Skip to content
BY 4.0 license Open Access Published by De Gruyter (O) August 25, 2020

Syntheses and crystal structures of the manganese hydroxide halides Mn5(OH)6Cl4, Mn5(OH)7I3, and Mn7(OH)10I4

  • Viktoria Falkowski , Alexander Zeugner , Stefan Seidel , Rainer Pöttgen , Klaus Wurst , Michael Ruck and Hubert Huppertz EMAIL logo

Abstract

Single-crystals of the new manganese hydroxide halides Mn5(OH)6Cl4, Mn5(OH)7I3, and Mn7(OH)10I4 were obtained by means of high-pressure/high-temperature synthesis in a Walker-type multianvil apparatus. The chloride crystallizes in the monoclinic space group P21/c (no. 14) with the lattice parameters (single-crystal data) a = 592.55(8), b = 1699.7(2), c = 597.33(8) pm, and β = 112.58(1)°. The iodides crystallize in the triclinic space group P1 (no. 2) with a = 653.16(2), b = 905.98(3), c = 1242.98(4) pm, α = 114.21(1)°, β = 99.91(1)°, and γ = 94.37(1)° for Mn5(OH)7I3 and a = 656.90(3), b = 906.59(4), c = 909.32(4) pm, α = 119.29(1)°, β = 97.99(1)°, and γ = 95.03(1)° for Mn7(OH)10I4. The crystal structures consist of edge-sharing Mn(OH)6–nXn (X = Cl, I; n = 1, 2, 3) octahedra arranged in stacked sheets. Adjacent layers are connected by hydrogen bonds of type O–H···X, confirmed by further characterization of single-crystals by IR-spectroscopy. The crystal chemical relationship with the aristotype Mg(OH)2 (brucite) is discussed on the basis of Bärnighausen trees (group–subgroup relations).

1 Introduction

The most common hydroxide halides of divalent metals have compositions M(OH)X and M2(OH)3X (M = divalent metal, X = Cl, Br, I). Many of them occur as minerals, such as Pb(OH)Cl (laurionite) and Cu(OH)Cl (belloite). Some possess polymorphs as in the case of Cu2(OH)3Cl (atacamite, botallackite, clinoatacamite, and paratacamite) [1], [2], [3], [4]. These compounds are formally double-salts mM(OH)2·nMX2, but with respect to their structures they are better described as halide-substituted forms of the metal hydroxides M(OH)2, indicated by the general formula Mm(OH)2m−nXn. Concerning synthesis and structural characterization, Oswald, Feitknecht, and co-workers were the main contributors, also summarizing their results in several review articles [4], [5], [6], [7], [8]. In addition, vibrational and magnetic properties as well as light absorption behavior of this compound class were topics of intense scrutiny [9], [10], [11], [12], [13], [14], [15], [16], [17], [18]. Lutz et al. investigated laurionite-type hydroxide halides with IR and Raman spectroscopy and defined new distance limits for O–H···X hydrogen bonds [9], [15]. In recent years, the hydroxide halide phases with magnetic 3d transition-metal ions, like Cu2(OH)3X, β-Co2(OH)3Cl, Ni2(OH)3Cl, and β-Fe2(OH)3Cl, gained attention due to their behavior as geometrically frustrated magnetic materials [17], [19], [20], [21], [22], [23], [24]. Their triangular, kagomé, and tetrahedral lattices of magnetically active sites can exhibit various exotic ground states and unconventional magnetic transitions. The compound Co2(OH)3Cl is also considered as auspicious electrode material for application in batteries and supercapacitors because of its outstanding electrochemical performance [25], [26], [27].

As hydroxide halides of the magnetically active 3d transition-metals show intriguing properties, our recent works focused on the less investigated manganese phases of this material class. Here, only Mn2(OH)3Cl, naturally occurring as the mineral kempite, and its bromine counterpart were thoroughly characterized [4], [18], [28]. In the case of the other previously reported phases, i.e., the α- and β-modification of Mn(OH)Cl as well as Mn2(OH)3I, the characterization is limited to their cell parameters and the isotypicity of Mn2(OH)3I to botallackite [4], [7].

Poor crystallinity and a tendency for polymorphism, polytypism, and stacking faults make detailed structural characterizations of the manganese hydroxide halides a challenging task. Taking advantage of the provided opportunities for synthesis and improved crystallization behavior offered by the high-pressure/high-temperature method, we investigated this system under such conditions. In our preliminary work, we could already present a new modification of Mn(OH)Cl and the new compounds Mn(OH)Br and Mn(OH)I, successfully synthesized by this method [29], [30]. In the course of our inquiries, it was possible to identify further representatives of this material class. In this contribution, the syntheses, structural characterizations through single-crystal X-ray diffraction, IR spectroscopy, and the structural relationships of the new phases Mn5(OH)6Cl4, Mn5(OH)7I3, and Mn7(OH)10I4 to Mg(OH)2 are presented.

2 Experimental section

2.1 Synthesis

The compounds Mn5(OH)6Cl4, Mn5(OH)7I3, and Mn7(OH)10I4 were obtained via high-pressure/high-temperature syntheses in a hydraulic 1000 t press (mavo press LPR 1000-400/50, Max Voggenreiter GmbH, Mainleus, Germany) with a modified Walker-type module (also Max Voggenreiter GmbH). The chloride phase could be obtained according to Eq. (1) at 1173 K and 3 GPa from a 3:2 molar mixture of MnO (99%, Sigma Aldrich, Steinheim, Germany) and MnCl2·4H2O (purum, p.a., Fluka, Buchs, Switzerland).

(1)3MnO+2MnCl24H2O3GPa,1173KMn5(OH)6Cl4+H2O

Crystals of the iodide phases Mn5(OH)7I3 and Mn7(OH)10I4 occurred as side-phases from several reactions (all at 2.5 GPa/1273 K) of Mn (99+%, Sigma Aldrich, Steinheim, Germany) and BiOI with different molar ratios, hydrolyzed by wetting with water vapor. Therefore, BiOI was synthesized by adding BiI3 (99.95%, Alfa Aesar, Karlsruhe, Germany) in distilled water and precipitated with a diluted KOH solution. The precipitate was filtered and washed with EtOH and H2O until the resulting BiOI (ICDD card No. 00-73-2062 [31]) was found to be phase pure by PXRD. Possible reactions for the formation of the iodide phase are presented in Eqs. (2) and (3). Further attempts to synthesize phase pure samples of the iodide phases failed so far.

(2)10Mn+6BiOI+8H2O2.5GPa,1273K2Mn5(OH)7I3+6Bi+H2
(3)7Mn+4BiOI+6H2O2.5GPa,1273KMn7(OH)10I4+4Bi+H2

Reactions were carried out in molybdenum capsules (0.025 mm foil, 99.95%, Alfa Aesar, Karlsruhe, Germany) and transferred into crucibles made from h-BN (HeBoSint® P100, Henze BNP GmbH, Kempten, Germany). The crucibles were built into 18/11 assemblies, which were compressed by eight tungsten carbide cubes (HA-7%Co, Hawedia, Marklkofen, Germany). Details on the construction of such assemblies can be found in the corresponding work [32].

Unlike the previously reported compounds Mn(OH)Br and Mn(OH)I [30], these phases, although slightly hygroscopic, showed reasonable air stability and did not require handling (but storage) in an inert atmosphere. The chloride phase shows the characteristic pale rose color of manganese(II) compounds, whereas the Mn5(OH)7I3 and Mn7(OH)10I4 platelets are initially clear, turning slightly yellow after several days of light exposure. The elemental bismuth formed during the reaction was present as small beads and could be separated from the hydroxide halides by hand.

2.2 Single-crystal X-ray diffraction (SCXRD)

For the single-crystal X-ray structure analysis, platelets of Mn5(OH)6Cl4 and Mn5(OH)7I3 were mounted on a Bruker D8 Quest diffractometer equipped with a Photon 100 detector and an Incoatec Microfocus source generator (Mo-Kα radiation with λ = 71.073 pm, multi-layered optics monochromatized) and intensity data was collected at 173 and 190 K, respectively. Single-crystals of Mn7(OH)10I4 were sealed in a glass capillary (0.5 mm, Hilgenberg, Malsfeld, Germany) and SCXRD data were collected with a Bruker CCD Kappa APEXII diffractometer, also using Mo-Kα radiation. Multi-scan absorption corrections based on equivalent and redundant intensities were applied with the program Sadabs-2014/5 [33], [34]. The crystal structures of the hydroxide halides were standardized with the program Structure Tidy[35–37]. The chloride phase shows a substitutional disorder of the ligands Cl1/O1 and Cl3/O4, whereby the occupancy factors were refined close to 0.75 (Cl1 and O4) and 0.25 (O1 and Cl3) and fixed at these values. The protons binding to O1 and O4 could not be detected as they are situated close to the positions of Cl1 and Cl3. For all phases, the O–H distance was restrained to a value of 83 ± 2 pm. All non-hydrogen atoms were refined with anisotropic displacement parameters. Table 1 summarizes the experimental details of the data collection and evaluations. Tables 2–4 contain the positional parameters, anisotropic displacement parameters, and selected interatomic distances and angles of the chloride phase and Tables 5–7 those of the iodide compounds.

Table 1:

Crystal data and structure refinement of Mn5(OH)6Cl4, Mn5(OH)7I3, and Mn7(OH)10I4 (standard deviations in parentheses).

FormulaMn5(OH)6Cl4Mn5(OH)7I3Mn7(OH)10I4
Crystal systemMonoclinicTriclinicTriclinic
Space groupP21/c (no. 14)P1 (no. 2)P1 (no. 2)
Single-crystal data
Single-crystal diffractometerBruker D8 QuestBruker APEXII Kappa CCD
RadiationMo-Kα (λ = 71.073 pm)
Temperature/K173(2)190(2)296(2)
a/pm592.55(8)653.16(2)656.90(3)
b/pm1699.7(2)905.98(3)906.59(4)
c/pm597.33(8)1242.98(3)909.32(4)
α90114.21(1)119.29(1)
β112.58(1)99.91(1)97.99(1)
γ9094.37(1)95.03(1)
V/106 pm3555.50652.00460.04
Formula units per cellZ = 2Z = 2Z = 1
Calculated density/(g·cm−3)3.103.953.83
Absorption coefficient/mm−16.511.811.4
Crystal size/mm³0.030 × 0.050 × 0.0700.010 × 0.020 × 0.0300.010 × 0.030 × 0.030
θ range/°2.4–32.52.4–30.02.6–32.6
Range in hkl−8 ≤ h ≤ 8

−25 ≤ k ≤ 25

−9 ≤ l ≤ 9
−9 ≤ h ≤ 9

−12 ≤ k ≤ 12

−17 ≤ l ≤ 17
−9 ≤ h ≤ 9

−13 ≤ k ≤ 13

−13 ≤ l ≤ 13
Total no. of reflections13,35423,1189659
Independent reflections1998 (Rint = 0.050)3802 (Rint = 0.051)3306 (Rint = 0.032)
Reflections with I≥2σ(I)1253 (Rσ = 0.031)3149 (Rσ = 0.037)2415 (Rσ = 0.050)
Data/restraints/parameters1998/2/943802/7/1673306/5/117
Goodness-of-fit on F21.0401.0350.945
R indices [I ≥ 2σ(I)]R1 = 0.030, wR2 = 0.054R1 = 0.027, wR2 = 0.043R1 = 0.032, wR2 = 0.051
R indices (all data)R1 = 0.066, wR2 = 0.062R1 = 0.041, wR2 = 0.045R1 = 0.058, wR2 = 0.056
Largest diff. peak and hole/(e·Å−3)0.73/−0.621.25/−1.271.46/−1.19
Table 2:

Atomic coordinates, occupancy, bond valence sums ∑ν (for oxygen atoms as O2–/OH) and equivalent isotropic displacement parameters (Ueq/10−4 pm2) of Mn5(OH)6Cl4. All atoms except Mn3 (2b) have the Wyckoff position 4e. Ueq is defined as one third of the trace of the orthogonalized Uij tensor (standard deviations in parentheses).

AtomOcc.νxyzUeq
Mn111.930000.0131(2)
Mn211.950.03843(8)0.90259(3)0.53171(8)0.0136(2)
Mn311.950.06984(7)0.80102(3)0.05025(6)0.0119(2)
Cl10.75−1.080.3443(2)0.91429(6)0.31052(18)0.0125(2)
Cl21−0.760.3728(2)0.80546(6)0.8240(2)0.0153(2)
Cl30.25−1.120.3048(5)0.0175(2)0.7795(5)0.0133(5)
O10.25−1.08/−1.950.207(2)0.9084(4)0.261(2)0.013(2)
O21−1.12/−1.990.1220(4)0.3915(2)0.7017(4)0.0162(4)
O31−1.13/−2.010.0894(3)0.2977(2)0.1811(3)0.0122(3)
O40.75−1.39/−2.270.1698(6)0.0116(2)0.7354(5)0.0129(6)
H210.271(3)0.384(2)0.744(6)0.015
H310.230(3)0.287(2)0.231(5)0.015
Table 3:

Anisotropic displacement parameters (Uij/10−4 pm2) of Mn5(OH)6Cl4 (standard deviations in parentheses).

AtomU11U22U33U23U13U12
Mn10.0195(3)0.0108(3)0.0112(3)−0.0001(2)0.0084(2)0.0014(2)
Mn20.0217(2)0.0101(2)0.0100(2)−0.0001(2)0.0072(2)−0.0005(2)
Mn30.0202(2)0.0088(2)0.0073(2)−0.0006(2)0.0060(2)−0.0010(2)
Cl10.0162(4)0.0116(4)0.0093(4)−0.0009(3)0.0046(3)−0.0032(4)
Cl20.0173(3)0.0165(3)0.0115(3)0.0002(3)0.0050(2)−0.0023(3)
Cl30.013(2)0.011(2)0.016(2)0.0010(9)0.005(2)−0.001(2)
O10.020(4)0.014(4)0.008(4)0.001(3)0.010(4)0.004(3)
O20.027(2)0.0113(9)0.014(2)−0.0009(8)0.011(2)−0.0018(8)
O30.0163(8)0.0114(7)0.0089(7)0.0003(9)0.0049(7)−0.0009(9)
O40.024(2)0.007(2)0.011(2)0.002(2)0.009(2)0.002(2)
Table 4:

Interatomic distances (/pm) in Mn5(OH)6Cl4, calculated with the single-crystal lattice parameters (standard deviations in parentheses).

Mn1–O2 (2×)217.0(2)Mn2–O3215.2(2)Mn3–O2214.6(2)
Mn1–O4 (2×)218.4(3)Mn2–O2215.3(2)Mn3–O3214.7(2)
Mn1–O1 (2×)221.1(8)Mn2–O4216.4(3)Mn3–O3215.2(2)
Mn1–Cl1 (2×)261.1(2)Mn2–O4218.9(3)Mn3–O1218.7(8)
Mn1–Cl3 (2×)262.9(3)Mn2–O1220.4(8)Mn3–Cl1261.3(2)
d ¯(Mn1-O)218.8Mn2–Cl3255.6(3)Mn3–Cl2 (2×)263.0(1)
d ¯(Mn1-Cl)262.0Mn2–Cl3258.9(3)d ¯(Mn3-O)215.8
Mn2–Cl1262.9(2)d ¯(Mn3-Cl)262.4
Mn2–Cl2265.4(1)
d ¯(Mn2-O)217.2d ¯(Mn-O)217.3
d ¯(Mn2-Cl)260.7¯(Mn-Cl)261.7
Table 5:

Atomic coordinates, bond valence sums (∑ν; for oxygen atoms: O2–/OH) and equivalent isotropic displacement parameters (Ueq/10−4 pm2) of Mn5(OH)7I3 and Mn7(OH)10I4. Ueq is defined as one third of the trace of the orthogonalized Uij tensor (standard deviations in parentheses).

AtomWyck.νxyzUeq
Mn5(OH)7I3
I12i−0.730.65765(4)0.70492(3)0.88703(2)0.01243(6)
I22i−0.670.62692(4)0.25245(3)0.45040(3)0.01412(6)
I32i−0.630.62111(4)0.80392(3)0.22178(2)0.01395(6)
Mn11a2.000000.0103(2)
Mn22i1.950.0426(2)0.58595(7)0.78827(5)0.0108(2)
Mn32i1.960.0082(2)0.19349(7)0.59568(5)0.0102(2)
Mn42i2.150.9695(2)0.67822(7)0.08979(5)0.0102(2)
Mn52i2.000.0579(2)0.90881(7)0.72416(5)0.0101(2)
Mn61g1.9001/21/20.0109(2)
O12i−1.14/−2.020.8897(5)0.3365(3)0.7481(2)0.0104(6)
O22i−1.18/−2.050.9053(4)0.9496(3)0.5738(2)0.0090(5)
O32i−1.12/−1.990.8363(5)0.5679(3)0.3661(2)0.0094(6)
O42i−1.12/−1.990.8292(4)0.1505(3)0.1281(2)0.0093(6)
O52i−1.07/−1.940.8466(4)0.4280(3)0.0415(2)0.0082(5)
O62i−1.12/−1.990.9050(4)0.0710(3)0.8557(2)0.0092(6)
O72i−1.13/−2.000.9006(5)0.6646(3)0.6511(2)0.0090(5)
H12i0.761(3)0.319(7)0.738(5)0.05(2)
H22i0.784(4)0.917(6)0.569(4)0.02(2)
H32i0.708(3)0.556(8)0.357(6)0.07(2)
H42i0.701(3)0.141(5)0.100(4)0.01(2)
H52i0.730(4)0.380(6)0.006(4)0.03(2)
H62i0.780(4)0.070(6)0.831(4)0.03(2)
H72i0.775(3)0.650(8)0.629(6)0.06(2)
Mn7(OH)10I4
I12i−0.600.36814(4)0.26630(3)0.08153(4)0.02300(7)
I22i−0.670.35498(4)0.74656(3)0.42638(4)0.02069(7)
Mn11a1.950000.0188(2)
Mn22i1.890.0282(2)0.43393(7)0.30934(8)0.0192(2)
Mn32i2.000.9843(2)0.87783(7)0.59164(7)0.0186(2)
Mn42i2.000.9946(2)0.29266(7)0.85544(8)0.0191(2)
O12i−1.11/−1.980.1519(5)0.9522(3)0.1939(4)0.0175(6)
O22i−1.10/−1.970.1471(5)0.3973(3)0.5169(4)0.0166(6)
O32i−1.15/−2.020.1214(5)0.8242(3)0.7906(4)0.0167(6)
O42i−1.10/−1.970.1327(5)0.1352(3)0.6470(4)0.0165(6)
O52i−1.13/−2.000.1059(5)0.5470(3)0.9057(3)0.0149(5)
H12i0.279(3)0.980(7)0.234(7)0.05(2)
H22i0.273(3)0.425(6)0.557(6)0.02(2)
H32i0.250(3)0.841(7)0.822(7)0.05(2)
H42i0.262(3)0.155(6)0.664(6)0.03(2)
H52i0.235(3)0.569(8)0.922(8)0.06(2)
Table 6:

Anisotropic displacement parameters (Uij/10−4 pm2) of Mn5(OH)7I3 and Mn7(OH)10I4 (standard deviations in parentheses).

AtomU11U22U33U23U13U12
Mn5(OH)7I3
I10.0114(2)0.0143(2)0.0125(2)0.0064(2)0.0032(2)0.0024(2)
I20.0114(2)0.0134(2)0.0174(2)0.0078(2)0.0005(2)0.0006(2)
I30.0137(2)0.0143(2)0.0141(2)0.0065(2)0.0029(2)0.0016(2)
Mn10.0180(5)0.0081(4)0.0075(4)0.0048(3)0.0047(3)0.0046(3)
Mn20.0225(3)0.0054(3)0.0061(3)0.0034(2)0.0044(3)0.0029(2)
Mn30.0192(3)0.0051(3)0.0067(3)0.0025(2)0.0041(2)0.0024(2)
Mn40.0192(3)0.0056(3)0.0078(3)0.0039(2)0.0051(2)0.0034(2)
Mn50.0185(3)0.0065(3)0.0067(3)0.0038(2)0.0034(2)0.0023(2)
Mn60.0209(5)0.0075(4)0.0052(4)0.0029(3)0.0039(3)0.0044(3)
O10.013(2)0.010(2)0.009(2)0.004(2)0.002(2)0.001(2)
O20.010(2)0.010(2)0.010(2)0.007(2)0.004(2)0.003(2)
O30.013(2)0.010(2)0.007(2)0.005(2)0.003(2)0.002(2)
O40.009(2)0.009(2)0.010(2)0.005(2)0.003(2)0.003(2)
O50.011(2)0.008(2)0.006(2)0.004(2)0.001(2)−0.000(2)
O60.012(2)0.008(2)0.008(2)0.005(2)0.001(2)0.002(2)
O70.011(2)0.009(2)0.009(2)0.005(2)0.002(2)0.001(2)
Mn7(OH)10I4
I10.0195(2)0.0216(2)0.0273(2)0.0131(2)0.0020(2)0.0016(2)
I20.0180(2)0.0202(2)0.0256(2)0.0125(2)0.0056(2)0.0040(2)
Mn10.0334(5)0.0138(4)0.0124(4)0.0081(3)0.0067(4)0.0068(4)
Mn20.0330(4)0.0138(3)0.0141(3)0.0092(2)0.0057(3)0.0054(3)
Mn30.0357(4)0.0116(3)0.0122(3)0.0080(2)0.0076(3)0.0064(3)
Mn40.0348(4)0.0100(3)0.0147(3)0.0066(2)0.0095(3)0.0051(3)
O10.021(2)0.017(2)0.018(2)0.012(2)0.005(2)0.003(2)
O20.018(2)0.015(2)0.019(2)0.010(2)0.004(2)0.003(2)
O30.021(2)0.015(2)0.016(2)0.008(2)0.004(2)0.005(2)
O40.018(2)0.017(2)0.015(2)0.009(2)0.004(2)0.003(2)
O50.017(2)0.016(2)0.015(2)0.010(2)0.004(2)0.003(2)
Table 7:

Selected interatomic distances (/pm) in Mn5(OH)7I3 and Mn7(OH)10I4, calculated with the single-crystal lattice parameters (standard deviations in parentheses).

Mn5(OH)7I3
Mn1–O6 (2×)215.4(3)Mn3–O1212.5(3)Mn5–O7211.2(3)
Mn1–O4 (2×)216.8(3)Mn3–O3214.2(3)Mn5–O2214.7(3)
Mn1–I1 (2×)300.83(3)Mn3–O2215.0(3)Mn5–O4214.7(3)
d¯ (Mn1–O)216.10Mn3–O2217.2(3)Mn5–O6215.5(3)
d¯ (Mn1–I)300.83Mn3–I3301.13(7)Mn5–I3297.85(7)
Mn3–I2303.22(7)Mn5–I2323.67(6)
Mn2–O5217.4(3)d¯ (Mn3–O)214.73d¯ (Mn5–O)214.03
Mn2–O3217.9(3)d¯ (Mn3–I)302.18d¯ (Mn5–I)310.76
Mn2–O7219.8(3)
Mn2–O4220.1(3)Mn4–O1212.7(3)Mn6–O7 (2×)210.8(3)
Mn2–O1221.7(3)Mn4–O6213.4(3)Mn6–O3 (2×)214.9(3)
Mn2–I1306.28(7)Mn4–O5214.1(3)Mn6–I2 (2×)297.75(3)
d¯ (Mn2–O)219.38Mn4–O5214.3(3)d¯ (Mn6–O)212.85
d¯ (Mn2–I)306.28Mn4–I3305.20(7)d¯ (Mn6–I)297.75
Mn4–I1306.07(7)
d¯ (Mn4–O)213.63d¯(Mn–O)215.12
d¯ (Mn4-I)305.64d¯(Mn–I)303.91
Mn7(OH)10I4
Mn1–O3 (2×)211.6(3)d¯ (Mn2-O)213.25Mn4–O4212.4(3)
Mn1–O1 (2×)215.1(3)d¯ Mn2-I)312.17Mn4–O1214.1(3)
Mn1–I1 (2×)299.20(3)Mn4–O5215.4(3)
d¯ (Mn1-O)213.35Mn3–O4214.2(3)Mn4–O5217.8(3)
d¯ (Mn1-I)299.20Mn3–O1218.3(3)Mn4–I2304.89(7)
Mn3–O3219.2(3)Mn4–I1308.00(7)
Mn2–O5212.3(3)Mn3–O2221.7(3)d¯ (Mn4-O)214.93
Mn2–O2212.9(3)Mn3–O4222.7(3)d¯ (Mn4-I)306.45
Mn2–O3213.0(3)Mn3–I2306.23(7)
Mn2–O2214.8(3)d¯ (Mn3-O)219.22d¯ (Mn–O)215.19
Mn2–I2302.00(7)d¯ (Mn3-I)306.23d¯ (Mn–I)306.01
Mn2–I1322.34(7)

Additional information on the crystal structure investigations can be obtained from the joint CCDC/FIZ Karlsruhe deposition service on quoting the deposition number 1940191 (Mn5(OH)6Cl4), 1940192 (Mn5(OH)7I3), and 1940193 (Mn7(OH)10I4).

2.3 Powder X-ray diffraction (PXRD)

A powdered sample containing the hydroxide halide Mn5(OH)6Cl4 was additionally characterized using powder X-ray diffraction (PXRD) technique. Therefore, data were collected with a Stoe Stadi P powder diffractometer in transmission geometry with Mo-Kα1 (λ = 70.930 pm; Ge(111) primary beam monochromator) radiation and a Mythen2 1K detector (Dectris, Baden, Switzerland).

2.4 Vibrational spectroscopy

To confirm the hydroxide groups, single-crystal transmission FTIR spectra were measured with a Bruker Vertex 70 FTIR spectrometer (resolution 4 cm−1), equipped with a KBr beam splitter, an LN-MCT (Mercury Cadmium Telluride) detector and a Hyperion 3000 microscope (Bruker, Vienna, Austria). The measurements were performed in absorption mode in the range of 600–4000 cm−1 using a silicon carbide rod (Globar) as mid-infrared source, and a 15×  IR objective as focus was used. Mn5(OH)6Cl4 was measured under a protective film of oil (Nujol for IR-spectroscopy, Alfa Aesar, Karlsruhe, Germany). About 260 scans of each single-crystal were acquired and a correction of atmospheric influences was performed with the software Opus 7.2 [38].

2.5 Bond-length bond-strength (BL/BS) concept

By applying the bond-length bond-strength concept [39], [40], the bond valence sums (∑ν) for all non-hydrogen atoms were calculated. The resulting values were compared to the formal ionic charges. The values for all oxygen atoms were calculated with and without the implication of a proton (O2–/OH) to affirm the presence of the hydroxide groups at each oxygen position.

3 Results and discussion

3.1 Crystal structures

Table 1 summarizes the results of the SCXRD analysis of the phases Mn5(OH)6Cl4, Mn5(OH)7I3, and Mn7(OH)10I4.

In contrast to γ-Mn(OH)Cl [29], which can be derived from the CdCl2 (C19) structure type, Mn5(OH)6Cl4 can be seen as a C6 structure with (AγB)(AγB) stacking sequence. The layers consist of edge-sharing distorted Mn(OH)6−xClx (x = 1, 2, 3, 4) octahedra forming corrugated, uncharged layers parallel to the (100) plane (Figure 1). This undulation prevents the formation of polytypes. The octahedral coordination around the manganese(II) atoms show substitutional disorder of the ligands Cl1/O1 (occupancy: 0.75/0.25) and Cl3/O4 (occupancy: 0.25/0.75).

Figure 1: The crystal structure of Mn5(OH)6Cl4 viewed along the a- and c-axis and coordination of the manganese atoms (displacement ellipsoids represent 90% probability at 173 K). The occupancies of the disordered atom pairs O1/Cl1 and O4/Cl3 are 0.25/0.75 and 0.75/0.25, respectively.
Figure 1:

The crystal structure of Mn5(OH)6Cl4 viewed along the a- and c-axis and coordination of the manganese atoms (displacement ellipsoids represent 90% probability at 173 K). The occupancies of the disordered atom pairs O1/Cl1 and O4/Cl3 are 0.25/0.75 and 0.75/0.25, respectively.

The Mn–Cl distances (Table 4) range from 255.6(3) to 265.4(1) pm with average values of 262.0 (Mn1–Cl), 260.7 (Mn2–Cl), and 262.4 pm (Mn3–Cl). These bonds are longer than in MnCl2 (254.8(2) pm [41]) and similar to the ones in γ-Mn(OH)Cl (average 260.9 pm [29]) The oxygen atoms of the hydroxide groups are situated at Mn–O distances ranging from 214.6(2) to 221.1(8) pm and in average (Mn1–O: 218.8, Mn2–O: 217.2, Mn3–O: 215.8 pm) shorter than the mean Mn–O distance in Mn(OH)2 (219.62 pm [42]). This behavior indicates that the hydroxide groups in Mn5(OH)6Cl4 are stronger ligands to manganese(II) than the chloride ions and can be attributed to hydrogen bonding. The Cl–Mn and O–H bonds are weakened as part of the valence is used for Cl binding to H. Correspondingly, the O–Mn bond strength increases with weaker O–H bonds.

The bond valence sums (∑ν) for all non-hydrogen atoms were calculated (Table 2), whereby oxygen atoms were treated without and with the implication of a proton (O2–/OH). The calculated values for the oxygen atoms are −1.08/−1.95 (O1), −1.12/−1.99 (O2), −1.13/−2.01 (O3), and −1.39/−2.27 (O4), confirming the presence of hydroxide groups. For the non-oxygen atoms the values of 1.93 (Mn1), 1.95 (Mn2, Mn3), −1.08 (Cl1), −0.76 (Cl2), and −1.12 (Cl3) correspond well to the formal ionic charges of the atoms.

The manganese hydroxide iodides Mn5(OH)7I3 and Mn7(OH)10I4 both crystallize in the triclinic space group P1 (no. 2) and similar to Mn(OH)I [30] and Mn5(OH)6Cl4 described above, their crystal structure is related to the brucite (C6, Mg(OH)2/CdI2) structure. The single layers are formed by distorted Mn(OH)6−xIx (x = 1, 2) octahedra parallel to the (100)-plane (Figures 2 and 3). Within the octahedra having two iodine atoms, the halogen atoms occur both in cis and trans position. In contrast to Mn5(OH)6Cl4, the iodides show no substitutional disorder of the ligands.

Figure 2: Octahedral coordination spheres of the manganese(II) atoms in Mn5(OH)7I3 (displacement ellipsoids represent 90% probability at 190 K) and polyhedral representation of the crystal structure viewed along the a- and b-axis.
Figure 2:

Octahedral coordination spheres of the manganese(II) atoms in Mn5(OH)7I3 (displacement ellipsoids represent 90% probability at 190 K) and polyhedral representation of the crystal structure viewed along the a- and b-axis.

Figure 3: Octahedral coordination of the manganese(II) atoms in Mn7(OH)10I4 (displacement ellipsoids represent 90% probability at 296 K) and polyhedral representation of the crystal structure viewed along the a- and b-axis.
Figure 3:

Octahedral coordination of the manganese(II) atoms in Mn7(OH)10I4 (displacement ellipsoids represent 90% probability at 296 K) and polyhedral representation of the crystal structure viewed along the a- and b-axis.

The Mn–I distances (Table 7) range from 297.75(3) to 323.67(6) pm (average 303.91 pm) in Mn5(OH)7I3, and 299.20(3) to 322.34(7) pm (average 306.01 pm) in Mn7(OH)10I4. These bonds are considerably longer than in MnI2 (292.02 pm [43]) and in MnI2·4H2O (292.16 pm [44]).

2The values for the Mn–O bond-lengths are between 210.8(3) and 221.7(3) pm (average 215.2 pm) for Mn5(OH)7I3, and 211.6(3) and 222.7(3) pm (average 215.1 pm) in Mn7(OH)10I4 and are shorter than the average Mn–O distances in Mn(OH)2 (219.62 pm [42]). Like in Mn5(OH)6Cl4, this behavior can be seen as a result of the existing hydrogen bonds.

Also for the iodides, bond valence sums (∑ν) for all non-hydrogen atoms were determined (Table 5). The non-oxygen atoms have values close to their formal ionic charges. The values for the oxygen atoms calculated as O2– and OH are −1.14/−2.02 (O1), −1.18/−2.05 (O2), −1.12/−1.99 (O3, O4, O6), −1.07/−1.94 (O5), and −1.13/−2.00 (O7) for Mn5(OH)7I3 and −1.11/−1.98 (O1), −1.10/−1.97 (O2, O4), −1.15/−2.02 (O3), and −1.13/−2.00 (O5) for Mn7(OH)10I4. These values confirm the presence of hydroxide groups. The rather small bond-valence sums of the iodine atoms are presumably a result of the influence of the O–H··· X hydrogen bonds, which were not considered in the BLBS calculations. The IR spectra (vide infra) show the presence of stronger hydrogen bonding in Mn5(OH)7I3 and Mn7(OH)10I4 than in Mn5(OH)6Cl4, thus having more influence on the I–Mn bond valence and leading to the underestimation of the valence sum.

3.2 Powder X-ray diffraction of Mn5(OH)6Cl4

A sample containing the chloride phase was additionally analyzed by PXRD. The measured diffraction pattern was compared to the theoretical pattern derived from SCXRD data (Figure 4). Besides Mn5(OH)6Cl4, which represented the main phase, the sample contained an unidentified side phase.

Figure 4: The measured diffraction pattern of a sample containing Mn5(OH)6Cl4 (top) compared to the simulated pattern from SCXRD data (bottom). Reflections marked with an asterisk originate from an unidentified side phase.
Figure 4:

The measured diffraction pattern of a sample containing Mn5(OH)6Cl4 (top) compared to the simulated pattern from SCXRD data (bottom). Reflections marked with an asterisk originate from an unidentified side phase.

3.3 Group–subgroup relationships

The manganese hydroxide halides Mn5(OH)6Cl4, Mn5(OH)7I3, and Mn7(OH)10I4 reported herein are crystal chemically closely related to the well-known layer structure of cadmium iodide or brucite (Mg(OH)2). CdI2 crystallizes with space group P3¯2/m1, which is a maximal subgroup of P63/mmc, the space group type of the hexagonal closest packing. In CdI2, every other layer of octahedral voids of the hcp iodine packing remains empty. In the present case, we have distorted hexagonal packings of the hydroxide and halide ions and the octahedral voids are filled by Mn2+.

These close relationships readily call for group–subgroup schemes, relating the three new structures with the aristotype CdI2. The first Bärnighausen tree (Figure 5) [45], [46], [47], [48] of the CdI2 family was worked out by Meyer during his PhD work [49] in the Bärnighausen group. General considerations on subgroups for hcp packings with partial ordered occupancies of octahedral voids were published by Müller [48], [50], [51].

Figure 5: Group–subgroup scheme for CdI2 superstructures. The indices for the translationengleiche (t), klassengleiche (k), and isomorphic (i) symmetry reductions are given. The highlighted part of the scheme was reproduced from [49].
Figure 5:

Group–subgroup scheme for CdI2 superstructures. The indices for the translationengleiche (t), klassengleiche (k), and isomorphic (i) symmetry reductions are given. The highlighted part of the scheme was reproduced from [49].

The Bärnighausen tree worked out by Meyer [49] has three different branches: (i) the structure of Na2Sn(OH)6, (ii) the structures of Li2Pt(OH)6, Bi2Pt (h2), Na2Pt(OD)6 and Tl2S, and (iii) the structure of CrBr2 and its lower symmetric derivatives NbTe2, AgAuCl4, AgAuTe4, and Cu2Cl(OH)3, nicely underpinning the broad crystal chemical diversity of the CdI2 family. The third part of the Bärnighausen tree from Meyer is included in Figure 5, since the superstructures of the manganese hydroxide halides Mn5(OH)6Cl4, Mn5(OH)7I3, and Mn7(OH)10I4 also derive from CrBr2 [52]. The structure of chromium dibromide is monoclinic. The space group symmetry is reduced by a translationengleiche symmetry reduction of index 3 (t3) from P3¯2/m1 to C12/m1.

Each, Mn5(OH)6Cl4, Mn5(OH)7I3, and Mn7(OH)10I4 deserve two subsequent steps starting from CrBr2. This is paralleled with unit cell enlargements. In Figure 6, we present a projection of the MnI2 structure along the trigonal axis. The unit cells of the three hydroxide halide are incorporated into this drawing for better comparison with the aristotype. The individual Bärnighausen trees along with the evolution of the atomic parameters are shown in Figures 7–9.

Figure 6: Structure of MnI2 [43] with view along the c-axis. The manganese and iodine atoms are drawn as gray and purple circles, respectively. The cell of MnI2 is highlighted (gray) as well as the cells of the CdI2 superstructures Mn5(OH)6Cl4 (blue), Mn7(OH)10I4 (green), and Mn5(OH)7I3 (red).
Figure 6:

Structure of MnI2 [43] with view along the c-axis. The manganese and iodine atoms are drawn as gray and purple circles, respectively. The cell of MnI2 is highlighted (gray) as well as the cells of the CdI2 superstructures Mn5(OH)6Cl4 (blue), Mn7(OH)10I4 (green), and Mn5(OH)7I3 (red).

Figure 7: Group–subgroup scheme in the Bärnighausen formalism [45], [46], [47], [48] for the structures of Mn(OH)2 [42], HP-CrBr2 [52], and Mn5(OH)6Cl4. The indices for the translationengleiche (t), klassengleiche (k), and isomorphic (i) symmetry reductions and the evolution of the atomic parameters are given.
Figure 7:

Group–subgroup scheme in the Bärnighausen formalism [45], [46], [47], [48] for the structures of Mn(OH)2 [42], HP-CrBr2 [52], and Mn5(OH)6Cl4. The indices for the translationengleiche (t), klassengleiche (k), and isomorphic (i) symmetry reductions and the evolution of the atomic parameters are given.

Figure 8: Group–subgroup scheme in the Bärnighausen formalism [45], [46], [47], [48] for the structures of MnI2 [43], HP-CrBr2 [52] and Mn7(OH)10I4. The indices for the translationengleiche (t), klassengleiche (k), and isomorphic (i) symmetry reductions and the evolution of the atomic parameters are given.
Figure 8:

Group–subgroup scheme in the Bärnighausen formalism [45], [46], [47], [48] for the structures of MnI2 [43], HP-CrBr2 [52] and Mn7(OH)10I4. The indices for the translationengleiche (t), klassengleiche (k), and isomorphic (i) symmetry reductions and the evolution of the atomic parameters are given.

Figure 9: Group–subgroup scheme in the Bärnighausen formalism [45], [46], [47], [48] for the structures of MnI2 [43], HP-CrBr2 [52] and Mn5(OH)7I3. The indices for the translationengleiche (t), klassengleiche (k), and isomorphic (i) symmetry reductions and the evolution of the atomic parameters are given.
Figure 9:

Group–subgroup scheme in the Bärnighausen formalism [45], [46], [47], [48] for the structures of MnI2 [43], HP-CrBr2 [52] and Mn5(OH)7I3. The indices for the translationengleiche (t), klassengleiche (k), and isomorphic (i) symmetry reductions and the evolution of the atomic parameters are given.

We start with the structure of Mn5(OH)6Cl4 (Figure 7). The subsequent i5 and k2 transitions lead to five crystallographically independent anion sites, which allow the hydroxide-halide ordering. Nevertheless, the structure shows some residual disorder. Especially these sites show the largest differences between the positional parameters calculated from the subcell and those refined from the X-ray data. The manganese cations also show small displacements from the ideal positions in order to account for the anion substructure (OH/Cl ordering).

For Mn7(OH)10I4, the second step in symmetry reduction (Figure 8) is an isomorphic one from P1¯ to P1;¯ with the rare index seven (i7). Consequently, we obtain seven anion sites for the OH/I ordering. The latter was well resolved from the single-crystal X-ray diffraction data. Again, the anions show the largest displacements from the ideal subcell sites. This is expressed in the undulated layers of condensed octahedra, drastically deviating from planarity as in the aristotype.

Mn5(OH)7I3 shows the most complicated of the three superstructures with 10 crystallographically independent anion sites, similar to Mn7(OH)10I4 with complete OH/I ordering. The displacements of the different anions from the ideal subcell positions are similar to the two compounds described above and also results in undulated layers of condensed octahedra.

The strong undulations of the Mn(OH/X)2 layers for all three compounds show the limits of group–subgroup relations. The strong shifts in the atomic parameters underpin the differences in the radii OH/Cl versus OH/I but also changes in chemical bonding, i.e., ionicity toward covalence. In that view, it is interesting to mention the intermetallic phase Bi2Pt [53]. The latter also derives from the CdI2 type and is included in the Bärnighausen tree in Figure 5. The evolution of the atomic parameters presented in [53] shows substantial displacements for both the platinum and bismuth atoms, leading to distorted PtBi6/3 octahedra. Also, this example is a limit of a group–subgroup scheme. Besides the purely geometrical relationship, the chemical bonding pattern should match as close as possible, in order to have a close crystal chemical similarity.

3.4 FTIR-spectroscopy

Figures 10–12 show the experimental IR spectra of single-crystals of Mn5(OH)6Cl4, Mn5(OH)7I3, and Mn7(OH)10I4 in the range from 600 to 4000 cm−1 in absorbance mode. In the fingerprint region, the strong bands from 600 to 800 cm−1 correspond to ν(Mn–O) stretching modes. In the range from 800 to 1200 cm−1, O–H bending modes of the hydroxide groups can be found (very weak in Mn5(OH)6Cl4, strong in the iodide phases). Marked peaks in Figure 10 are due to the protective Nujol film on the Mn5(OH)6Cl4 crystal and the bands in the region from 1500 to 3000 cm−1 in the spectra of the iodide compounds (Figures 11 and 12) originate from residual grease of the sample preparation procedure. In the spectrum of Mn5(OH)7I3 (Figure 11), the band from 3000 to 3300 cm−1 can be attributed to absorbed water and the small shoulder at around 1600 cm−1 to the corresponding bending vibration of O–H groups in the absorbed water molecules.

Figure 10: FTIR spectrum of a single-crystal of Mn5(OH)6Cl4 under a protective layer of Nujol (bands marked with an asterisk).
Figure 10:

FTIR spectrum of a single-crystal of Mn5(OH)6Cl4 under a protective layer of Nujol (bands marked with an asterisk).

Figure 11: FTIR spectrum of a single-crystal of Mn5(OH)7I3.
Figure 11:

FTIR spectrum of a single-crystal of Mn5(OH)7I3.

Figure 12: FTIR spectrum of a single-crystal of Mn7(OH)10I4.
Figure 12:

FTIR spectrum of a single-crystal of Mn7(OH)10I4.

All spectra show sharp ν(O–H) stretching vibrations at around 3500 cm−1 (Mn5(OH)6Cl4: 3550 cm−1; Mn5(OH)7I3: 3491, 3533, and 3573 cm−1; Mn7(OH)10I4: 3503, 3552, and 3595 cm−1) and confirm the presence of the hydroxide groups. The red-shift of the values for these modes compared to the value for a free hydroxide group (3620 cm−1 [54]) is an indication for the occurrence of hydrogen bonding of the type O–H···X (X = Cl, I) between the layers. The occurring peak splitting in the IR spectra of the iodide phases (Figures 11 and 12) can be interpreted as a result of existing hydrogen bonds of different strengths arising from different O–H···I distances.

4 Conclusions

Three new manganese(II) hydroxide halides were obtained by means of high-pressure/high-temperature synthesis using a multianvil apparatus. Single-crystals of the compounds Mn5(OH)6Cl4, Mn5(OH)7I3, and Mn7(OH)10I4 were characterized via X-ray diffraction, showing layered crystal structures derived from the CdI2/Mg(OH)2 crystal structure. The chloride phase crystallizes in the space group P21/c and exhibits a substitutional disorder concerning half of the ligand positions, whereas the iodide phases both crystallize in P1¯ and do not exhibit any disorder. In contrast to the aristotype, these hydroxide halide phases show a distortion of the coordination polyhedra and strong undulation of the layers that prevents the formation of polytypes. Additional FTIR spectroscopy affirm the hydroxide groups of the compounds and indicate the occurrence of hydrogen bonding of type O–H···X.


Dedicated to Professor Dr. Ulrich Müller on the occasion of his 80th birthday.



Corresponding author: Hubert Huppertz, Institut für Allgemeine, Anorganische und Theoretische Chemie, Universität Innsbruck, Innrain 80–82, 6020Innsbruck, Austria, E-mail:

Funding source: Universität Innsbruck

Funding source: Austrian Science Fund

Award Identifier / Grant number: I 2179

Funding source: German Research Council (DFG)

Award Identifier / Grant number: RU 776/15‐1

Acknowledgment

The authors would like to thank Assoz.-Prof. Dr. G. Heymann for the collection of single-crystal data. Special thanks go to Prof. Dr. R. Stalder (Institute for Mineralogy and Petrography, University of Innsbruck) for granting us access to the FTIR-microscope and Dr. D. Vitzthum for the IR measurements.

  1. Author contribution: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.

  2. Research funding: This work was financially supported by the Austrian Science Fund (FWF, Project no. I 2179) and the German Research Council (DFG, Project no. RU 776/15‐1).

  3. Conflict of interest statement: The authors declare no conflicts of interest regarding this article.

References

1. Venetopoulos, C. H., Rentzeperis, J. P. Z. Kristallogr. 1975, 141, 246–259. https://doi.org/10.1524/zkri.1975.141.3-4.246.Search in Google Scholar

2. Schlüter, J., Klaska, K.-H., Gebhard, G. Neues Jahrb. Mineral. Monatsh. 2000, 67–73.Search in Google Scholar

3. Krivovichev, S. V., Hawthorne, F. C., Williams, P. A. Struct. Chem. 2017, 28, 153–159. https://doi.org/10.1007/s11224-016-0792-z.Search in Google Scholar

4. Oswald, H. R., Feitknecht, W. Helv. Chim. Acta 1964, 47, 272–289. https://doi.org/10.1002/hlca.19640470136.Search in Google Scholar

5. Feitknecht, W., Held, F. Helv. Chim. Acta 1944, 27, 1480–1495. https://doi.org/10.1002/hlca.194402701189.Search in Google Scholar

6. Feitknecht, W., Maget, K. Helv. Chim. Acta 1949, 32, 1639–1653. https://doi.org/10.1002/hlca.19490320535.Search in Google Scholar

7. Oswald, H. R., Feitknecht, W. Helv. Chim. Acta 1961, 44, 847–858. https://doi.org/10.1002/hlca.19610440329.Search in Google Scholar

8. Feitknecht, W., Oswald, H. R., Forsberg, H. E. Chimia (Aarau) 1959, 13, 113.Search in Google Scholar

9. Peter, S., Cockcroft, J. K., Roisnel, T., Lutz, H. D. Acta Crystallogr. 1996, 52B, 423–427. https://doi.org/10.1107/s0108768195013449.Search in Google Scholar

10. Cudennec, Y., Riou, A., Gérault, Y., Lecerf, A. J. Solid State Chem. 2000, 151, 308–312. https://doi.org/10.1006/jssc.2000.8659.Search in Google Scholar

11. Jung, C., Lutz, H. D. J. Mol. Struct. 1996, 381, 15–22. https://doi.org/10.1016/0022-2860(96)09229-0.Search in Google Scholar

12. Weckler, B., Lutz, H. D. Spectrochim. Acta 1996, A52, 1507–1513. https://doi.org/10.1016/0584-8539(96)01693-5.Search in Google Scholar

13. Ludi, A., Feitknecht, W. Helv. Chim. Acta 1963, 46, 2226–2238. https://doi.org/10.1002/hlca.19630460641.Search in Google Scholar

14. Ludi, A., Feitknecht, W. Helv. Chim. Acta 1963, 46, 2238–2248. https://doi.org/10.1002/hlca.19630460642.Search in Google Scholar

15. Lutz, H. D., Beckenkamp, K., Peter, S. Spectrochim. Acta 1995, A51, 755–767. https://doi.org/10.1016/0584-8539(94)01292-o.Search in Google Scholar

16. Liu, X. D., Zheng, X. G., Meng, D. D., Xu, X. L. Appl. Mech. Mater. 2013, 341-342, 242–245. https://doi.org/10.4028/www.scientific.net/amm.341-342.242.Search in Google Scholar

17. Liu, X.-D., Meng, D.-D., Hagihala, M., Zheng, X.-G., Guo, Q.-X. Chin. Phys. B 2011, 20, 077801. https://doi.org/10.1088/1674-1056/20/7/077801.Search in Google Scholar

18. Danilovich, I. L., Volkova, O. S., Vasiliev, A. N. Low Temp. Phys. 2017, 43, 529–542. https://doi.org/10.1063/1.4985210.Search in Google Scholar

19. Liu, X.-D., Hagihala, M., Zheng, X.-G., Guo, Q.-X. Vib. Spectrosc. 2011, 56, 177–183. https://doi.org/10.1016/j.vibspec.2011.02.002.Search in Google Scholar

20. Zheng, X. G., Yamashita, T., Hagihala, M., Fujihala, M., Kawae, T. Phys. B 2009, 404, 680–682. https://doi.org/10.1016/j.physb.2008.11.221.Search in Google Scholar

21. Liu, X.-D., Hagihala, M., Zheng, X.-G., Tao, W.-J., Meng, D.-D., Zhang, S.-L., Guo, Q.-X. Chin. Phys. Lett. 2011, 28, 017803. https://doi.org/10.1088/0256-307x/28/1/017803.Search in Google Scholar

22. Liu, X.-D., Hagihala, M., Zheng, X.-G., Meng, D.-D., Guo, Q.-X. Chin. Phys. Lett. 2011, 28, 087805. https://doi.org/10.1088/0256-307x/28/8/087805.Search in Google Scholar

23. Fujihala, M., Hagihala, M., Zheng, X. G., Kawae, T. Phys. Rev. 2010, B82, 024425. https://doi.org/10.1103/PhysRevB.82.024425.Search in Google Scholar

24. Zheng, X.-G., Nishiyama, K. Phys. B 2006, 374–375, 156–159. https://doi.org/10.1016/j.physb.2005.11.041.Search in Google Scholar

25. Zhang, Z., Yin, L. J. Mater. Chem. 2015, A3, 17659–17668. https://doi.org/10.1039/c5ta03426d.Search in Google Scholar

26. Ranganatha, S., Kumar, S., Penki, T. R., Kishore, B., Munichandraiah, N. J. Solid State Electrochem. 2017, 21, 133–143. https://doi.org/10.1007/s10008-016-3348-7.Search in Google Scholar

27. Ma, W., Feng, Y., Wang, L., Li, Y., Shi, M., Cui, H. Adv. Powder Technol. 2017, 28, 2642–2647. https://doi.org/10.1016/j.apt.2017.07.016.Search in Google Scholar

28. Masato, H., Zheng, X. G., Takato, T., Kawae, T. J. Phys. Condens. Matter. 2007, 19, 145281. https://doi.org/10.1088/0953-8984/19/14/145281.Search in Google Scholar

29. Falkowski, V., Zeugner, A., Isaeva, A., Ruck, M., Huppertz, H. Eur. J. Inorg. Chem. 2018, 42, 4630–4637. https://doi.org/10.1002/ejic.201800928.Search in Google Scholar PubMed PubMed Central

30. Falkowski, V., Zeugner, A., Isaeva, A., Wurst, K., Ruck, M., Huppertz, H. Z. Anorg. Allg. Chem. 2019, 645, 919–926. https://doi.org/10.1002/zaac.201900103.Search in Google Scholar

31. Sillén, L. G. Sven. Kem. Tidskr. 1941, 53, 39–43.10.1177/0040571X4104325306Search in Google Scholar

32. Huppertz, H. Z. Kristallogr. 2004, 219, 330–338. https://doi.org/10.1524/zkri.219.6.330.34633.Search in Google Scholar

33. Sadabs-2014/5; Madison, WI: Bruker AXS, 2015.Search in Google Scholar

34. Krause, L., Herbst-Irmer, R., Sheldrick, G. M., Stalke, D. J. Appl. Crystallogr. 2015, 48, 3–10. https://doi.org/10.1107/s1600576714022985.Search in Google Scholar PubMed PubMed Central

35. Parthé, E., Gelato, L. M. Acta Crystallogr. 1984, 40A, 169–183. https://doi.org/10.1107/s0108767384000416.Search in Google Scholar

36. Gelato, L. M., Parthé, E. J. Appl. Crystallogr. 1987, 20, 139–143. https://doi.org/10.1107/s0021889887086965.Search in Google Scholar

37. Hu, S.-Z., Parthé, E. Chin. J. Struct. Chem. 2004, 23, 1150–1160.Search in Google Scholar

38. Opus, v. 7.2. Billerica, MA: Bruker Optik GmbH, 2012.Search in Google Scholar

39. Brese, N. E., O’Keeffe, M. Acta Crystallogr. 1991, 47B, 192–197. https://doi.org/10.1107/s0108768190011041.Search in Google Scholar

40. Brown, I. D., Altermatt, D. Acta Crystallogr. 1985, 41B, 244–247. https://doi.org/10.1107/s0108768185002063.Search in Google Scholar

41. Tornero, J. D. J. Fayos Z. Kristallogr. 1990, 192, 147–148. https://doi.org/10.1524/zkri.1990.192.14.147.Search in Google Scholar

42. Christensen, A. N., Ollivier, G. Solid State Commun. 1972, 10, 609–614. https://doi.org/10.1016/0038-1098(72)90602-3.Search in Google Scholar

43. Cable, J. W., Wilkinson, M. K., Wollan, E. O., Koehler, W. C. Phys. Rev. 1962, 125, 1860–1864. https://doi.org/10.1103/physrev.125.1860.Search in Google Scholar

44. Moore, J. E., Abola, J. E., Butera, R. A. Acta Crystallogr. 1985, 41C, 1284. https://doi.org/10.1107/s0108270185007466.Search in Google Scholar

45. Bärnighausen, H. Commun. Math. Chem. 1980, 9, 139.Search in Google Scholar

46. Müller, U. Z. Anorg. Allg. Chem. 2004, 630, 1519. https://doi.org/10.1002/zaac.200400250.Search in Google Scholar

47. Müller, U. Relating crystal structures by group-subgroup relations. In International Tables for Crystallography, Vol. A1, Symmetry Relations between Space Groups, 2nd ed.; Wondratschek, H., Müller, U., Eds. Chichester: John Wiley Sons, 2010; pp. 44–56.10.1107/97809553602060000795Search in Google Scholar

48. Müller, U. Symmetry relationships between crystal structures; Oxford: Oxford Universiry Press, 2013; Relaciones de simetría entre estructuras cristalinas; Madrid: Editorial Síntesis, 2013.10.1093/acprof:oso/9780199669950.001.0001Search in Google Scholar

49. Meyer, A. Symmetriebeziehungen zwischen Kristallstrukturen des Formeltyps AX2, ABX4 und AB2X6 sowie deren Ordnungs- und Leerstellenvarianten. Dissertation. Germany: Universität Karlsruhe (TH), 1981.Search in Google Scholar

50. Müller, U. Acta Crystallogr. 1992, B48, 972.10.1107/S010876819101340XSearch in Google Scholar

51. Müller, U. Z. Anorg. Allg. Chem. 1998, 624, 529–532.10.1002/(SICI)1521-3749(199803)624:3<529::AID-ZAAC529>3.0.CO;2-RSearch in Google Scholar

52. Tracy, J. W., Gregory, N. W., Lingafelter, E. C. Acta Crystallogr. 1962, 15, 672. https://doi.org/10.1107/s0365110x6200184x.Search in Google Scholar

53. Kaiser, M., Baranov, A. I., Ruck, M. Z. Anorg. Allg. Chem. 2014, 640, 2742–2746. https://doi.org/10.1002/zaac.201400331.Search in Google Scholar

54. Kagunya, W., Baddour-Hadjean, R., Kooli, F., Jones, W. Chem. Phys. 1998, 236, 225–234. https://doi.org/10.1016/s0301-0104(98)00234-1.Search in Google Scholar

Received: 2020-04-15
Accepted: 2020-07-07
Published Online: 2020-08-25
Published in Print: 2020-09-25

© 2020 Viktoria Falkowski et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 11.6.2024 from https://www.degruyter.com/document/doi/10.1515/zkri-2020-0040/html
Scroll to top button