin.

Table 2:

Atomic coordinates and equivalent isotropic displacement parameters (pm2) for the crystal structures of Na3CoS3 (above) and Na5[CoS2]2(Br) (below).

AtomWyck. pos.xyzUeq.
Na(1)4a00.23195(9)0.75393(19)348(5)
Na(2)4a00.56700(9)0.07258(17)292(4)
Na(3)8b0.22024(14)0.21726(6)0.45409(13)339(3)
Na(4)8b0.22592(12)0.17246(6)0.17106(12)261(2)
Na(5)8b0.23505(15)0.41955(6)0.16545(12)285(3)
Na(6)8b0.29427(11)0.37627(6)0.42259(11)221(2)
Na(7)8b0.29732(13)0.03698(6)0.36524(13)303(3)
Co(1)4a00.04654(3)0.22191(4)157(1)
Co(2)4a00.00903(2)0.00954(5)166(1)
S(12)8b0.20462(7)0.04352(3)0.10844(6)161(2)
S(21)4a00.04123(5)0.82836(10)193(2)
S(22)4a00.12877(5)0.34277(11)264(2)
S(23)4a00.09809(5)0.51032(11)287(3)
Co(3)4a00.33289(2)0.34362(5)148(1)
Co(4)4a00.29242(2)0.11895(4)148(1)
S(34)8b0.20367(7)0.29615(3)0.23975(6)154(2)
S(31)4a00.30486(5)0.53128(9)186(2)
S(32)4a00.44032(5)0.33244(10)183(2)
S(41)4a00.20926(5)0.00429(10)175(2)
S(42)4a00.38100(5)0.00001(10)233(2)
S(43)4a00.64020(5)0.32543(10)223(2)
Na(1)8c0.2166(4)x0.3606(15)355(12)
Na(2)2a000250(20)
Co4b01/20.3065(13)185(5)
S8d0.3098(2)00.0573(15)189(6)
Br2a000.5347(14)374(7)

The same sodium source Na2S (together with Co, S and NaBr) and Tmax of 1100°C were used to obtain the bromide Na5[CoS2]2(Br) in X-ray pure form. Samples consisting of, e.g. 353.2 mg/4.53 mmol Na2S, 267.5 mg/4.54 mmol Co, 145.6 mg/4.54 mmol S and 233.3 mg/2.27 mmol NaBr yielded homogeneous reguli of black-metallic shiny needles, which became dark-brown transparent when broken into thin pieces. Attempts to obtain the OH or SH derivatives of Na5[CoS2]2(Br) following this protocol and adding NaOH/NaSH instead of NaBr resulted in the formation of the hydrogen-free products Na2[CoS2] and Na6[CoS4] only. At a much lower Tmax of 400°C, Na5[CoS2]2(SH) could finally be obtained (together with Na6[CoS4] and Co9S8) from 5:2 mixtures of NaSH and elemental cobalt. Unfortunately, the hydrogensulfide was yielded in the form of dark-gray powder only. The Rietveld refinement of the powder data resulted in the lattice parameters a=913.41(8) and c=621.94(7) pm, which agree within 0.2% with the crystal data reported by Klepp and Bronger for Na5[CoS2]2(S) [11]. In an argon atmosphere Na5[CoS2]2(SH) decomposes at 600°C into Na6[CoS4] and Co9S8.

2.2 Crystal structure determination

For the single crystal structure analyses, dark-metallic crystals of Na3CoS3 and tiny needles of Na5[CoS2]2(Br) were selected using a stereo microscope. The crystals, which are very sensitive to moisture and air, were fixed in glass capillaries (diameter <0.1 mm) under dried paraffine oil and centered on a diffractometer equipped with a microsource and a CCD area detector (Na3CoS3) or an image plate detector [Na5[CoS2]2(Br)].

The reflections of the crystals of Na3CoS3 could be indexed using an orthorhombic C-centered lattice. The extra extinction conditions ‘reflections h0l only present for h and l=2 n’ and ‘reflections 00l only present for l=2 n’ restricted the possible space groups to Cmcm and Cmc21. The structure solution using Direct Methods (program Shelxs-2013 [23]) was successful in the noncentrosymmetric space group Cmc21 only. The solution already yielded the four cobalt, all sulfur and four of the seven sodium atomic sites. The positions of the three missing Na+ cations could be easily determined from the difference Fourier maps calculated after the first structure refinements (program Shelxl-2013 [23]). After standardization (program Structure Tidy [24]) the atomic parameters and anisotropic displacement parameters (ADP) could be refined to a final R1 value of 0.0205.

The diffraction images of Na5[CoS2]2(Br) show a tetragonal pattern of high Laue symmetry. The reflections could be indexed by the expected tetragonal I-centered lattice and do not show any extra extinction conditions. The lattice parameter a virtually coincides with the value observed for Na5[CoS2]2(S) (915.0 pm, [11]); the c parameter is only slightly (0.5%) increased compared to the sulfide (622.2 pm, [11]). Using the atomic parameters of the I4mm structure model of Na5[CoS2]2(S) as starting values [11] and substituting the isolated sulfur position by bromine, the atomic and ADP parameters of Na5[CoS2]2(Br) could be directly refined to a conclusively low R1 value of 0.0410 with the methods and programs described above. Whereas the absolute structure of Na3CoS3 could be reliably determined, all examined crystals of Na5[CoS2]2(Br) needed to be refined as inversion twins.

The crystal data and the results of the anisotropic structure refinement of both title compounds are collected in the Tables 1 and 2 (see also [25]). Selected interatomic distances can be found in Tables 4 and 5 (for Na3CoS3) and in Table 6 [for Na5[CoS2]2(Br)].

2.3 Band structure calculations

DFT calculations of the electronic band structure were performed for the title compounds Na3CoS3 and Na5 [CoS2]2(Br) and the reference phase Na2[CoS2] (K2ZnO2-type structure, crystal data taken from [5]) using the FP-LAPW method (program Wien2k [26]). The exchange-correlation contribution was described by the generalized gradient approximation (GGA) of Perdew, Burke and Ernzerhof [27]. Muffin-tin radii were chosen as 111 pm (2.1 a.u.) for Na, Co and S, and 103 pm (1.95 a.u.) for the sulfur atoms of the disulfide groups in Na3CoS3. Cutoff energies used are Emaxpot=190 eV (potential) and Emaxwf=116 eV (interstitial PW). In the spin-polarized GGA+U calculation the additional Coulomb potential (Hubbard parameter Ueff) used for the strongly correlated Co d electrons was 2 eV [28], [29]. This value was chosen following the proven parameters, which were used for the calculation of other Fe and Co chalcogenides [28], [29]. In addition, the SIC (self-interaction correction) [30], [31] double counting correction method has been applied. Due to the magnetic ordering experimentally observed in several reference compounds, e.g. K2[CoS2] [5], an antiferromagnetic spin ordering was applied for the cobalt cations in the dimer and along the chains. For this AFM treatment of the Co spins, the crystal structures of Na5[CoS2]2(Br) and Na2[CoS2] (space group Ibam) were transformed into the appropriate subgroups P4mm and I222 (cf. Section 3.4 for details). A Bader analysis of the electron density maps was performed to evaluate the charge distribution between the atoms [32]. Electron and spin densities were calculated and visualized using the programs XCrysDen [33] and DRAWxtl [34]. Further parameters and selected results of the calculations are summarized in Table 3. The total (tDOS) and selected partial Co and S density of states (pDOS) are depicted in Fig. 5.

Table 3:

Details and selected results of the calculation of the electronic structures of Na2[CoS2], Na5[CoS2]2(Br) and Na3CoS3.

CompoundNa2[CoS2]Na5[CoS2]2BrNa3CoS3
Crystal data[5]Tables 1 and 2
Space groupI222P4mmCmc21
Rmt/pmNa, Co, S, Br: 111
S of S2 ligands: 103
Rmt·Kmax8.0
k-points/BZ1000768847
k-points/IBZ17060144
Monkhorst-Grid10×10×108×8×1211×11×7
DOS plot s. Fig. 5top leftbottom leftbottom right
ρBCP/e×10−6 pm−3 (d/pm)
a0.464 (232.8)0.461 (233.6)0.534 (226.1)
b0.455 (234.3)0.495 (229.5)
c0.497 (229.9)
d0.528 (227.3)
e0.507 (228.7)
f0.462 (233.2)
g0.476 (232.1)
h0.462 (232.8)
i0.453 (234.3)
j0.526 (226.9)
k0.485 (230.8)
l0.406 (239.5)
r0.886 (210.8)
s0.842 (213.4)
qBader (VBB/106 pm3)
 Na(1)+0.838 (10.3)+0.837 (10.3)+0.835 (10.3)
 Na(2)+0.840 (9.7)+0.834 (10.2)
 Na(3)+0.855 (10.4)
 Na(4)+0.831 (9.8)
 Na(5)+0.839 (9.9)
 Na(6)+0.823 (9.3)
 Na(7)+0.851 (10.2)
 Co(1)+0.742 (13.2)+0.793 (14.3)+0.721 (12.8)
 Co(2)+0.726 (13.0)
 Co(3)+0.738 (13.2)
 Co(4)+0.744 (13.5)
 S/S(12)−1.209 (35.2)−1.231 (33.6)−1.167 (32.1)
 S(21)−1.165 (32.4)
 S(22)−0.716 (30.0)
 S(23)−0.679 (32.2)
 S(31)−1.356 (37.1)
 S(32)−1.377 (37.8)
 S(34)−1.218 (33.0)
 S(41)−1.336 (35.4)
 S(42)−0.772 (32.1)
 S(43)−0.828 (34.5)
 Br−0.870 (46.3)

3 Results and discussion

3.1 Syntheses

The new chain sulfido cobaltate(II) Na3CoS3 was obtained in the course of a systematic synthetic study on the phase formation of sodium sulfido cobaltates, starting from mixtures of elemental Co and S together with elemental sodium or Na2S as sodium sources. For the first time, the phase appeared in samples composed of the three pure elements in the sulfur-rich composition range of the ternary system Na–Co–S. Later, it could be obtained as a nearly X-ray pure phase in targeted syntheses from stoichiometric mixtures of Na2S, cobalt and sulfur at a maximum temperature of 1100°C (cf. Experimental Section 2.1).

Due to the doubt about the exact chemical composition of the mixed valent Co(II/III) chain compound Na5[CoS2]2(S) [11] and its potential SH content, the isotypic bromine Na5[CoS2]2(Br) was successfully synthesized using a similar temperature program and the same sodium source (together with NaBr) to obtain crystals suitable for a single crystal structure analysis. At a decreased maximum temperature of 400°C the hydrogen sulfide Na5[CoS2]2(SH) could also be obtained when substituting NaBr by NaSH, unfortunately only in the form of powdered material. In contrast to the hydrogen-free sulfido cobaltate, Na5[CoS2]2(SH) decomposes at 600°C under an argon atmosphere, which is not inconsistent with the original synthesis of ‘Na5[CoS2]2(S)’ at 730°C in a stream of hydrogen sulfide. The lattice parameters of Na5[CoS2]2(SH) refined by the Rietveld method from powder data are in perfect agreement with those reported in the single crystal study of ‘Na5[CoS2]2(S)’ by Klepp and Bronger [11].

3.2 Crystal structure description of Na3CoS3

According to Na3CoS3=Na12[Co2S5(S2)][Co2S3(S2)] the orthorhombic structure of the first title compound contains two different types of sulfido/disulfido cobaltate anions (Fig. 1), which both exhibit m symmetry in occupying the mirror plane of the space group Cmc21 (Fig. 2b). The cobalt atoms Co(1) and Co(2) together with the sulfido and disulfido ligands S(1X) and S(2X) form chains [Co2S3(S2)]4− (Fig. 1b), whereas Co(3), Co(4), S(3 X) and S(4 X) form the dinuclear units [Co2S5(S2)]8− depicted in Fig. 1a. The sulfur atom labels are chosen to indicate the actually connected cobalt atom(s). The interatomic distances within the two anions are labeled using small bold characters and are collected in Table 4.

Fig. 1: Ortep representations (90% probability ellipsoids [35]) of the two sulfido cobaltate anions in the crystal structure of Na3CoS3. (a) Dimers [Co2IIS5 (η1−S2)]8−; (b) chains [Co2IIS3 (µ−1,2−S2)]4− (cf. Table 4 for the labeled interatomic distances).
Fig. 1:

Ortep representations (90% probability ellipsoids [35]) of the two sulfido cobaltate anions in the crystal structure of Na3CoS3. (a) Dimers [Co2IIS5 (η1−S2)]8−; (b) chains [Co2IIS3 (µ−1,2−S2)]4− (cf. Table 4 for the labeled interatomic distances).

Fig. 2: Different views of the crystal structure of Na3CoS3. (a) ‘Layers’ around the mirror plane (x≈0) perpendicular to the a axis of the space group Cmc21; (b) hexagonal rod packing along the c axis; (c) perspective view of the overall structure; (S: green balls; Co: red balls; Na: yellow balls; [CoS4]: light-gray (anion a) and dark-gray (anion b) polyhedra, [34]).
Fig. 2:

Different views of the crystal structure of Na3CoS3. (a) ‘Layers’ around the mirror plane (x≈0) perpendicular to the a axis of the space group Cmc21; (b) hexagonal rod packing along the c axis; (c) perspective view of the overall structure; (S: green balls; Co: red balls; Na: yellow balls; [CoS4]: light-gray (anion a) and dark-gray (anion b) polyhedra, [34]).

Table 4:

Selected interatomic distances (pm) in the crystal structure of Na3CoS3.

atomsdistancemult.lbl.CNatomsdistancemult.lbl.CNatomsdistancemult.lbl.CN
Co(1) – S(12)226.1(1)aS(12) – Co(1)226.1(1)aS(22) – S(23)210.8(2)r
– S(21)229.5(1)b– Co(2)228.7(1)e– Co(1)229.9(1)c
– S(22)229.9(1)c4– Na(2)269.6(1)– Na(4)301.6(2)
– Co(2)266.3(1)u– Na(6)282.3(2)– Na(3)304.5(2)
– Co(2)363.9(1)v– Na(5)283.4(2)– Na(7)331.3(1)1+1+6
– Na(4)291.1(2)
Co(2) – S(21)227.3(1)d– Na(7)317.5(2)S(23) – S(22)210.8(2)r
– S(12)228.7(1)e– Na(7)348.8(2)0+2+6– Co(2)233.2(1)f
– S(23)233.2(1)f4– Na(5)301.0(2)
– Co(1)266.3(1)uS(21) – Co(2)227.3(1)d– Na(3)331.3(2)
– Co(1)363.9(1)v– Co(1)229.5(1)b– Na(7)341.7(2)1+1+6
– Na(6)279.3(2)
– Na(5)316.2(2)
– Na(7)316.3(1)0+2+6
Co(3) – S(31)232.1(1)gS(31) – Co(3)232.1(1)gS(41) – Co(4)226.9(1)j
– S(34)232.8(1)h– Na(2)283.3(2)– Na(6)278.0(1)
– S(32)234.3(1)i4– Na(3)287.7(2)– Na(4)293.1(2)
– Co(4)282.2(1)w– Na(4)298.3(1)– Na(3)300.6(1)
– Na(1)309.5(2)– Na(1)302.8(3)0+1+7
Co(4) – S(41)226.9(1)j– Na(6)329.7(1)0+1+8
– S(34)230.8(1)kS(42) – S(43)213.4(2)s
– S(42)239.5(1)l4S(32) – Co(3)234.3(1)i– Co(4)239.5(1)l
– Co(3)282.2(1)w– Na(7)279.2(2)– Na(5)298.7(2)
– Na(2)287.0(2)– Na(7)299.8(2)
S(34) – Co(4)230.8(1)k– Na(5)291.5(2)– Na(3)331.6(2)1+1+6
– Co(3)232.8(1)h– Na(6)314.2(1)0+1+7
– Na(1)269.6(1)S(43) – S(42)213.4(2)s
– Na(4)282.2(2)– Na(1)291.1(2)
– Na(5)284.3(2)– Na(7)291.4(2)
– Na(6)290.6(2)– Na(4)312.4(1
– Na(3)308.4(2)– Na(3)336.0(2)
– Na(3)348.7(2)0+2+6– Na(2)341.2(2)1+0+8
  1. In the anions [Co2S5(S2)]8− (Fig. 1a), two [Co(3,4)S4] tetrahedra are connected via a common [S(34)2] edge. Whereas Co(3) is additionally coordinated by the two terminating S(31) and S(32) ligands (Co–S distances g, i), Co(4) exhibits only one terminal sulfido ligand [S(41)] and the fourth tetrahedron corner is occupied by the η1-disulfido ligand [S(42)–S(43)]2−. The S–S distance of dS(42)S(43)s=213.4 pm within this S2 dumbbell compares to other disulfido ligands (s. further discussion below). All Co(3)–S distances (g, h, i) are found in the expected narrow range 232.1 to 234.3 pm. In contrast, the terminal Co(4)–S(41) distance j is considerably decreased to 226.9 pm, an effect which is also observed in the chain anion for the Co(II) ions carrying disulfido ligands in addition (see below). As expected, the Co–S(42) bond l towards the η1-S2 dumbbell is significantly larger (239.5 pm) than the Co–S2− bond lengths. Due to the edge-sharing and the kink in the double tetrahedra, the Co–Co distance w (282.2 pm) is somewhat shorter than the value in the simple linear chains of Na2[CoS2] (292.5 pm, [5]). To the best of our knowledge, such double tetrahedra with one η1-disulfido ligand represent a new anion among sulfido metalates.

  2. The chain anions [Co2S3(S2)]4− in the structure of Na3CoS3 (Fig. 1b) are formed by [Co(1,2)S4] tetrahedra, which are alternately connected via common sulfido edges [S(12)2] and corners [S(21)]. At these common corners, the tetrahedra are additionally bridged by the μ-1,2-disulfido ligands [S(22)−S(23)]2−. In an equivalent description, the [CoS4] tetrahedra are alternatingly connected via sulfido [(μ-S)2] and mixed sulfido/disulfido [(μ-S)(μ-S2)] edges forming slightly folded Co2S2 four-membered and planar Co2S3 five-membered rings. Thus, the formula of the chain can be summarized after Niggli as 1[Co(S)3/2(S2)1/2]2=1[Co2S3(S2)]4 and belongs to a general series of chain metalates [M(μ-S)2−x(μ-S2)x]n. Similar chains of [MS4] tetrahedra, which are alternatingly bridged by [(S2−)2] and [(S2−)(S22−)] edges, are also found in the gallates Cs2Ga2S5 [36] and Cs2Ga2Se5 [37] and in a series of chalcogenido gallates and indates with organic ammonium cations and metal complexes as countercations ([38] and references therein). The Co–(S2−) distances (a, b, d, e) in the chain anion [Co2S3(S2)]4− of Na3CoS3 are in the narrow range from 226.9 to 229.5 pm (Table 4). As already detailed above for the dinuclear anion, these values are considerably shorter than Co–S contacts in simple chain sulfido cobaltates(II), evidently due to the further coordination by disulfido ligands. The Co–(S2)2− distances c and f of 229.9 and 233.2 pm are again somewhat larger than those involving simple sulfide ions. The Co–Co distances dCo(1)Co(2)u=266.3 pm between the centers of tetrahedra sharing sulfido edges are much shorter than in the simple chain compounds A2[CoS2] and in Na5[CoS2]2(Br) or Na5[CoS2]2(SH) (Section 3.3). Compared to layered cobaltate anions like those found in the structure of KCo2S2 (dCo−Co=263.7 pm [6], [9]) the observed CoII–CoII distances are again within the already known limits.

Compared to free disulfide ions (e.g. in Na2S2), where dS−S amounts to 215.8 pm [39], the S–S bond lengths decrease on increasing bonding of the S22− ligands: In the η1-ligand of the dinuclear anion (a) the bond length (s) is reduced to 213.4 pm. In the bridging disulfide ligand of the chain anion (b) the S–S distance (r) is further decreased to 210.8 pm. This characteristic feature of dichalcogenide ligands is sufficiently documented in the literature [38], [40]. It indicates an increased π donation of S22−, which leads to a slight depopulation of the π* states of the ligand and additionally causes the – compared to monosulfido metalates – decreased band gaps of corresponding disulfido metalate salts (cf. Section 3.4).

Fig. 2a shows a projection of the orthorhombic unit cell of Na3CoS3 along the a axis with both anions located at the mirror plane at x=0. The chains [Co2S3(S2)] (b) are aligned along [001], i.e. along the direction of the 21 screw axis, at y≈0. The double tetrahedra (a) are lined up in the same direction, whereby their two [CoS4] tetrahedra match four [CoS4] tetrahedra of the chain. Along [010], two such columns (at y13and23) are arranged between the chains, which finally leads to an overall 1:1 ratio of dinuclear [Co2IIS5 (S2)]8− and chain segments [Co2IIS3(S2)]4− (=Na12Co4S12=4×Na3CoS3) in each layer. According to the C-centered lattice, the adjacent layers (at x±12) are shifted by 12 in the a and b direction, which becomes obvious in the view of the structure along the chain direction [001] shown in Fig. 2b. This projection also illustrates the pseudo-hexagonal rod packing, which is a characteristic feature of all sulfido metalates containing chains of edge-sharing tetrahedra [41], [42], and also for cutouts of such chains, e.g. the tetraferrates like (Rb/Cs)8 [Fe4S10] [12], [13]. Each chain anion (dark-gray polyhedra in Fig. 2) is surrounded by six rods of aligned double tetrahedra (light-gray polyhedra). In contrast to the distribution of the countercations in the simple chain metalates, the Na+ ions in Na3CoS3 are not homogeneously interspersed among the chains/aligned dimers: With the only exception of Na(1) and Na(2), which occupy m/4 a positions, the sodium cations are arranged in puckered layers (x=0.22 – 0.30) between the Co/S ‘layers’ around the mirror planes (cf. projection of the structure in Fig. 2b, Na(1) and Na(2) are hidden by the anions). The seven crystallographically distinct cations in the structure of Na3CoS3 are coordinated by five to eight sulfur atoms (Table 5); the belonging arrangements are depicted in Fig. 3. The shortest Na–S distance of 269.6 pm is slightly shorter than the sum of Shannon’s ionic radii of 286 pm (for CN=6, [43]). The effective coordination numbers (ECoN) of the sodium cations are within the typical range observed in sodium salts: For Na(1) and Na(2) located at the mirror plane the ECoN values are with 4.31/4.88 somewhat reduced compared to the coordination numbers of the remaining Na+ ions (5.09 to 6.95, Table 5).

Table 5:

Selected interatomic distances Na–S (pm) and coordination numbers (CN) together with the effective coordination numbers (ECoN, in parentheses) in the crystal structure of Na3CoS3 (cf. Figure 3 for the coordination spheres).

atomsdistanceCNatomsdistanceCNatomsdistanceCNatomsdistanceCN
Na(1) – S(34)269.6(1)Na(2) – S(12)269.6(1)Na(3) – S(31)287.7(2)Na(4) – S(34)282.2(2)
– S(43)291.1(2)– S(31)283.3(2)– S(41)300.6(1)– S(12)291.1(2)
– S(41)302.8(3)– S(32)287.0(2)– S(22)304.5(2)– S(41)293.1(2)
– S(31)309.5(2)5– S(43)341.2(2)4+1– S(34)308.4(2)– S(31)298.3(1)
(4.31)(4.88)– S(23)331.3(2)– S(22)301.6(2)
– S(42)331.6(2)– S(43312.4(1)5+1
– S(43)336.0(2)(5.91)
– S(34)348.7(2)4+4
(6.95)
Na(5) – S(12)283.4(2)Na(6) – S(41)278.0(1)Na(7) – S(32)279.2(2)
– S(34)284.3(2)– S(21)279.3(2)– S(43)291.4(2)
– S(32)291.5(2)– S(12)282.3(2)– S(42)299.8(2)
– S(42)298.7(2)– S(34)290.6(2)– S(21)316.3(1)
– S(23)301.0(2)– S(32)314.2(1)– S(12)317.5(2)
– S(21)316.2(2)5+1– S(31)329.7(1)4+2– S(22)331.3(1)
(5.72)(5.09)– S(23)341.7(2)
– S(12)348.8(2)3+5
(5.93)
Fig. 3: Sodium coordination spheres in the crystal structure of Na3CoS3 (thick/thin red lines: distances below/above 310 pm, [34]). (a) Na(1) and Na(2); (b) Na(3) to Na(5); (c) Na(6) and Na(7).
Fig. 3:

Sodium coordination spheres in the crystal structure of Na3CoS3 (thick/thin red lines: distances below/above 310 pm, [34]). (a) Na(1) and Na(2); (b) Na(3) to Na(5); (c) Na(6) and Na(7).

3.3 Crystal structure description of Na5[CoS2]2 (X) (X=Br, SH)

The second title compound, Na5[CoS2]2(Br), is a ‘double salt’ containing linear SiS2-analogous chains and isolated bromide ions. Figure 4 shows the two anions (a, b) and a perspective view of the unit cell (c). In the tetragonal acentric structure, sulfido cobaltate chains [CoS4/2]2− of edge-sharing tetrahedra, comparable to those in the common chain compounds of the K2ZnO2-type structure, like e.g. Na2[CoS2], are running along the tetragonal c axis at 0, 12,z. The Br ions are located in between these chains. They are surrounded by five Na+ cations forming square pyramids (point group symmetry 4mm, blue polyhedra in Fig. 4c). The pyramids are equally oriented along [001], with one [BrNa5] pyramid per two [CoS4] tetrahedra of the cobaltate chains. The structure of Na5[CoS2]2(Br) is isotypic to the reported mixed-valent phase ‘Na5[CoS2](S)’ ([11], see Discussion below). Both compounds formally belong to the Na6PbO5-type structure, with the unusual assignment (Na6PbO5≡Na5Co2S4Br) according to Pb≡Br (2 a) and O(1)≡Na(2) (2 a), but O(2)≡S (8 d), Na(2)≡Co (4 b) and Na(1)≡Na(1) (8 c).

Fig. 4: Crystal structure of Na5[CoS2]2(Br). (a, b: Ortep representations with 90% probability ellipsoids [35]; cf. Table 6 for the interatomic distances; c: unit cell: green balls: S; red balls: Co; yellow balls: Na, [34]).
Fig. 4:

Crystal structure of Na5[CoS2]2(Br). (a, b: Ortep representations with 90% probability ellipsoids [35]; cf. Table 6 for the interatomic distances; c: unit cell: green balls: S; red balls: Co; yellow balls: Na, [34]).

The sulfido cobaltate chain in the tetragonal structure of Na5[CoS2]2(Br) is fully comparable to the similar chain anion present in the orthorhombic phase Na2[CoS2] [5]: The Co–S distances of 233.6 (a) and 234.3 pm (b, Table 6) (Co site symmetry: 2mm/C2v) are well comparable to the values found for Na2[CoS2] (dCo−S=232.8 pm; 222/D2h). In contrast, Na5[CoS2]2(SH) exhibits two significantly different Co–S distances along the chains (dCo−S=226.3 and 240.4 pm); the mean value of 233.4 pm nevertheless coincides with the Co–S distances in the common chain sulfido cobaltates(II). Apparently, the underlying shift of the S2− ligand enables the formation of S–H···S contacts between the SH groups and the surrounding sulfido ligands of the chain through the trianglar faces of the square pyramid (dS−S=418.3 pm). Due to the incorporation of the extra Br/SH ions in between the cobaltate anions, the Co–Co distances of 312.8/311.1 pm in the linear chain of Na5[CoS2]2(Br) are significantly enlarged compared to the values in Na2[CoS2] (dCo−Co=292.5 pm, [5]). Characteristic for trans edge-connected tetrahedra, the ∠S−Co−S of the two connecting edges (95.9/96.3° (X=Br) and 92.6/100.3° (SH) are much smaller than the four of the non-connecting tetrahedra edges (≈116°, 4×). In agreement with the increased Co–Co distances, this stretching of the tetrahedra along the 4̅ axis is even more pronounced in the ‘double salts’ than in the reference structure of Na2[CoS2] [102.2° (2×) and 110.1/116.4° (4×)].

Table 6:

Selected interatomic distances (pm) in the crystal structure of Na5[CoS2]2(Br) (for distance labels cf. Figure 4).

atomsdistancemult.lbl.CNatomsdistancemult.lbl.CNatomsdistancemult.CN
Co – S233.6 (3)aS – Co233.6 (3)aBr – Na (2)291.1 (9)
  – S234.3 (3)b4 – Co234.3 (3)b  – Na (1)300.6 (6
  – Co312.8u – Na (2)285.6 (1)  – Na (2)334.5 (9)5+1
 – Na (1)287.2 (4)
 – Na (1)287.9 (4)2+5
Na(1) – S287.2 (4)Na (2) – S285.6 (2)
   – S287.9 (4)   – Br291.1 (9)
   – Br300.6 (6)5   – Br334.5 (9)5+1

The extra anions X are coordinated by a square pyramid of sodium cations with X–Na distances of 291.1 and 300.6 pm for the bromide and the hydrogensulfide, respectively. A further Na+ ion is located above the square face of the pyramid at a distance of 334.5 pm (Fig. 4b). The Br–Na+ and SH–Na+ distances as well as the coordination numbers of X are thus in good agreement with the values found in the (for X=SH: distorted) rocksalt structures of NaSH (299.2 pm) and NaBr (298.1 pm). The sum of the Shannon radii [43] of 299 pm (for Na+/Br) are also similar.

The sodium coordination numbers in both X-containing compounds of 5 (4 S2−+1 Br) for Na(1) and 5+2 (4 S2−+(1+1) Br) for Na(2) (Table 6) also fit the values in simple binary sodium salts, as well as in Na2[CoS2] and Na3CoS3. The Na–X distances (dNa−S=289−333 pm, average value: 301 pm) perfectly compare to the distances observed in the structure of NaBr (298.1 pm) and NaSH (299.2 pm). In contrast (and despite the enlarged coordination number of 8) the Na+–S2− distance in the sulfide Na2S (anti-CaF2-type structure) is much smaller (283 pm). This provides a further argument for the presence of SH in ‘Na5[CoS2]2(S)’. Unfortunately, the detection of the SH vibrational mode by means of IR or Raman spectroscopy failed due to metallic lustre of the compound.

The molar volumes of Na5[CoS2]2(SH) and Na5[CoS2]2(Br) are practically equal, which agrees with the likewise similar molar volumes of NaSH (52.47×106 pm3) und NaBr (52.98×106pm3). For both compounds, the ‘double salt’ character leads to a nearly perfect match of the observed molar volumes (Vobs) with the values calculated from the doubled volume of Na2[CoS2] and the volume of NaBr/NaSH (X=SH: 261.5/260.5; X=Br: 261.6/262.0×106 pm3 for Vcalc/ Vobs).

3.4 Electronic structure and chemical bonding

The band structures of the title compounds and of the more simple reference chain sulfido cobaltate Na2[CoS2] were calculated using FP-LAPW DFT methods (GGA+ U) including spin-polarization and a suitable Hubbard U parameter. Computational details can be found in the Experimental Section 2.3 and in Table 3. For the orthorhombic structure of Na2[CoS2] a symmetry reduction from space group Ibam to I222 mimics the magnetic space group Ibam’, which was experimentally verified for the potassium compound K2[CoS2] by a Rietveld refinement of neutron powder diffraction data [5]. For the tetragonal structure of Na5[CoS2]2(Br) a symmetry lowering from space group I4mm to P4mm was applied to enable the antiferromagnetic ordering of the Co spins along the chains. In the case of Na3CoS3 the implementation of the AFM ordering of Co spins within the dinuclear unit as well as along the chain (which are not yet experimentally verified) is possible without a symmetry reduction [Co(1)/Co(3): ↑ and Co(2)/Co(4): ↓].

The total (tDOS, gray shaded) and selected partial (pDOS) Co and S density of states of Na2[CoS2], Na5[CoS2]2(Br) and Na3CoS3 are depicted in Fig. 5. Consistently, the tDOS of the three sodium salts show definite band gaps and a valence band region of approx. 5 eV formed by Co- d and S- p states. The conduction band is equally determined by the single-occupied Co-dxy/dyz/dxz states of the upper Hubbard band. Thus, in the ZSA scheme [44] all compounds belong to the class 2B of ‘charge transfer semiconductors’. In all compounds, the Bader charges (q) and volumes (VBB) as well as the heights of the bond critical points at the Co–S bonds (ρBCP) are similar: The sodium and cobalt cations exhibit the expected positive charges (Na+: q=+0.82 – +0.86; VBB=9.7 – 10.4 e×10−6 pm−3; Co2+: q=+0.72 – +0.79; VBB=12.8 – 14.3 e×10−6 pm−3, Table 3). The sulfur charges differ only for Na3CoS3 with its mixed sulfide/disulfide ligands (see below). For all compounds, the values of ρBCP for the Co–S bonds are found in a narrow range (ρBCP=0.41 – 0.53 e×10−6 pm−3) and steadily decrease with increasing bond lengths (Table 3). They are somewhat smaller than the values calculated for sulfido ferrates(III) (e.g. Cs[FeS2]: 0.57 – 0.61 e×10−6 pm−3), which is evidently only an effect of the reduced ionic radius of Fe(III) compared to Co(II). These values also agree with those found for pyrite in an experimental electron density study (ρBCP=0.53 e×10−6 pm−3 [45]). The Laplacian of the electron density at the BCPs (∇2ρBCP), which is discussed as an additional indicator for the character of the bonding, is also very comparable in the cobaltates (Na2[CoS2]: +3.63 e×10−10 pm−5) and the ferrates(III) (Cs[FeS2]: +3.55–3.80 e×10−10 pm−5). They are significantly smaller than the experimental as well as the calculated values of ∇2ρBCP in pyrite (Fe(II) in an octahedral coordination, exp.: 6.20 e×10−10 pm−5 [45]) indicating an increased covalency for the tetrahedral coordination.

Fig. 5: Calculated total density of states (tDOS, gray) together with the Co and S partial DOS (colored lines) of Na2[CoS2] (a), Na5[CoS2]2(Br) (b) and Na3CoS3 (d) in the range between −4.9 and 3.8 eV relative to EF. (c): schematic representation of the different Co d state energies; possible superexchange path; calculated spin density map in Na2[CoS2] at a ±0.01 e−×10−6pm−3 level.
Fig. 5:

Calculated total density of states (tDOS, gray) together with the Co and S partial DOS (colored lines) of Na2[CoS2] (a), Na5[CoS2]2(Br) (b) and Na3CoS3 (d) in the range between −4.9 and 3.8 eV relative to EF. (c): schematic representation of the different Co d state energies; possible superexchange path; calculated spin density map in Na2[CoS2] at a ±0.01 e×10−6pm−3 level.

Na2[CoS2] contains 19 valence electrons per formula unit (v.e./f.u.), excluding the sulfur 2 s states [2 (2×Na) +9 (Co) +8 (2×S)]. Due to the AFM ordering, the total DOS of the α () and β () spins are equal, but the splitting of the Co sites allows to assign the individual d states of the CoIId7 ions. For Na2[CoS2], the z direction of the local cartesian coordinate system of Co points along the chain direction, x and y are also oriented between the ligands (cf. scheme in Fig. 5, top right). In the calculation of Na5[CoS2]2(Br) x and y are rotated by 45 °. This difference was taken into account for by the coloring of the different d states in the pDOS plots at the left hand side of Fig. 5. As expected, Na2[CoS2] exhibits a distinct band gap of 1.4 eV and the 19 occupied valence states (Z=2), which spread from −4.7 eV to the Fermi level EF, are formed by S2−- p and CoII- d7 orbitals. The position of the individual d states of Co was used to extract the simplified scheme of Fig. 5 (top right). The t2 Co- d orbitals (Mulliken symbols as for an ideal tetrahedron, green/blue/cyan) together with the two a1 S- p ligand SALCs are located at the bottom of the valence band region. This region represents the σ-bonding contribution between S p and the Co d orbitals dxy/dyz/dxz pointing towards the ligands, which are also responsible for the strong antiferromagnetic coupling of the Co spins along the chain (see below). The broad region between −3.7 and −1.7 eV, where the Co- d states of e symmetry (red/orange) and six S- p orbitals coincide, represents the mainly π-bonding contribution between Co dx2y2/dz2 and the ligand p orbitals both of ideal symmetry t2. The upper region of the valence band is formed by the remaining four sulfur orbitals (non/π*-bonding) and the e set of the d-Co states of the upper Hubbard band. The relative energy of the three narrow empty t bands of the upper Hubbard band determines the band gap of the compound. Its value of 1.42 eV fits the dark-metallic color of the crystals of Na2[CoS2].

The occupation of the different d states is reflected in the electron and spin densities in the vicinity of the Co atom. The v. e. density map exhibits the octahedral shape of the fully occupied superimposed e-type orbitals, i.e. the lobes are situated between the Co–S bonds. In contrast, the spin density map is of the ‘cubic’ shape resulting from the single-occupied superimposed t2 atomic orbitals and it thus extends towards the ligands. In Fig. 5, top right, the spin density map is shown at the ±0.01 e×10−6 pm−3 level. In accordance with the comparably long Co–Co distance, the distinct spin density at the ligands clearly shows that the magnetic ordering is due to a superexchange via the sulfur atoms. A simple explanation (e.g. applying the GKA rules [46], [47]) for an AFM ordering via a 90° dt- p- dt superexchange is not possible for tetrahedral d7 ions like Co(II) [48]. The scheme of possible σ bonding interactions of the relevant dxy (cyan) and the dxy+dyz (blue, green) combination of Co show, that only a pz-type ligand orbital is able to mediate the AFM exchange for this configuration. Although the pDOS exhibits a preferred admixture of S pz into the singly occupied Co d states and the sign of the spin density fits a simple pz orbital, the calculated spin distribution exhibits distinct lobes along the Co–S bonds (and thus looks like two pz+px/y ‘orbitals’.)

The calculated magnetic moment of Co (2.33 μB) in Na2[CoS2] is in good agreement with the experimental value of the isotypic potassium compound K2[CoS2] (2.5 μB, [5]) and is somewhat smaller than the expected ‘spin only’ value of 3 μB. A reduced magnetic moment of the 3 d metal is observed in most of the sulfido metalates containing [MS4] tetrahedra and was attributed to covalency effects, in comparison with oxides or compounds containing an octahedral coordination of M. In contrast to the complex situation in the d5 ferrates(III), where the reduction of the magnetic moment is pronounced and depends on the connectivity of the [FeS4] tetrahedra [49], the CoII- d7 ions in sulfido cobaltates(II) exhibit a very similar magnetic moment [3], [8], [50].

In accordance with the crystal-chemically shown ‘double salt’ character of Na5[CoS2]2(Br), the electronic structure (Fig. 5, bottom left) is directly comparable to Na2[CoS2] (but note the rotation of x/y and Z=4). The additional pDOS of the Br- p ions (6 v. e., magenta line) is situated between −1.5 and −3.5 eV. The slight increase of the band gap to 1.6 eV on formal addition of the salt NaBr to Na2[CoS2] is reasonable and also becomes apparent from the slight transpacency of the crystals (even regarding the DFT intrinsic underestimation of band gaps).

The DOS of the complex structure of Na3CoS3 (Fig. 5, bottom right) is by far too complicated for an in-depth interpretation. While the Bader charges and volumes of the seven Na+ and the four Co2+ sites of Na3CoS3 are very similar and also comparable to those of the simpler sodium cobaltates, the differences for the S atoms arise from their different bonding situation: Sulfur atoms of bridging ligands exhibit smaller Bader volumes than those of terminal ligands. As expected, the negative charge of the pure sulfido ligand’s atoms (q=−1.16 to −1.38) is larger than the charge of the disulfide atom (q=−0.68 to −0.83). The latter values again agree with the experimentally observed charge of this anion in FeS2 (ρBCP=0.83 e×10−6 pm−3, [45]). The band gap of Na3CoS3 is slightly reduced with respect to the pure sulfido cobaltates(II) (1.14 eV). A similar trend has been observed in several main group metalates containing dichalcogenide ligands; it was traced to Q-π contributions of these ligands to the top of the valence band [38].

4 Summary

The new sulfido cobaltates(II) Na3CoS3 and Na5[CoS2]2(Br) were synthesized at 1100°C from stoichiometric mixtures of Na2S, cobalt and sulfur (and NaBr). The isotypic SH salt was obtained from NaSH and Co at 400°C and is most likely identical to the phase described in the literature as mixed valent sulfide, Na5[CoS2]2(S). The bromide and the hydrogensulfide Na5[CoS2]2(X) are ‘double salts’ composed of linear edge-sharing tetrahedra chains [CoS4/2]2− (like in Na2[CoS2]) and isolated X anions. The new complex orthorhombic structure of Na3CoS3 (=Na12[Co2S5(S2)][Co2S3(S2)]) contains dinuclear units [Co2S5(S2)]8− of two edge-sharing [CoS4] tetrahedra with five sulfido and one η1-disulfido ligands, which is a new type of anion among sulfido metalates. The second type of anions of Na3CoS3 are chains 1[Co2S3(S2)]4 of [CoS4] tetrahedra connected via μ-sulfido and μ-1,2-disulfido ligands. Similar chains have been observed in Ga and In chalcogenides, but are a new structure element in transition metal chalcogenides, too.

DFT studies of the two title compounds and the reference Na2[CoS2] exhibit comparable DOS and band gaps, uniform Bader charges and volumes of all atoms, and similar magnetic moments for the Co(II) ions, which are also in good agreement with experimental values reported in the literature. Even though the semiempirical GKA rules fail to explain the AFM order for a 90° superexchange of tetrahedral d7 ions, the calculated spin density map of Na2[CoS2] clearly substantiates the participation of the ligand p states in the superexchange process.

Acknowledgments

We would like to thank the Deutsche Forschungsgemeinschaft for financial support.

References

[1] K. Klepp, W. Bronger, Z. Naturforsch.1983, 38b, 12.10.1515/znb-1983-0105Search in Google Scholar

[2] W. Bronger, C. Bomba, J. Less-Common Met.1990, 162, 309.10.1016/0022-5088(90)90346-LSearch in Google Scholar

[3] W. Bronger, C. Bomba, W. Koelman, Z. Anorg. Allg. Chem.1995, 621, 409.10.1002/zaac.19956210312Search in Google Scholar

[4] W. Bronger, W. Koelmann, P. Müller, Z. Anorg. Allg. Chem.1995, 621, 412.10.1002/zaac.19956210313Search in Google Scholar

[5] W. Bronger, C. Bomba, J. Less-Common Met.1990, 158, 169.10.1016/0022-5088(90)90444-OSearch in Google Scholar

[6] M. Schwarz, P. Stüble, J. Kägi, C. Röhr, Z. Kristallogr. Suppl.2015, 35, 108.Search in Google Scholar

[7] W. Bronger, P. Böttcher, Z. Anorg. Allg. Chem.1972, 390, 1.10.1002/zaac.19723900102Search in Google Scholar

[8] W. Bronger, U. Hendriks, P. Müller, Z. Anorg. Allg. Chem.1988, 559, 95.10.1002/zaac.19885590109Search in Google Scholar

[9] G. Huan, M. Greenblatt, M. Croft, Eur. J. Solid State Inorg. Chem.1989, 26, 193.Search in Google Scholar

[10] W. Bronger, W. Koelmann, D. Schmitz, Z. Anorg. Allg. Chem.1995, 621, 405.10.1002/zaac.19956210311Search in Google Scholar

[11] K. O. Klepp, W. Bronger, J. Less-Common Met.1984, 98, 165.10.1016/0022-5088(84)90287-XSearch in Google Scholar

[12] M. Schwarz, M. Haas, C. Röhr, Z. Anorg. Allg. Chem.2013, 639, 360.10.1002/zaac.201200397Search in Google Scholar

[13] M. Schwarz, C. Röhr, Inorg. Chem.2015, 54, 1038.10.1021/ic502382vSearch in Google Scholar PubMed

[14] M. Schwarz, C. Röhr, Z. Anorg. Allg. Chem.2015, 641, 1053.10.1002/zaac.201500097Search in Google Scholar

[15] M. Schwarz, C. Röhr, Z. Anorg. Allg. Chem.2014, 640, 2792.10.1002/zaac.201400357Search in Google Scholar

[16] P. Stüble, C. Röhr, Acta Crystallogr. Suppl.2016, A72, s265.Search in Google Scholar

[17] M. Schwarz, P. Stüble, C. Röhr (in preparation).Search in Google Scholar

[18] K. Yvon, W. Jeitschko, E. Parthé, Program Lazy-Pulverix, University Geneve, Switzerland 1976.Search in Google Scholar

[19] A. C. Larson, R. B. V. Dreele, General Structure Analysis System (Gsas), Los Alamos National Laboratory Report LAUR 86–748, 2000.Search in Google Scholar

[20] B. H. Toby, J. Appl. Crystallogr.2001, 34, 210.10.1107/S0021889801002242Search in Google Scholar

[21] G. M. Sheldrick, Sadabs: Program for Absorption Correction for Data from Area Detector Frames, Bruker Analytical X-ray Systems Inc., Madison Wisconsin, USA 2008.Search in Google Scholar

[22] Stoe & Cie GmbH, Darmstadt (Germany), X-Shape (version 1.03), Crystal Optimization for Numerical Absorption Correction 2005.Search in Google Scholar

[23] G. M. Sheldrick, Acta Crystallogr.2008, A64, 112.10.1107/S0108767307043930Search in Google Scholar PubMed

[24] L. M. Gelato, E. Parthé, J. Appl. Crystallogr.1990, A46, 467.Search in Google Scholar

[25] Further details of the crystal structure investigations may be obtained from Fachinformationszentrum Karlsruhe, D-76344 Eggenstein-Leopoldshafen, Germany (fax +49-7247-808-666; e-mail: crysdata@fiz-karlsruhe.de; http://www.fiz-karlsruhe.de/request_for_deposited_data.html) on quoting the deposition numbers CSD 430156 (Na3CoS3) and 430157 (Na5[CoS2]2(Br)).Search in Google Scholar

[26] P. Blaha, K. Schwarz, G. K. H. Madsen, D. Kvasnicka, J. Luitz, Wien2k, An Augmented Plane Wave and Local Orbital Program for Calculating Crystal Properties, TU Wien, Vienna (Austria), ISBN3-9501031-1-2, 2006.Search in Google Scholar

[27] J. P. Perdew, S. Burke, M. Ernzerhof, Phys. Rev. Lett.1996, 77, 3865.10.1103/PhysRevLett.77.3865Search in Google Scholar PubMed

[28] A. Jain, G. Hautier, C. J. Moore, S. P. Ong, C. C. Fischer, T. Mueller, K. A. Persson, G. Ceder, Comp. Mater. Sci.2011, 50, 2295.10.1016/j.commatsci.2011.02.023Search in Google Scholar

[29] S. Selcuk, A. Selloni, J. Phys. Chem. C2015, 119, 9973.10.1021/acs.jpcc.5b02298Search in Google Scholar

[30] V. I. Anisimov, I. V. Solovyev, M. A. Korotin, M. T. Czyzyk, G. A. Sawatzky, Phys. Rev. B1993, 48, 16929.10.1103/PhysRevB.48.16929Search in Google Scholar

[31] A. I. Lichtenstein, V. I. Anisimov, J. Zaanen, Phys. Rev. B1995, 52, R5467.10.1103/PhysRevB.52.R5467Search in Google Scholar

[32] R. W. F. Bader. Atoms in Molecules. A Quantum Theory. International Series of Monographs on Chemistry, Clarendon Press, Oxford 1994.Search in Google Scholar

[33] A. Kokalj, J. Mol. Graphics Modelling1999, 17, 176.10.1016/S1093-3263(99)00028-5Search in Google Scholar

[34] L. W. Finger, M. Kroeker, B. H. Toby, J. Appl. Crystallogr.2007, 40, 188.10.1107/S0021889806051557Search in Google Scholar

[35] M. N. Burnett, C. K. Johnson, Ortep-III. ORNL-6895, Oak Ridge National Laboratory, Oak Ridge, Tennessee (USA) 1996.Search in Google Scholar

[36] D. Friedrich, A. Pfitzner, M. Schlosser, Z. Anorg. Allg. Chem.2012, 638, 1572.10.1002/zaac.201204011Search in Google Scholar

[37] D. Friedrich, M. Schlosser, A. Pfitzner, Z. Anorg. Allg. Chem.2014, 640, 826.10.1002/zaac.201400046Search in Google Scholar

[38] S. J. Ewing, M. L. Romero, J. Hutchinson, A. V. Powell, P. Vaqueiro, Z. Anorg. Allg. Chem.2012, 638, 2526.10.1002/zaac.201200255Search in Google Scholar

[39] P. Böttcher, J. Getzschmann, R. Keller, Z. Anorg. Allg. Chem.1993, 619, 476.10.1002/zaac.19936190309Search in Google Scholar

[40] M. S. Devi, K. Vidyasagar, J. Chem. Soc., Dalton Trans.2002, 40, 4751.10.1039/b207741hSearch in Google Scholar

[41] M. R. Harrison, M. G. Francesconi, Coord. Chem. Rev.2011, 255, 451.10.1016/j.ccr.2010.10.008Search in Google Scholar

[42] W. Bronger, U. Ruschewitz, P. Müller, J. Alloys Compd.1995, 218, 22.10.1016/0925-8388(94)01344-6Search in Google Scholar

[43] R. D. Shannon, Acta Crystallogr.1976, A32, 751.10.1107/S0567739476001551Search in Google Scholar

[44] J. Zaanen, G. A. Sawatzky, J. W. Allen, Phys. Rev. Lett1985, 55, 418.10.1103/PhysRevLett.55.418Search in Google Scholar

[45] M. Schmøkel, L. Bjerg, S. Cenedese, M. R. V. Jørgensen, Y.-S. Chen, J. Overgaard, B. B. Iversen, Chem. Sci.2014, 5, 1408.10.1039/C3SC52977KSearch in Google Scholar

[46] J. B. Goodenough, J.-S. Zhou, Struct. Bond.2001, 98, 17.10.1007/3-540-45503-5_2Search in Google Scholar

[47] J. B. Goodenough, Magnetism and the Chemical Bond, Interscience-Wiley, New York (USA) 1963.Search in Google Scholar

[48] W. Geertsma, D. Khomskii, Phys. Rev. B1996, 54, 3011.10.1103/PhysRevB.54.3011Search in Google Scholar

[49] W. Bronger, Angew. Chem., Int. Ed. Engl.1981, 20, 52.10.1002/anie.198100521Search in Google Scholar

[50] W. Bronger, P. Müller, J. Alloys Compds.1997, 246, 27.10.1016/S0925-8388(96)02459-0Search in Google Scholar

Received: 2016-8-22
Accepted: 2016-9-8
Published Online: 2016-11-24
Published in Print: 2016-12-1

©2016 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 9.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/znb-2016-0179/html
Scroll to top button