Abstract
Metallic black platelet single crystals of a new ternary compound, Sr7N2Sn3, were obtained by heating Sr and Sn in a Na flux together with NaN3 as a nitrogen source at 1073 K, followed by slow cooling. Single-crystal X-ray analysis revealed that this compound crystallizes in an orthorhombic cell with the cell parameters a = 10.4082(2), b = 18.0737(4), and c = 7.43390(10) Å (space group Pmna, Z = 2), and has a layered (modular) antiperovskite-type structure which could be related to the inverse structure of Ca2Nb2O7 ((Ca2)[Ca2Nb4O14]). Four-membered zigzag [Sn4] chains are situated between slabs comprising four antiperovskite layers cut by the (110) plane of the ideal anitiperovskite structure, and Sr7N2Sn3 can be expressed as [Sn4][Sn2N4Sr14]. Although an electron-precise valence electron distribution according to the formula (Sr2+)14(N3−)4(Sn4−)2([Sn4]8−) is proposed for this ternary compound, yet, there are certain structural peculiarities which cannot be explained by this idealized picture. Therefore, first principles-based means were employed to account for the aforementioned structural features.
1 Introduction
A wide variety of oxides having the perovskite-type structure (ABC
3), or related structures with components containing blocks of perovskite structure units, have been synthesized, and the properties and applications of these materials have been examined [1]. Recently, compounds having the perovskite-type structure in which the cation and anion sites are exchanged have been synthesized and characterized [2, 3]. As an example, there are 118 records in the inorganic crystal structure data base of ideal antiperovskite compounds with the space group Pm
In contrast, there have been few reports on layered antiperovskite compounds. (PrCa3N)Bi2 has been shown to have the antistructure to that of K2NiF4, which is a Ruddlesden–Popper phase [5]. The crystal structure of Sr11Ge4N6 contains layers of the nitridogermanate group [GeN2]4− and a distorted antiprism layer of [Sr8/2Ge]4+ sandwiched between slabs of double antiperovskite layers cut in the perovskite [100] direction [6]. Some subnitrides and nitride-related compounds containing N3− at the centers of Sr/Ba octahedra as also occurs in antiperovskite subnitrides, and incorporating nitridometallate and Zintl polyanions, have been prepared by the Na flux method [7]. The present paper reports the preparation, crystal structure, and electronic structure of a new strontium nitride stannide having a layered antiperovskite structure.
2 Experimental
All experimental manipulations described herein were carried out in an Ar-filled glove box (MBRAUN, O2, H2O < 1 ppm). Sn powder (99.9%, grain size ≤45 μm, Fujifilm Wako Pure Chemical Co.), Sr (99.99%, Aldrich), NaN3 (99.9%, Tokyo Chemical Industry Co., Ltd.), and Na (99.95%, Nippon Soda Co. Ltd.) were used to synthesize compounds in the Sr–Sn–N system. In these experiments, 0.046 g of Sn, 0.068 g of Sr, 0.030 g of NaN3, and 0.036 g of Na (for a Sn:Sr:NaN3:Na molar ratio of 0.5:1.0:0.6:2.0, Sample I) or 0.033 g of Sn, 0.099 g of Sr, 0.022 g of NaN3, and 0.066 g of Na (Sn:Sr:NaN3:Na molar ratio of 0.5:2.0:0.6:2.0, Sample II) were placed in boron nitride crucibles (99.5%; inner diameter, 6.5 mm; depth, 18 mm; Showa Denko K.K.), that were then sealed in a stainless steel (SUS 316) container (inner diameter, 10.7 mm; depth, 80 mm). The container was subsequently heated to T = 1073 K for 4 h, held at this temperature for 2 h, and then cooled to 573 K for 96 h. Following this, the heater power was turned off and the sample was allowed to cool naturally in the furnace. The cooled container was opened in the glove box and the specimen was transferred to a glass container and then heated under a reduced pressure of approximately 0.1 Pa at 573 K for 12 h to evaporate residual Na. Single crystals were extracted from the samples while observing the specimens using an optical microscope attached to the glove box.
The chemical composition of each single crystal was analyzed using electron probe microanalysis (EPMA) and wavelength dispersive X-ray spectrometry (WDX) (A-8200, JEOL). The single crystals were fixed in glass capillaries using grease after which the capillaries were sealed on a hot Pt filament. Single crystal X-ray diffraction (XRD) data were collected using a single-crystal diffractometer (D8 QUEST, Bruker AXS) using MoKα radiation and employing the Apex3 software package [8]. Numerical and multi-scan absorption corrections were performed using the program Sadabs [9]. Refinement of the atomic coordinates and displacement parameters was performed with the Shelxl-2014 program [10]. The Vesta software [11] was employed to visualize the crystal structures.
CCDC 095786 (Sr7N2Sn3), 2095787 (SrSn) and 2095839 (Sr3N0.77Sn) contain the supplementary crystallographic data for this paper. These data can be obtained free of charge from The joint ICSD and CCDC database via www.ccdc.cam.ac.uk/data_request/cif.
To provide an insight into the electronic structure of the ternary nitride stannide, density-functional-based methods were employed by means of the projector augmented wave (PAW) method [12] as implemented in the Vienna ab initio simulation package (VASP) [13], [14], [15], [16], [17]. Prior to the bonding analyses, full structural optimizations which included both lattice parameters as well as atomic positions were accomplished utilizing the aforementioned approach. Correlation and exchange in all computations were described by the generalized gradient approximation of Perdew, Burke, and Ernzerhof (GGA-PBE) [18], while the energy cut-off of the plane wave basis set was 500 eV. A 7 × 4 × 10 k-points set was used to sample the first Brillouin zone, and all computations were expected to be converged as the energy difference between two iterative steps fell below 10−8 (and 10−6) eV per cell of the electronic (and ionic) relaxations.
A chemical bonding analysis was completed based on an examination of the Mulliken and Löwdin charges [19] as well as the projected crystal orbital Hamilton populations (pCOHP) [20] for Sr7N2Sn3. The Mulliken and Löwdin charges were obtained by subtracting the gross population of a given atom from its number of valence electrons, while the Hamilton populations [21, 22] are determined by weighting the off-site entries of the densities-of-states with the respective Hamilton matrix elements. The constructions of the pCOHPs and the computations of the Mulliken and Löwdin charges require the use of local basis sets, whose nature is in stark contrast to that of the delocalized one of the plane-wave basis sets. Therefore, the results of the plane-wave-based computations had to be projected into a local representation using transfer matrices as well as all-electron Slater-type orbitals. These transformations were accomplished by means of the Local-Orbital Basis Suite Towards Electronic-Structure Reconstruction (LOBSTER) code [23, 24]. The DOS and pCOHP curves of Sr7N2Sn3 were visualized using the wxDragon code [25], while the program Diamond4 [26] was employed to create the representation showing the distributions of the Mulliken and Löwdin charges in the crystal structure of the ternary nitride stannide. In addition to the pCOHP and Mulliken as well as Löwdin population analyses, the electron localization function (ELF) [27, 28] of Sr7N2Sn3 was also computed by means of the aforementioned PAW method and analyzed as well as visualized utilizing the program Vesta [11].
3 Results and discussion
Sample I is mainly comprised of aggregates of black metallic platelet single crystals with a maximum length along the sides of 500 μm and a thickness of 20–50 μm. Figure 1 shows a scanning electron microscopic (SEM) image taken for a single crystal of Sample I with the EPMA apparatus. The EPMA data obtained from the platelet single crystal indicated that the specimen was made of 58.1% Sr, 36.7% Sn, 2.6% N, and 1.9% O (for a total 99.3 mass%), while the Na concentration was less than 0.01%. The oxygen in this specimen was likely present in a surface oxide or hydroxide layer generated while transferring the sample into the EPMA apparatus under ambient air. The Sr:Sn molar ratio in this crystal was calculated to be approximately 7:3, which agreed with the formula of Sr7N2Sn3 determined from the subsequent single crystal XRD analysis. In addition to the platelet single crystals of Sr7N2Sn3, black granular single crystals of SrSn with a maximum size of approximately 60 μm were also found in Sample I. The orthorhombic lattice parameters of this compound were a = 5.0658(3), b = 12.0252(6), and c = 4.4945(2) Å, consistent with the reported values (a = 5.064, b = 12.04, c = 4.494 Å [29]; a = 5.033(6), b = 12.00(2), c = 4.493(3) Å [30]; a = 5.045(5), b = 12.040(1), c = 4.495(5) Å [31]). Sample II was found to comprise granular single crystals of the antiperovskite Sr3N x Sn less than 70 μm in diameter. The cubic lattice parameter of these crystals was a = 5.24180(10) Å, similar to the value of a = 5.2351(5) Å reported for a powder sample of Sr3N0.74Sn [32].

Scanning electron microscopic (SEM) image of a Sr7N2Sn3 single crystal.
All XRD reflections from a platelet single crystal of Sr7N2Sn3 could be indexed using the orthorhombic lattice parameters a = 10.4082(2), b = 18.0737(4), and c = 7.43390(10) Å, and the possible space groups determined from systematic extinctions were P2na and Pmna. A crystal structure model for the Pmna space group was generated using the Intrinsic Phasing module of the Apex3 program package and the structure was refined to wR2 and R1 (all data) values of 3.98 and 1.80%, respectively. The results of this structure analysis are summarized in Table 1, and the refined atomic coordinates, anisotropic displacement parameters, and selected interatomic distances are provided in Tables 2, 3, and 4, respectively. The results for the SrSn and Sr3N x Sn single crystals obtained in the present study are also included in these tables.
Crystal data and refinement results for Sr7N2Sn3, SrSn, and Sr3N0.77Sn.
Chemical formula | Sr7N2Sn3 | SrSn | Sr3N0.77Sn |
---|---|---|---|
Formula weight, M r | 997.43 | 206.31 | 392.34 |
Habit/color | Platelet/black | Granule/black | Granule/black |
Size, mm3 | 0.173 × 0.117 × 0.032 | 0.065 × 0.050 × 0.050 | 0.070 × 0.052 × 0.049 |
Temperature T, K | 298(2) | 298(2) | 301(2) |
Crystal system | Orthorhombic | Orthorhombic | Cubic |
Space group | Pmna (No. 53) | Cmcm (No. 63) |
|
Unit cell dimensions | |||
a, Å | 10.4082(2) | 5.0658(3) | 5.24117(15) |
b, Å | 18.0737(4) | 12.0252(6) | |
c, Å | 7.43390(10) | 4.4945(2) | |
Unit cell volume V, Å3 | 1398.43(5) | 273.79(2) | 143.974(12) |
Z | 4 | 4 | 1 |
Calculated density, D cald, g cm−3 | 4.74 | 5.01 | 4.53 |
Wavelength λ, Å | 0.71073 | 0.71073 | 0.71073 |
Absorption correction | Numerical (Sadabs) | Multi-scan (Sadabs) | Multi-scan (Sadabs) |
Absorption coefficient μ, mm−1 | 31.7 | 28.2 | 31.8 |
Limiting indices hkl | −13 ≤ h ≤ 13 | −7 ≤ h ≤ 7 | −7 ≤ h ≤ 7 |
−23 ≤ k ≤ 22 | −18 ≤ k ≤ 18 | −7 ≤ k ≤ 7 | |
−9 ≤ l ≤ 9 | −6 ≤ l ≤ 6 | −7 ≤ l ≤ 7 | |
F(000), e | 1720 | 352 | 169 |
θ range data collection, deg | 2.25–27.51 | 3.39–33.18 | 3.89–31.39 |
Reflections collected/unique | 24,519/1708 | 6610/314 | 3151/73 |
R int | 0.0290 | 0.0218 | 0.0232 |
Data/restraints/parameters | 1708/0/65 | 314/0/10 | 73/0/6 |
R1a/wR2b (I > 2 σ(I)) | 0.0165/0.0390 | 0.0088/0.0203 | 0.0143/0.0343 |
R1a/wR2b (all data) | 0.0180/0.0398 | 0.0088/0.0203 | 0.0143/0.0343 |
Weight parametersb a/b | 0.0171/2.0590 | 0.0061/0.3301 | 0.0195/0.0220 |
Extinction coefficient χ | 0.00031(3) (Shelxl) | 0.0047(4) (Shelxl) | – |
Goodness-of-fit S c on F 2 | 1.081 | 1.329 | 1.340 |
Largest diff. peak/hole Δρ, e Å−3 | 1.31/−0.85 | 0.64/−0.63 | 1.09/−0.55 |
-
a R1 = Σ||F o| − |F c||/Σ|F o|. b wR2 = [Σw(F o 2 − F c 2)2/Σ(wF o 2)2]1/2, w = 1/[σ 2(F o 2) + (aP)2 + bP], where F o is the observed structure factor, F c is the calculated structure factor, σ is the standard deviation of F c 2, and P = (F o 2 + 2F c 2)/3. c S = [Σw(F o 2 − F c 2)2/(n − p)]1/2, where n is the number of reflections and p is the total number of parameters refined.
Site occupancies, atomic coordinates and equivalent isotropic displacement parameters (U eq in Å2) for Sr7N2Sn3, SrSn, and Sr3N0.77Sn.
Atom | Wyckoff position | Occup. | x | y | z | U eq a |
---|---|---|---|---|---|---|
Sr7N2Sn3 | ||||||
Sr1 | 8i | 1.0 | 0.25708(2) | 0.20426(2) | 0.00584(3) | 0.01686(7) |
Sr2 | 8i | 1.0 | 0.26526(3) | 0.40247(2) | 0.00438(3) | 0.01226(7) |
Sr3 | 4h | 1.0 | 0 | 0.10004(2) | 0.76327(5) | 0.02233(10) |
Sr4 | 4h | 1.0 | 0 | 0.30257(2) | 0.22833(5) | 0.01538(8) |
Sr5 | 4e | 1.0 | 0.25621(3) | 0 | 0 | 0.02063(9) |
N1 | 4g | 1.0 | 1/4 | 0.31064(14) | 1/4 | 0.0091(5) |
N2 | 4g | 1.0 | 1/4 | 0.89448(14) | 1/4 | 0.0138(6) |
Sn1 | 4h | 1.0 | 0 | 0.51271(2) | 0.20574(3) | 0.01284(8) |
Sn2 | 4h | 1.0 | 0 | 0.68405(2) | 0.24692(3) | 0.01247(7) |
Sn3 | 4h | 1.0 | 0 | 0.10342(2) | 0.24984(3) | 0.01248(7) |
SrSn | ||||||
Sr1 | 4c | 1.0 | 0 | 0.36391(2) | 1/4 | 0.01428(7) |
Sn1 | 4c | 1.0 | 0 | 0.07895(2) | 1/4 | 0.01370(6) |
Sr3N0.77Sn | ||||||
Sr1 | 3d | 1.0 | 1/2 | 0 | 0 | 0.02292(18) |
Sn1 | 1b | 1.0 | 1/2 | 1/2 | 1/2 | 0.01561(17) |
N1 | 1a | 0.767(17) | 0 | 0 | 0 | 0.0178(15) |
-
a U eq = (∑i ∑j U ij a i * a j * a i· a j)/3.
Anisotropic displacement parameters (U ij in Å2) for Sr7N2Sn3, SrSn, and Sr3N0.77Sn.
Atom | U 11 | U 22 | U 33 | U 13 | U 13 | U 23 |
---|---|---|---|---|---|---|
Sr7N2Sn3 | ||||||
Sr1 | 0.01960(14) | 0.01563(14) | 0.01534(13) | 0.00025(9) | −0.00057(8) | −0.00436(8) |
Sr2 | 0.01418(12) | 0.01196(12) | 0.01064(12) | 0.00151(8) | 0.00125(7) | 0.00344(7) |
Sr3 | 0.01407(18) | 0.0304(2) | 0.02256(19) | 0 | 0 | −0.00034(13) |
Sr4 | 0.00649(15) | 0.02354(19) | 0.01611(16) | 0 | 0 | −0.00009(12) |
Sr5 | 0.02295(19) | 0.0213(2) | 0.01763(18) | 0 | 0 | 0.00076(12) |
N1 | 0.0081(12) | 0.0103(12) | 0.0089(11) | 0 | −0.0002(9) | 0 |
N2 | 0.0137(13) | 0.0118(14) | 0.0160(13) | 0 | 0.0012(10) | 0 |
Sn1 | 0.01061(11) | 0.01321(12) | 0.01470(13) | 0 | 0 | −0.00066(9) |
Sn2 | 0.01144(12) | 0.01346(13) | 0.01250(11) | 0 | 0 | −0.00070(8) |
Sn3 | 0.01260(12) | 0.01145(13) | 0.01339(12) | 0 | 0 | 0.00019(8) |
SrN | ||||||
Sr1 | 0.01382(10) | 0.01460(11) | 0.01443(11) | 0 | 0 | 0 |
Sn1 | 0.01457(9) | 0.01432(8) | 0.01220(8) | 0 | 0 | 0 |
Sr3N0.77Sn | ||||||
Sr1 | 0.0188(2) | 0.02498(19) | 0.02498(19) | 0 | 0 | 0 |
Sn1 | 0.01561(17) | 0.01561(17) | 0.01561(17) | 0 | 0 | 0 |
N1 | 0.0178(15) | 0.0178(15) | 0.0178(15) | 0 | 0 | 0 |
Selected interatomic distances (Å) for Sr7N2Sn3, SrSn, and Sr3N0.77Sn.
Sr7N2Sn3 | |||
---|---|---|---|
N1–Sr2 (2×) | 2.4726(18) | Sn2–Sr2 (2×) | 3.4753(4) |
N1–Sr4 (2×) | 2.61111(16) | Sn2–Sr4 | 3.5412(5) |
N1–Sr1 (2×) | 2.6451(19) | Sn2–Sr2 (2×) | 3.6821(4) |
av. | 2.576 | Sn2–Sr1 (2×) | 3.7646(4) |
Sn2–Sr1 (2×) | 3.8425(4) | ||
N2–Sr3 (2×) | 2.60580(12) | Sn2–Sr3 | 3.9030(6) |
N2–Sr1 (2×) | 2.6091(18) | Sn2–Sr4 | 3.9084(5) |
N2–Sr5 (2×) | 2.6637(19) | av. | 3.717 |
av. | 2.626 | ||
Sn3–Sr4 | 3.6029(6) | ||
Sn1–Sn1 | 3.0932(5) | Sn3–Sr1 (2×) | 3.6074(4) |
Sn1–Sn2 | 3.1118(6) | Sn3–Sr3 | 3.6176(5) |
Sn3–Sr5 (2×) | 3.6593(4) | ||
Sn1–Sr2 (2×) | 3.5231(4) | Sn3–Sr3 | 3.6786(6) |
Sn1–Sr2 (2×) | 3.6398(4) | Sn3–Sr1 (2×) | 3.7110(4) |
Sn1–Sr2 (2×) | 3.7193(4) | Sn3–Sr5 (2×) | 3.7489(4) |
Sn1–Sr4 | 3.8017(6) | Sn3–Sr3 | 3.8173(5) |
Sn1–Sr2 (2×) | 3.8188(4) | av. | 3.681 |
av. | 3.689 | ||
SrSn | |||
Sn1–Sn1 | 2.9420(4) | Sn1–Sr1 | 3.4267(4) |
Sn1–Sr1 (4×) | 3.45512(15) | ||
Sn1–Sr1 (2×) | 3.6197(3) | ||
av. | 3.498 | ||
Sr3N0.77Sn | |||
N1–Sr1 (6×) | 2.62058(8) | Sn1–Sr1 (12×) | 3.70607(11) |
The N1 and N2 occupancies in Sr7N2Sn3 were refined to be 1.034(10) and 0.991(11), and these occupancies were fixed at 1.0 in the final refinement. The nitrogen content, x, in Sr3N x Sn determined based on the N1 occupancy refinement was 0.767(17), which was close to the value reported for Sr3N0.74Sn powder [32]. A lattice parameter of a = 10.4708(7) Å (twice that for the antiperovskite-type structure due to nitrogen defect ordering) has been reported for Sr3N0.59Sn prepared using a Li flux [33]. One-half of this value, 5.2354 Å, agreed with that of Sr3N0.74Sn (5.2351(5) Å), and no superstructure reflections corresponding to the nitrogen defect ordering were observed in the single-crystal XRD images of the present study.
The crystal structure of Sr7N2Sn3 consists of four layers of corner-sharing NSr6 octahedra separated by [Sn4] four-atom chains as shown in Figure 2(a) and (b). The plane of the four-layer slabs corresponds to the (110) plane of the ideal antiperovskite structure. This structure can be expressed as an inverse structure related to that of Ca2Nb2O7, in which slabs made of four perovskite layers [Ca2Nb4O14] are separated by two layers of Ca atoms (Figure 2(c) and (d)) [34]. For this reason, Sr7N2Sn3 can also be expressed as [Sn4][Sn2N4Sr14], following the representation of Ca2[Ca2Nb4O14]. However, the [Sn2N4Sr14] layers are stacked in the b-axis direction, while the [Ca2Nb4O14] layers are stacked in the a-axis direction with a c/4 shift (Figure 2(b) and (d)). The results of single crystal XRD experiments indicate that the plate surface of the single crystal shown in Figure 1 equated to the slab surface perpendicular to the b axis, while the step lines observed on the crystal were in the a-axis direction.
![Figure 2:
Projections of the crystal structure of Sr7N2Sn3 on (a) the crystallograpic bc plane (0.0 ≤ a < 0.5) and (b) the ab plane, and for Ca2Nb2O7 [34] on (c) the ab plane (0.0 ≤ c < 0.5) and (d) the ac plane.](/document/doi/10.1515/znb-2021-0097/asset/graphic/j_znb-2021-0097_fig_002.jpg)
Projections of the crystal structure of Sr7N2Sn3 on (a) the crystallograpic bc plane (0.0 ≤ a < 0.5) and (b) the ab plane, and for Ca2Nb2O7 [34] on (c) the ab plane (0.0 ≤ c < 0.5) and (d) the ac plane.
Figure 3(a) shows the antiperovskite region of Sr7N2Sn3, in which Sn3 is at the A site surrounded by 12 Sr atoms, and there are two N1-centered Sr12Sr22Sr42 octahedra and six N2-centered Sr12Sr32Sr52 octahedra. The Sn3‒Sr interatomic distances range from 3.6029(6) to 3.8173(5) Å and the average distance of 3.717 Å is close to the Sn1‒Sr1 distance of 3.70651(8) Å determined for antiperovskite Sr3N0.77Sn (Table 4). N1‒Sr and N2‒Sr distances are in the ranges of 2.4726(18)–2.6451(19) and 2.60580(12)–2.6637(19) Å, respectively, and the N1-centered Sr octahedra are more distorted than the N2-centered ones. The average N1‒Sr and N2‒Sr distances are 2.576 and 2.626 Å, respectively, and the average values of 2.601 Å is a little shorter than the N1‒Sr1 distance of Sr3N0.77Sn (2.62090(6) Å).

Atomic arrangements in crystals of Sr7N2Sn3 (a) in the anti-perovskite slab and (b) around Sn3 and the polyanion Sn2‒Sn1‒Sn1‒Sn2.
Each Sn1 in the [Sn4] four-membered zigzag chain is surrounded by nine Sr atoms. The coordination environment of each Sn2 is similar to that of Sn3 in the antiperovskite unit except that one of the 12 Sr atoms at the C site is replaced with a Sn atom at Sn1. The Sn2‒Sr distances are in the range from 3.4753(4) to 3.9084(5) Å, and the average distance of 3.717 Å is longer than the averages of the Sn1‒Sr (3.689 Å) and Sn3‒Sr (3.681 Å) distances, but close to the Sn1‒Sr1 distance of Sr3N0.77Sn (3.70607(11) Å). The shortest value of 3.4753(4) Å among the Sn2‒Sr distances is in the range of Sn1‒Sr1 distances (3.4267(4)–3.6197(3) Å, av. 3.498 Å) observed for sevenfold coordinated Sn atoms in SrSn (Table 4).
The Sn1‒Sn1 and Sn1‒Sn2 bond lengths in the [Sn4] four-membered chains (that is, the Sn2‒Sn1‒Sn1‒Sn2 chains) were determined to be 3.0932(5) and 3.1118(6) Å, respectively, and the ∠Sn2‒Sn1‒Sn1 bond angle was 104.187(16)°. The bond lengths are comparable to the Sn‒Sn distances in metallic β-Sn (3.02192(10)–3.1810(3) Å [35]).
Similar [Sn4] four-membered zigzag chains have been reported in the structure of Ca7Sn6 [36], in which the Sn1‒Sn1 and Sn1‒Sn2 bond lengths of [Sn4] are 2.8987(6) and 2.9238(6) Å, respectively. These are shorter than the lengths found in Sr7N2Sn3, although a ∠Sn2‒Sn1‒Sn1 angle of 104.353(16)° in Ca7Sn6 is consistent with that in Sr7N2Sn3. [Sn2] dumbbells are also contained in Ca7Sn6. Assuming Sn‒Sn single bonds, formal charges of −6 and −10 are expected for [Sn2] and [Sn4] based on the (8−N) rule [37]. In this model, two electrons are unaccounted for because 14 electrons are provided from the seven electropositive Ca atoms in Ca7Sn6. Siggelkow et al. proposed the resonance hybrid structures [Ge–Ge]6− +
The formal charge of Sn atoms at Sn3 in the antiperovskite slab of Sr7N2Sn3 is −4, so that the charge of the [Sn4] four-membered chain must be −8 to meet the requirement of charge neutrality ({(Sr14)28+(N4)12−(Sn2)8−}[Sn4]8−). However, the Sn1‒Sn1 single bond length obtained for the
Assuming that each [Sn4] chain has single Sn‒Sn bonds (even though the bond lengths are longer than the expected values of approximately 2.9420(4) Å), the formal positive and negative charges could be balanced by introducing oxygen as {(Sr14)28+ (N2)6−(O2)4−(Sn2)8−}[Sn4]10−. The wR2 (all data) factors obtained from crystal structure refinements with the Sr7(N0.5O0.5)2Sn3, Sr7(N1O2)Sn3, and Sr7(O1N2)Sn3 models were 4.32, 4.58, and 4.26%, respectively, all of which were larger than the value for Sr7N2Sn3 (3.98%).
Gärtner and Korber have pointed out that the bond lengths in Zintl polyanions are determined by the configuration and the valence electron numbers of the constituent anions as well as the size of the electropositive cations around the polyanions [40]. In fact, SrSn is isostructural with CaSn, and the Sn‒Sn distance in the infinite zigzag chain,
An examination of the DOS curves for Sr7N2Sn3 (Figure 4) reveals that the states in the energy regions near the Fermi level, E F, largely originate from the Sn-5p and N-2p atomic orbitals with minor contributions from the Sr-5s atomic orbitals. Because the states related to the latter type of atomic orbitals are mainly located above E F, it can be concluded that the strontium atoms are oxidized, and, hence, act as valence electron donors in Sr7N2Sn3. A closer inspection of the DOS in the energy regions near E F indicates that the Fermi level is located below a broad pseudogap in the DOS of Sr7N2Sn3, which should lead to metallic conductivity. Furthermore, it is remarkable that E F is located below and not in the pseudogap, whereas the location of a Fermi level in a pseudogap is typically related to an electronically favorable situation for a given solid-state compound [41], [42], [43], [44], [45]. This circumstance can be explained based on an examination of the –pCOHP curves for Sr7N2Sn3 (Figure 4): as the −pCOHP of all interactions change from bonding to antibonding states below E F, the populations of further states would lead to a decrease of the bond energy and, accordingly, an electronically less favorable situation [41].

(a)–(c) Net as well as atom- and orbital-projected densities-of-states (DOS) curves of Sr7N2Sn3: the orbital-projected DOS correspond to those states providing the largest contributions to the DOS in the energy regions near the Fermi level, E F; (d) projected crystal orbital Hamilton populations (pCOHP) of diverse interactions in Sr7N2Sn3: the cumulative −IpCOHP per cell values and their percentages to the net bonding capacities have been included using the font color of a respective sort of interaction.
To analyze the nature of bonding in Sr7N2Sn3 in more detail, we also projected the diverse cumulative −IpCOHP per cell values, i.e. the sums of all −IpCOHP per bond values of a given sort of interaction within a unit cell, as percentages to the net bonding capacities – a procedure that has been described in detail elsewhere [46]. The comparison of the percentages to net bonding capacities for the diverse interactions reveals that the largest contributions arise from the Sr–N and Sr–Sn interactions. In this connection it is remarkable that the percentages of Sn–Sn interactions are nearly half as great as those of the Sr–N contacts, while the amount of the latter separations per cell (48) is much higher than that of the homoatomic interactions (6). This is because the Sr−N interactions correspond to much smaller −IpCOHP per bond values (<−IpCOHP per bond> = 0.517 eV) than the Sn–Sn interactions (<−IpCOHP per bond> = 1.937 eV), which also correspond to higher −IpCOHP per bond values than those of the heteroatomic Sr–Sn interactions (<−IpCOHP per bond> = 0.302 eV). How can this finding be explained? As the analysis of the DOS curves for Sr7N2Sn3 clearly indicates that the strontium atoms act as valence electron donors, the Sr–Sn and Sr–N interactions should be depicted as rather polar bonds, whose nature translates into a less bonding character relative to that of the Sn–Sn interactions.
The suggested presence of rather polar Sr–Sn and Sr–N bonds in Sr7N2Sn3 is also backed by Mulliken and Löwdin (Figure 5) population analyses, which underline the role of strontium as a valence-electron donor for the tin and nitrogen atoms. In this connection it is remarkable that the Sn1, Sn2, and Sn3 atoms differ in their respective Mulliken and Löwdin charges: the Sn3 atoms are most reduced among all the different Sn atoms, while the Sn2 atoms are more reduced than the Sn1 atoms. This result indicates that the Sn1 and Sn2 atoms correspond to different electronic configurations which could translate into a structural configuration as expected for a [Sn4]8− polyanion. This conclusion is further backed by an examination of the electron localization function (ELF; Figure 5), as localization domains for high values of the localization parameter (η > 0.8) are located around the Sn1 and Sn2 atoms and could be regarded as Sn1 and Sn2 lone pairs. In addition, the relatively large −IpCOHP per bond values corresponding to the separations between the tin atoms also point to the presence of Sn–Sn interactions with a strong bonding character as expected for covalent bonds; yet, the antibonding character of the Sn–Sn interactions at E F weakens these homoatomic bonds and, ultimately, translates into enlarged Sn–Sn distances – a circumstance that has also been encountered by previous research on tin-containing compounds [47]. In this connection it should also be noted that the computed Mulliken and Löwdin charges are evidently smaller than the charges predicted by the Zintl–Klemm treatment. This outcome means that the Sr–Sn and Sr–N bonds are not solely constructed of closed-shell species with opposite electric charges as expected for ionic bonds, but that also certain covalent contributions must be present in the heteroatomic interactions in derogation from the Zintl–Klemm ideal.

(a) Averaged Mulliken and Löwdin (in parentheses) charges in Sr7N2Sn3: the diverse types of strontium polyhedra enclosing the tin and nitrogen atoms are shown in the separate representations; (b) representation of the electron localization function (ELF) for a Sr7N2Sn3 unit cell in the bc plane at x = 0.5; (c) ELF isosurface with η = 0.8: sectional views of localization domains which correspond to values of the ELF larger than 0.8 have been included.
4 Conclusions
Herein, we have reported on the crystal and electronic structure of the ternary compound Sr7N2Sn3, which was obtained from reactions of strontium and tin in the presence of a sodium flux and N2 generated in the thermal decomposition of sodium azide. The crystal structure of this compound can be related to the inverse structure of the previously reported Ca2Nb2O7, and it is composed of antiperovskite-fashioned slabs separated by four-membered tin units, [Sn4]. Although for Sr7N2Sn3 the Zintl–Klemm formalism predicts an electron-precise valence electron distribution according to the formula (Sr2+)14(N3−)4(Sn4−)2([Sn4]8−), there are certain structural peculiarities that cannot be explained by applying the aforementioned formalism. From a first principles-based examination of the electronic structure, it is clear that there is a significant valence electron transfer from the strontium to the tin and nitrogen atoms within these heteroatomic contacts, while the bonding nature of the Sn–Sn interactions should be depicted as covalent – a picture that is in full agreement with the predictions based on the Zintl–Klemm treatment. And yet, there are also certain differences between the predictions of the Zintl–Klemm treatment and the results of the first principles-based bonding analyses, which allow us to understand the aforementioned structural peculiarities.
Dedicated to: Professor Richard Dronskowski of the RWTH Aachen on the occasion of his 60th birthday.
Funding source: Japan Society for the Promotion of Science
Award Identifier / Grant number: JP19H02794
Funding source: Verband der Chemischen Industrie
Acknowledgments
The authors wish to thank Prof. Richard Dronskowski for the allocation of the computer cluster of the Chemistry Department of RWTH Aachen University. The authors are also grateful to Mitsuyo Takaishi for her assistance with the sample preparation and Takashi Kamaya for performing the EPMA.
-
Author contributions: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.
-
Research funding: A Kakenhi Grant-in-Aid (No. JP19H02794) from the Japan Society for the Promotion of Science (JSPS). The Verband der Chemischen Industrie (FCI) through a Liebig stipend to S.S.
-
Conflict of interest statement: The authors declare no conflicts of interest regarding this article.
References
1. Tilley, R. J. D. Perovskites: Structure–Property Relationships; John Wiley & Sons Ltd.: Chichester, 2016.10.1002/9781118935651Suche in Google Scholar
2. Wang, Y., Zhang, H., Zhu, J., Lü, X., Li, S., Zou, R., Zhao, Y. Adv. Mater. 2020, 32, 1905007; https://doi.org/10.1002/adma.201905007.Suche in Google Scholar
3. Niewa, R. Eur. J. Inorg. Chem. 2019, 2019, 3647–3660; https://doi.org/10.1002/ejic.201900756.Suche in Google Scholar
4. Inorganic Crystal Structure Database (ICSD). Version 2021-1; FIZ Karlsruhe - Leibniz Institute for Information Infrastructure: Karlsruhe, 2020.Suche in Google Scholar
5. Gäbler, F., Bräunling, D., Schnelle, W., Schellenberg, I., Pöttgen, R., Niewa, R. Z. Anorg. Allg. Chem. 2011, 637, 977–982; https://doi.org/10.1002/zaac.201000416.Suche in Google Scholar
6. Gál, Z. A., Clarke, S. J. Chem. Commun. 2005, 728–730; https://doi.org/10.1039/b413534b.Suche in Google Scholar
7. Yamane, H., Di Salvo, F. J. Prog. Solid State Chem. 2018, 51, 27–40; https://doi.org/10.1016/j.progsolidstchem.2017.08.002.Suche in Google Scholar
8. Apex3 (version 2018/1); Bruker AXS Inc.: Madison, Wisconsin (USA), 2018.Suche in Google Scholar
9. Sheldrick, G. M. Sadabs (version 2014/2); Bruker AXS Inc.: Madison, Wisconsin (USA), 2014.Suche in Google Scholar
10. Sheldrick, G. Acta Crystallogr. 2015, C71, 3–8.Suche in Google Scholar
11. Momma, K., Izumi, F. J. Appl. Crystallogr. 2011, 44, 1272–1276; https://doi.org/10.1107/s0021889811038970.Suche in Google Scholar
12. Blöchl, P. E. Phys. Rev. B 1994, 50, 17953–17979; https://doi.org/10.1103/physrevb.50.17953.Suche in Google Scholar
13. Kresse, G., Furthmüller, J. Comput. Mater. Sci. 1996, 6, 15–50; https://doi.org/10.1016/0927-0256(96)00008-0.Suche in Google Scholar
14. Kresse, G., Furthmüller, J. Phys. Rev. B 1996, 54, 11169–11186; https://doi.org/10.1103/physrevb.54.11169.Suche in Google Scholar PubMed
15. Kresse, G., Hafner, J. Phys. Rev. B 1993, 47, 558–561; https://doi.org/10.1103/physrevb.47.558.Suche in Google Scholar PubMed
16. Kresse, G., Joubert, D. Phys. Rev. B 1999, 59, 1758–1775; https://doi.org/10.1103/physrevb.59.1758.Suche in Google Scholar
17. Kresse, G., Marsman, M., Furthmüller, J. Vienna Ab Initio Simulation Package (VASP), The Guide, Computational Materials Physics; Faculty of Physics, Universität Wien: Vienna (Austria), 2014.Suche in Google Scholar
18. Perdew, J. P., Burke, K., Ernzerhof, M. Phys. Rev. Lett. 1996, 77, 3865–3868; https://doi.org/10.1103/physrevlett.77.3865.Suche in Google Scholar
19. Ertural, C., Steinberg, S., Dronskowski, R. RSC Adv. 2019, 9, 29821–29830; https://doi.org/10.1039/c9ra05190b.Suche in Google Scholar PubMed PubMed Central
20. Deringer, V. L., Tchougreeff, A. L., Dronskowski, R. J. Phys. Chem. A 2011, 115, 5461–5466; https://doi.org/10.1021/jp202489s.Suche in Google Scholar PubMed
21. Dronskowski, R., Blöchl, P. E. J. Phys. Chem. 1993, 97, 8617–8624; https://doi.org/10.1021/j100135a014.Suche in Google Scholar
22. Steinberg, S., Dronskowski, R. Crystals 2018, 8, 225; https://doi.org/10.3390/cryst8050225.Suche in Google Scholar
23. Maintz, S., Deringer, V. L., Tchougreéff, A. L., Dronskowski, R. J. Comput. Chem. 2013, 34, 2557–2567; https://doi.org/10.1002/jcc.23424.Suche in Google Scholar PubMed
24. Maintz, S., Deringer, V. L., Tchougreéff, A. L., Dronskowski, R. J. Comput. Chem. 2016, 37, 1030–1035; https://doi.org/10.1002/jcc.24300.Suche in Google Scholar PubMed PubMed Central
25. Eck, B. wxDragon (version 2.2.3); RWTH Aachen University: Aachen (Germany), 2020.Suche in Google Scholar
26. Brandenburg, K. Diamond (version 4.6.3), Crystal and Molecular Structure Visualization; Crystal Impact - K. Brandenburg & H. Putz GbR: Bonn (Germany), 2020.Suche in Google Scholar
27. Silvi, B., Savin, A. Nature (London) 1994, 371, 683–686; https://doi.org/10.1038/371683a0.Suche in Google Scholar
28. Savin, A., Nesper, R., Wengert, S., Fässler, T. F. Angew. Chem. Int. Ed. Engl. 1997, 36, 1808–1832; https://doi.org/10.1002/anie.199718081.Suche in Google Scholar
29. Merlo, F., Fornasini, M. L. J. Less Common. Met. 1967, 13, 603–610; https://doi.org/10.1016/0022-5088(67)90105-1.Suche in Google Scholar
30. Rieger, W., Parthé, E. Acta Crystallogr. 1967, 22, 919–922; https://doi.org/10.1107/s0365110x67001793.Suche in Google Scholar
31. Widera, A., Schäfer, H. J. Less Common. Met. 1981, 77, 29–36; https://doi.org/10.1016/0022-5088(81)90005-9.Suche in Google Scholar
32. Gäbler, F., Kirchner, M., Schnelle, W., Schmitt, M., Rosner, H., Niewa, R. Z. Anorg. Allg. Chem. 2005, 631, 397–402.10.1002/zaac.200400344Suche in Google Scholar
33. Pathak, M., Stoiber, D., Bobnar, M., Ormeci, A., Prots, Y., Niewa, R., Höhn, P. Z. Anorg. Allg. Chem. 2018, 644, 161–167; https://doi.org/10.1002/zaac.201700368.Suche in Google Scholar
34. Scheunemann, K., Müller-Buschbaum, H. J. Inorg. Nucl. Chem. 1974, 36, 1965–1970; https://doi.org/10.1016/0022-1902(74)80709-8.Suche in Google Scholar
35. Wołcyrz, M., Kubiak, R., Maciejewski, S. Phys. Status Solidi B 1981, 107, 245–253; https://doi.org/10.1002/pssb.v107:2.10.1002/pssb.2221070125Suche in Google Scholar
36. Palenzona, A., Manfrinetti, P., Fornasini, M. L. J. Alloys Compd. 2000, 312, 165–171; https://doi.org/10.1016/s0925-8388(00)01150-6.Suche in Google Scholar
37. Schäfer, H., Eisenmann, B., Müller, W. Angew Chem. Int. Ed. Engl. 1973, 12, 694–712; https://doi.org/10.1002/anie.197306941.Suche in Google Scholar
38. Thewlis, J., Davey, A. R. Nature 1954, 174, 1011; https://doi.org/10.1038/1741011a0.Suche in Google Scholar
39. Leon-Escamilla, E. A., Corbett, J. D. Inorg. Chem. 2001, 40, 1226–1233; https://doi.org/10.1021/ic0010306.Suche in Google Scholar
40. Gärtner, S., Korber, N. Struct. Bonding Berlin Ger. 2011, 140, 25–57; https://doi.org/10.1007/430_2011_43.Suche in Google Scholar
41. Steinberg, S., Brgoch, J., Miller, G. J., Meyer, G. Inorg. Chem. 2012, 51, 11356–11364; https://doi.org/10.1021/ic300838a.Suche in Google Scholar
42. Zhang, M., Cheng, K., Yan, H., Wei, Q., Zheng, B. Sci. Rep. 2016, 6, 36911; https://doi.org/10.1038/srep36911.Suche in Google Scholar
43. Bigun, I., Steinberg, S., Smetana, V., Mudryk, Y., Kalychak, Y., Havela, L., Pecharsky, V., Mudring, A.-V. Chem. Mater. 2017, 29, 2599–2614; https://doi.org/10.1021/acs.chemmater.6b04782.Suche in Google Scholar
44. Peterson, G. G. C., Yannello, V. J., Fredrickson, D. C. Angew. Chem. Int. Ed. 2017, 56, 10145–10150; https://doi.org/10.1002/anie.201702156.Suche in Google Scholar
45. Clark, W. P., Steinberg, S., Dronskowski, R., McCammon, C., Kupenko, I., Bykov, M., Dubrovinsky, L., Akselrud, L. G., Schwarz, U., Niewa, R. Angew. Chem. Int. Ed. 2017, 56, 7302–7306; https://doi.org/10.1002/anie.201702440.Suche in Google Scholar
46. Gladisch, F. C., Steinberg, S. Crystals 2018, 8, 80; https://doi.org/10.3390/cryst8020080.Suche in Google Scholar
47. Zürcher, F., Nesper, R., Hoffmann, S., Fässler, T. F. Z. Anorg. Allg. Chem. 2001, 627, 2211–2219; https://doi.org/10.1002/1521-3749(200109)627:9<2211::aid-zaac2211>3.0.co;2-2.10.1002/1521-3749(200109)627:9<2211::AID-ZAAC2211>3.0.CO;2-2Suche in Google Scholar
© 2021 Walter de Gruyter GmbH, Berlin/Boston
Artikel in diesem Heft
- Frontmatter
- In this issue
- Laudatio/Preface
- Celebrating the 60th birthday of Richard Dronskowski
- Review
- Orbital-selective electronic excitation in phase-change memory materials: a brief review
- Research Articles
- Solving the puzzle of the dielectric nature of tantalum oxynitride perovskites
- d- and s-orbital populations in the d block: unbound atoms in physical vacuum versus chemical elements in condensed matter. A Dronskowski-population analysis
- Single-crystal structures of A 2SiF6 (A = Tl, Rb, Cs), a better structure model for Tl3[SiF6]F, and its novel tetragonal polymorph
- Na2La4(NH2)14·NH3, a lanthanum-rich intermediate in the ammonothermal synthesis of LaN and the effect of ammonia loss on the crystal structure
- Linarite from Cap Garonne
- Salts of octabismuth(2+) polycations crystallized from Lewis-acidic ionic liquids
- High-temperature diffraction experiments and phase diagram of ZrO2 and ZrSiO4
- Thermal conversion of the hydrous aluminosilicate LiAlSiO3(OH)2 into γ-eucryptite
- Crystal structure of mechanochemically prepared Ag2FeGeS4
- Effect of nanostructured Al2O3 on poly(ethylene oxide)-based solid polymer electrolytes
- Sr7N2Sn3: a layered antiperovskite-type nitride stannide containing zigzag chains of Sn4 polyanions
- Exploring the frontier between polar intermetallics and Zintl phases for the examples of the prolific ALnTnTe3-type alkali metal (A) lanthanide (Ln) late transition metal (Tn) tellurides
- Zwitterion coordination to configurationally flexible d 10 cations: synthesis and characterization of tetrakis(betaine) complexes of divalent Zn, Cd, and Hg
- An approach towards the synthesis of lithium and beryllium diphenylphosphinites
- Synthesis, crystal and electronic structure of CaNi2Al8
- Crystal and electronic structure of the new ternary phosphide Ho5Pd19P12
- Synthesis, structure, and magnetic properties of the quaternary oxysulfides Ln 5V3O7S6 (Ln = La, Ce)
- Synthesis, crystal and electronic structure of BaLi2Cd2Ge2
- Structural variations of trinitrato(terpyridine)lanthanoid complexes
- Preparation of CoGe2-type NiSn2 at 10 GPa
- Controlled exposure of CuO thin films through corrosion-protecting, ALD-deposited TiO2 overlayers
- Experimental and computational investigations of TiIrB: a new ternary boride with Ti1+x Rh2−x+y Ir3−y B3-type structure
- Synthesis and crystal structure of the lanthanum cyanurate complex La[H2N3C3O3]3 · 8.5 H2O
- Cd additive effect on self-flux growth of Cs-intercalated NbS2 superconducting single crystals
- 14N, 13C, and 119Sn solid-state NMR characterization of tin(II) carbodiimide Sn(NCN)
- Superexchange interactions in AgMF4 (M = Co, Ni, Cu) polymorphs
- Copper(I) iodide-based organic–inorganic hybrid compounds as phosphor materials
- On iodido bismuthates, bismuth complexes and polyiodides with bismuth in the system BiI3/18-crown-6/I2
- Synthesis, crystal structure and selected properties of K2[Ni(dien)2]{[Ni(dien)]2Ta6O19}·11 H2O
- First low-spin carbodiimide, Fe2(NCN)3, predicted from first-principles investigations
- A novel ternary bismuthide, NaMgBi: crystal and electronic structure and electrical properties
- Magnetic properties of 1D spin systems with compositional disorder of three-spin structural units
- Amine-based synthesis of Fe3C nanomaterials: mechanism and impact of synthetic conditions
- Enhanced phosphorescence of Pd(II) and Pt(II) complexes adsorbed onto Laponite for optical sensing of triplet molecular dioxygen in water
- Theoretical investigations of hydrogen absorption in the A15 intermetallics Ti3Sb and Ti3Ir
- Assembly of cobalt-p-sulfonatothiacalix[4]arene frameworks with phosphate, phosphite and phenylphosphonate ligands
- Chiral bis(pyrazolyl)methane copper(I) complexes and their application in nitrene transfer reactions
- UoC-6: a first MOF based on a perfluorinated trimesate ligand
- PbCN2 – an elucidation of its modifications and morphologies
- Flux synthesis, crystal structure and electronic properties of the layered rare earth metal boride silicide Er3Si5–x B. An example of a boron/silicon-ordered structure derived from the AlB2 structure type
Artikel in diesem Heft
- Frontmatter
- In this issue
- Laudatio/Preface
- Celebrating the 60th birthday of Richard Dronskowski
- Review
- Orbital-selective electronic excitation in phase-change memory materials: a brief review
- Research Articles
- Solving the puzzle of the dielectric nature of tantalum oxynitride perovskites
- d- and s-orbital populations in the d block: unbound atoms in physical vacuum versus chemical elements in condensed matter. A Dronskowski-population analysis
- Single-crystal structures of A 2SiF6 (A = Tl, Rb, Cs), a better structure model for Tl3[SiF6]F, and its novel tetragonal polymorph
- Na2La4(NH2)14·NH3, a lanthanum-rich intermediate in the ammonothermal synthesis of LaN and the effect of ammonia loss on the crystal structure
- Linarite from Cap Garonne
- Salts of octabismuth(2+) polycations crystallized from Lewis-acidic ionic liquids
- High-temperature diffraction experiments and phase diagram of ZrO2 and ZrSiO4
- Thermal conversion of the hydrous aluminosilicate LiAlSiO3(OH)2 into γ-eucryptite
- Crystal structure of mechanochemically prepared Ag2FeGeS4
- Effect of nanostructured Al2O3 on poly(ethylene oxide)-based solid polymer electrolytes
- Sr7N2Sn3: a layered antiperovskite-type nitride stannide containing zigzag chains of Sn4 polyanions
- Exploring the frontier between polar intermetallics and Zintl phases for the examples of the prolific ALnTnTe3-type alkali metal (A) lanthanide (Ln) late transition metal (Tn) tellurides
- Zwitterion coordination to configurationally flexible d 10 cations: synthesis and characterization of tetrakis(betaine) complexes of divalent Zn, Cd, and Hg
- An approach towards the synthesis of lithium and beryllium diphenylphosphinites
- Synthesis, crystal and electronic structure of CaNi2Al8
- Crystal and electronic structure of the new ternary phosphide Ho5Pd19P12
- Synthesis, structure, and magnetic properties of the quaternary oxysulfides Ln 5V3O7S6 (Ln = La, Ce)
- Synthesis, crystal and electronic structure of BaLi2Cd2Ge2
- Structural variations of trinitrato(terpyridine)lanthanoid complexes
- Preparation of CoGe2-type NiSn2 at 10 GPa
- Controlled exposure of CuO thin films through corrosion-protecting, ALD-deposited TiO2 overlayers
- Experimental and computational investigations of TiIrB: a new ternary boride with Ti1+x Rh2−x+y Ir3−y B3-type structure
- Synthesis and crystal structure of the lanthanum cyanurate complex La[H2N3C3O3]3 · 8.5 H2O
- Cd additive effect on self-flux growth of Cs-intercalated NbS2 superconducting single crystals
- 14N, 13C, and 119Sn solid-state NMR characterization of tin(II) carbodiimide Sn(NCN)
- Superexchange interactions in AgMF4 (M = Co, Ni, Cu) polymorphs
- Copper(I) iodide-based organic–inorganic hybrid compounds as phosphor materials
- On iodido bismuthates, bismuth complexes and polyiodides with bismuth in the system BiI3/18-crown-6/I2
- Synthesis, crystal structure and selected properties of K2[Ni(dien)2]{[Ni(dien)]2Ta6O19}·11 H2O
- First low-spin carbodiimide, Fe2(NCN)3, predicted from first-principles investigations
- A novel ternary bismuthide, NaMgBi: crystal and electronic structure and electrical properties
- Magnetic properties of 1D spin systems with compositional disorder of three-spin structural units
- Amine-based synthesis of Fe3C nanomaterials: mechanism and impact of synthetic conditions
- Enhanced phosphorescence of Pd(II) and Pt(II) complexes adsorbed onto Laponite for optical sensing of triplet molecular dioxygen in water
- Theoretical investigations of hydrogen absorption in the A15 intermetallics Ti3Sb and Ti3Ir
- Assembly of cobalt-p-sulfonatothiacalix[4]arene frameworks with phosphate, phosphite and phenylphosphonate ligands
- Chiral bis(pyrazolyl)methane copper(I) complexes and their application in nitrene transfer reactions
- UoC-6: a first MOF based on a perfluorinated trimesate ligand
- PbCN2 – an elucidation of its modifications and morphologies
- Flux synthesis, crystal structure and electronic properties of the layered rare earth metal boride silicide Er3Si5–x B. An example of a boron/silicon-ordered structure derived from the AlB2 structure type