Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

On the structure of heptalithium distannide: Li7Sn2 or Li7−xSn2 (0.3 < x < 0.5)

Salina Rahmana, Kowsik Ghosha, Alexander Ovchinnikovab, Anirudh Nandakumarc, Candace K. Chanc and Svilen Bobev*a
aDepartment of Chemistry and Biochemistry, University of Delaware, Newark, Delaware 19716, USA. E-mail: bobev@udel.edu
bFaculty of Chemistry and Food Chemistry, Technische Universität Dresden, 01062 Dresden, Germany
cMaterials Science and Engineering, School for Engineering of Matter, Transport and Energy, Arizona State University, P.O. Box 876106, Tempe, Arizona 85827, USA

Received 30th April 2025 , Accepted 22nd July 2025

First published on 29th July 2025


Abstract

Li–Sn intermetallics are promising candidates for high-capacity Li-ion battery anodes, yet, the precise crystal structures of some “known” binary phases remain unresolved. One such phase, Li7Sn2, originally reported in 1975, is re-examined here using single-crystal X-ray diffraction methods. Its orthorhombic structure (space group Cmmm, no. 65) features isolated Sn atoms and Sn–Sn dumbbells in a matrix of Li atoms. The present study indicates occupational disorder on one of the Li sites in Li7Sn2, meaning that it is a non-stoichiometric phase with revised composition of Li7−xSn2 (0.3 < x < 0.5). In the most extreme case, when the Li atomic site in question is completely empty, the composition becomes Li6.5Sn2 or Li13Sn4. Analysis of the chemical bonding through electronic structure calculations indicates significant covalency of Sn–Sn and Li–Sn interactions.


Introduction

Lithium-ion batteries (LIBs) are rechargeable energy storage devices that have become widely used since their introduction in the early 1990s. Graphite is the typical anode in LIBs, as it has a good cycling performance with a theoretical capacity of 372 mAh g−1 (LiC6).1 The other group 14 elements, referred to as tetrels, Si, Ge, and Sn, have also been studied for application as potential anode materials in LIBs. The tetrels (Tt = Si, Ge, and Sn, hereafter) form a variety of intermetallic LixTty compounds with Li, some of which are rather Li-rich. Hence, compared to graphite, Si, Ge, and Sn can provide much higher theoretical capacities: 4200 mAh g−1 for Li17Si4, 1600 mAh g−1 for Li17Ge4, and 990 mAh g−1 Li17Sn4.1

However, the volume expansion in fully lithiated states is problematic (viz. 434% (399%) for amorphous (crystalline) Li–Si2), leading to diminished capacity and reduced lifespan. Although not as attractive from the point of view of increased weight, Li–Ge/Sn compounds are gaining attention as more viable alternatives for LIB applications since their volume expansions are significantly smaller, especially for Sn.2 To design and develop improved anode materials, one can argue that it is essential to gain a deeper understanding of the structural chemistry of the various Li–Ge and Li–Sn phases that form during the lithiation of Ge- and Sn-based anodes.

The early structural work on the binary Li–Sn phases is from the 1950s/1960s. Based on literature searches and the information found in the ICSD (Inorganic Crystal Structure Database), there are eight crystallographically characterized Li–Sn phases and these are: Li2Sn5[thin space (1/6-em)]3 (the correct composition and structure of a phase initially reported as LiSn2[thin space (1/6-em)]3) LiSn,4 Li7Sn3,5 Li5Sn2 (or rather Li5−xSn2+x),6,7 Li13Sn5,8 Li7Sn2,9 Li17Sn4 (or Li17.05Sn4, previously known as Li21Sn5 and/or Li22Sn5 and often referred to as Li4.4Sn),10,11 and Li5Sn, which has the highest amount of Li.12 The latter is also the most recently discovered phase, which is metastable and not found in the experimental phase diagram.13 Theoretical predictions indicate the presence of additional, hitherto unidentified members near the bottom of the convex hull. For example, high-throughput density functional theory (DFT) calculations on the Li–Sn systems indicate several ground-state and metastable crystal structures, such as Li2Sn3, Li2Sn, Li3Sn2, Li5Sn3, Li8Sn3, Li3Sn, Li4Sn, and Li7Sn,14 none of which have been experimentally confirmed to date.

In addition to the structures of the two thermodynamically stable phases that have already been re-assessed, there remain controversies surrounding the Li7Sn2 phase. The latter was discovered in 1975 and reported to be isotypic with Li7Ge2 (published in 197215). However, in 1988, the composition of the Li7Ge2 phase has been questioned,16 and suggested to be Li6.5Ge2[triple bond, length as m-dash]Li13Ge4. The proposed reformulation of Li7Ge2 as Li13Ge4[thin space (1/6-em)]16 precedes by more than two decades the re-examination of the structure of Li13Si4, which had also been previously known as Li7Si2 (N. B. Li13Si4 and Li13Ge4 are not isostructural).17 In the case of the Li–Si system, very careful refinement of the structure from single-crystal X-ray diffraction data carried out by Fässler et al. identified discrepancies with regards to the occupation of Li on one of the positions,18 which validates a revised model. It is acknowledged though, right at the outset, that because lithium may show high mobility within the structure and it has only three electrons (compared to germanium's 32 and tin's 51), achieving a more accurate description of lithium positions and occupancies in Li–Ge/Sn compounds may be more challenging.

Having said that, we note that in our previous work,7 the crystal structures of Li5Sn2 and Li5Ge2 were re-examined, highlighting the structural intricacies among the heavier carbon congeners. Specifically, it was established that the Li and Sn atoms in Li5−xSn2+x (0 < x < 0.1) exhibit partial disorder.7 This partial substitution of lithium by tin atoms leads to more complex atomic bonding, with the Sn atoms forming larger covalently bonded fragments. In contrast, both experimental and computational evidence indicated that in the structure of the phase hitherto known as Li5Ge2, there are approximately 30% vacancies at one of the lithium sites, with the actual chemical formula being Li5−xGe2 (x ≈ 0.3). The latter composition is very close to that of the “Li7Ge3” phase, which has only been theoretically predicted.19 One should be aware that the work computing Li7Ge3 was carried out using the atom swapping method by combining the ab initio random structure searching procedure (AIRSS) to perform structure prediction from density functional theory (DFT).19 Our experimental work7 cannot quite validate the computationally suggested structure, but it does show that in rhombohedral Li5−xGe2 (x ≈ 0.3), the [Ge2] dimers are slightly more “oxidized” and show bond order intermediate between 1 and 2.

The above-mentioned studies raise questions of potential undetected disorders in other compositionally/structurally similar Li–Ge/Sn phases. Therefore we set out to revisit some of the other structures with smaller, molecular-like fragments, e.g. [Tt2] dimers and [Tt3] trimers (Fig. 1). In particular, our interest was piqued by the reported Li7Tt2 (Tt = Ge, Sn) phases, which are the Li-richest phases whose structures feature homoatomic Tt–Tt bonding; we found the structure of both heptalithium germanide and stannide to be Li-deficient, i.e., Li7−xTt2. As part of our investigation, using modern single-crystal X-ray diffraction methods, we also re-examined the crystal structures of Li7Sn3, and Li13Sn5, which are unique to the Li–Sn phases and have no structural analogs among the known Li–Ge phases; we found both structures to be devoid of any disorder.


image file: d5dt01022e-f1.tif
Fig. 1 Schematic representations of the crystal structures of (a) Li7Sn2, (b) Li13Sn5, and (c) Li7Sn3. Blue represents Li atoms, Sn atoms are drawn in yellow.

Experimental

Synthetic procedures

Handling of the raw and synthesized materials required the use of a glovebox filled with argon gas (maintaining O2/H2O levels below one ppm) and vacuum techniques. This was done in order to protect the air-sensitive reactants and products. Samples were prepared using the traditional solid-state method in sealed tubes. Li and Sn (or Ge) were the starting materials, which were sourced from Sigma–Aldrich and Alfa Aesar with purities of at least 99.9% by weight. Lithium rods were cleaned with a scalpel immediately before being used to remove surface impurities from Li3N. The elements were weighed with molar ratios varying from ca. 4[thin space (1/6-em)]:[thin space (1/6-em)]1 to 2[thin space (1/6-em)]:[thin space (1/6-em)]1. Some of the findings from these experiments have already been reported.7 In this article we focus on the nominal compositions Li[thin space (1/6-em)]:[thin space (1/6-em)]Sn = 7.5[thin space (1/6-em)]:[thin space (1/6-em)]2, 7.0[thin space (1/6-em)]:[thin space (1/6-em)]2, 6.7[thin space (1/6-em)]:[thin space (1/6-em)]2, 6.5[thin space (1/6-em)]:[thin space (1/6-em)]2, 6.25[thin space (1/6-em)]:[thin space (1/6-em)]2, and 6.0[thin space (1/6-em)]:[thin space (1/6-em)]2. In the typical experiments, 300–400 mg elemental mixtures were placed into niobium tubes, which were then welded shut under partial pressure of high purity argon gas. Then, the niobium tubes were enclosed in fused silica tubes, which were evacuated and flame sealed.

The processing was done in programmable tube furnaces, which were brought to 1123 K at a rate of 100 K per hour. The reactions took place over two hours, following which the furnaces were set to controlled cooling at a rate of 25 K per hour. Once at room temperature, the reaction vessels were removed from the furnaces; the fused silica tubes were cut and the welded tubes were returned to the glovebox and carefully opened to retrieve the final products.

Caution! Lithium vapors can leak from the niobium tubes at high temperature and cause the breakage of the fused silica jackets. Therefore, all described experimental procedures have to be performed with utmost care and by using additional precautions to minimize the risk of failure of the silica tubes.

Crystallography

Powder X-ray diffraction. Powder X-ray diffraction (PXRD) experiments were conducted at room temperature on freshly prepared polycrystalline samples. Small portions from each reaction's products were ground in the glovebox into powders, using agate mortars and pestles. The samples were loaded into thin-walled capillary tubes (0.5 mm diameter special glass capillary tubes from Charles Supper). Samples were prepared in the Ar glovebox and the capillary ends were sealed with clay. Then the capillaries were placed in a plastic tube, which was closed and sealed with parafilm, then heat sealed inside a polybag for transport to the diffractometer to minimize air exposure until the measurement (N. B. the measurement was done at ambient conditions). A STOE STADI P diffractometer with monochromatized Ag Kα radiation (λ = 0.55941 Å) with four single module strip detectors (Mythen2 1K Dectris) was utilized and data were collected up to 73.335° in 2θ with a step size of 0.015° for a total scan-time of 30 minutes.
Single-crystal X-ray diffraction. Single-crystal X-ray diffraction (SCXRD) data were gathered for crystals from samples of different nominal compositions. Abbreviated crystallographic parameters for two Li7−xSn2 crystals selected from samples with nominal compositions Li[thin space (1/6-em)]:[thin space (1/6-em)]Sn = 7.0[thin space (1/6-em)]:[thin space (1/6-em)]2 (crystal A), and 6.0[thin space (1/6-em)]:[thin space (1/6-em)]2 (crystal B) are tabulated in Table 1; the corresponding data for three additional Li7−xSn2 crystals are available as SI (Table S1). Also in the SI section, one can find complementary structure refinements for three Li7−xGe2 crystals (Table S2). To ascertain the presented crystallographic work, additional analyses were also performed for Li7Sn3 (from two independent samples), Li13Sn5 (from two independent samples), and Li17Sn4 crystals, with the data available in Tables S3–S5.
Table 1 Crystallographic information for two representative Li7−xSn2 (0.3 < x < 0.5) crystals. T = 200(2) K; Mo Kα, λ = 0.71073 Å; crystal system: orthorhombic, space group: Cmmm, Z = 4
  Crystal A Crystal B
a R1 = ∑||Fo| − |Fc||/∑|Fo|; wR2 = [∑[w(Fo2Fc2)2]/∑[w(Fo2)2]]1/2, where w = 1/[σ2Fo2 + (xP)2 + (yP)], and P = (Fo2 + 2Fc2)/3. x = 0.0126, y = 0.4756 for Li6.67(3)Sn2; x = 0.0229, y = 0.2352 for Li6.5Sn2.
Refined chemical formula Li6.67(3)Sn2 Li6.5Sn2
Formula weight (g mol−1) 283.69 282.49
a (Å) 9.7882(13) 9.8176(13)
b (Å) 13.842(2) 13.826(2)
c (Å) 4.7333(6) 4.7187(6)
V3) 641.29(15) 640.52(15)
ρcalc (g cm−3) 2.94 2.93
μ (cm−1) 76.3 76.4
Goodness-of-fit on F2 1.083 1.056
R1 (I ≥ 2σI)a 0.0220 0.0359
wR2 (I ≥ 2σI)a 0.0395 0.0580
R1 (all data)a 0.0275 0.0558
wR2 (all data)a 0.0408 0.0627
Δρmax,min/e Å−3 1.13/−1.00 1.56/−1.51
CCDC number 2444619 2444622


The general procedure was the following. First, single crystals were selected in an argon-filled glovebox under a microscope. They were emersed in a droplet of dry Paratone-N oil and cut with a scalpel, ensuring their dimensions were ≤0.2 mm. The samples were them taken out of the glovebox and the selected crystals were scooped from the oil drop by MiTeGen plastic loops and transferred to a Bruker APEX III diffractometer with monochromatized Mo Kα radiation (λ = 0.71073 Å). Data collection was performed at a temperature of 200(2) K, which was maintained using a stream of cold nitrogen gas. Raw data sets were processed using the Bruker software and corrected for absorption with the SADABS software package. Crystal structures were solved by the intrinsic phasing method using SHELXT and refined by full-matrix least-squares methods on F2 with SHELXL, with the OLEX2 program as graphical interface.20–22 Atomic coordinates were standardized with Structure TIDY.23 Sn (and Ge) atoms were refined with anisotropic displacement parameters, while the Li atoms in Li7−xTt2 were treated isotropically without further constraints. For the structure refinements of Li7Sn3 and Li13Sn5, both of which are free of any disorder, all or nearly all atoms could be refined with anisotropic displacement parameters. For the structure refinements of Li17Sn4 (Table S5), one of the 13 Li sites had to have its Uiso constrained (vide infra).

The Cambridge Crystallographic Data Centre (CCDC) contains the full supplementary crystallographic data. The respective CIFs have depository numbers 2444618–2444627, 2444630, 2444633 and 2444635.

Electronic structure calculations

The electronic band structures were computed using the Stuttgart TB-LMTO-ASA code, applying the local density approximation (LDA) for the exchange–correlation functional.24 Specifically, the von Barth–Hedin version of the LDA functional was used.25 Additional empty spheres were introduced where necessary to fulfil the atomic sphere approximation (ASA) requirements. The Brillouin zone was sampled with a k-point grid of 6 × 6 × 10. Chemical bonding analysis was carried out using the Crystal Orbital Hamilton Population (COHP) method.26

Total energy calculations were carried out using the Quantum ESPRESSO 7.4 program package27 for Li7Sn2, Li13Sn5, as well as two Li7−xSn2 models: Li13Sn4 (x = 0.5) and Li27Sn8 (x = 0.25). Exchange and correlation were described in the generalized gradient approximation (GGA) with the Perdew–Burke–Ernzerhof exchange–correlation functional (PBE).28 A plane wave basis with a kinetic energy cutoff of 70 Ry and ultrasoft Vanderbilt pseudopotentials were employed.29 The Brillouin zones were sampled with a k-point grid spacing of about 0.15 Å−1. The starting structure of Li6.5Sn2[triple bond, length as m-dash]Li13Sn4 was taken from directly from the single-crystal refinements, where the 2a Li site was found to be completely empty (vide infra). The Li27Sn8 model was generated from the structure of Li13Sn4 by reducing the space group symmetry down to Pmmm without changing the conventional unit cell size and placing a Li atom in 1d or 1g position (both options result in symmetry equivalent structures). All structures were fully relaxed to residual forces of less than 0.01 eV Å−1.

Results and discussion

Synthesis and crystal structure

As indicated in the previous section, several Li7−xSn2 nominal compositions were tried and the corresponding PXRD patterns are shown in Fig. 2. A close look at the PXRD patterns for the nominal compositions Li7Sn2 and Li6.7Sn2 shows that they match very well with the simulated pattern, with only faint additional peaks originating from a minor secondary phase. As the lithium content decreases to the nominal composition Li6.5Sn2, stronger additional peaks (black asterisk) that do not match the reference Li7Sn2 pattern emerge. Upon further reduction in lithium content to nominal composition of 6.25[thin space (1/6-em)]:[thin space (1/6-em)]2, the PXRD pattern still shows peaks consistent with Li7Sn2; however, more intense peaks corresponding to the additional phase, identified as Li13Sn5 become more evident. Finally, the sample with nominal composition Li6Sn2 exhibits distinct diffraction peaks associated with a different phase, Li13Sn5, together with peaks matching the Li7Sn2 phase. The peaks’ relative intensities here are suggestive of Li7Sn2 being the minor product. These findings are in a reasonably good agreement with the established Li–Sn phase diagram,13 which indicates Li7Sn2 and Li13Sn5 to be in equilibrium around 75 at% Li.
image file: d5dt01022e-f2.tif
Fig. 2 Experimental powder X-ray diffraction (PXRD) patterns of samples with varying nominal compositions (Li[thin space (1/6-em)]:[thin space (1/6-em)]Sn = 7[thin space (1/6-em)]:[thin space (1/6-em)]2, 6.7[thin space (1/6-em)]:[thin space (1/6-em)]2, 6.5[thin space (1/6-em)]:[thin space (1/6-em)]2, 6.25[thin space (1/6-em)]:[thin space (1/6-em)]2, and 6[thin space (1/6-em)]:[thin space (1/6-em)]2) compared with the simulated PXRD pattern based on the reported data for Li7Sn2.9 Asterisks denote peaks from side products, which in the cases of Li[thin space (1/6-em)]:[thin space (1/6-em)]Sn = 6.25[thin space (1/6-em)]:[thin space (1/6-em)]2, and Li[thin space (1/6-em)]:[thin space (1/6-em)]Sn = 6[thin space (1/6-em)]:[thin space (1/6-em)]2 have been determined to be Li13Sn5. Data are acquired with monochromatized Ag Kα radiation (λ = 0.55941 Å).

Additionally, we note that upon increasing of lithium content (samples with nominal compositions of 7.25[thin space (1/6-em)]:[thin space (1/6-em)]2 and 7.50[thin space (1/6-em)]:[thin space (1/6-em)]2), the pathway to Li7−xSn2 appears to shut off. Such experiments predominantly yielded the cubic Li17Sn4 phase,10 with indications of another, yet unidentified phase. This may suggest more complex phase relationships than the equilibria near 80 at% Li that are indicated in the phase diagram.13

The PXRD patterns show no systematic changes in the lattice parameters that could point to sample dependence on the experimental conditions, which helps to rule out significant stoichiometry range in Li7−xSn2. This scenario is distinctly different from what was observed for the phase Li5−xSn2+x (0 < x < 0.1), whose structure was recently revisited from Li5Sn2 and found to be partially disordered.7 Nonetheless, from single-crystal X-ray diffraction data, small variations in the lattice parameters were found, even for crystals selected from the same batch. This observation is indicative of small disordering, possibly resulting from partial occupation of Li site(s). By comparison, contrasting behavior can be seen for the other phase that was a common by-product of our syntheses, Li13Sn5—when we examined crystals from two independent preparations, no statistically significant deviations in the lattice parameters were found (Table S3 in SI).

To gain better understanding of the potential structural changes that can be inferred from the discussion above, single-crystal X-ray diffraction analysis was conducted for five crystals from independent Li7−xSn2 reaction batches. The results from the structure refinements confirmed that none of the collected data sets can be matched properly to the reported Li7Sn2 structure.9 Specifically, one of the lithium sites in the published “7–2” structure, the special 2a Wyckoff position (Table 2, Fig. 3), was found to be empty, or rather, not entirely filled by Li. Acknowledging the inaccuracy of refining Li occupancy factors from X-ray diffraction, we draw attention to the fact that in all refined structures, the difference Fourier maps are flat with highest residual density of less than 1.3 e Å−3. In only one case, namely, the crystal from the nominal composition 7[thin space (1/6-em)]:[thin space (1/6-em)]2, a small residual peak was located at coordinates 0,0,0. When refined as partially occupied Li, the corresponding site occupation factor was approximately 35%. If this position is assigned as Li6 at full occupancy, its Uiso parameter becomes 20 times larger than the average Uiso of the other Li atoms in the structure. One may also notice that Li6 has four Li2 neighbors at a distance of approx. 2.5 (2.3 Å in Li7−xGe2, the structure of which was also refined as a part of this study – Tables S2 and S7 in SI) which is very short for a Li–Li contact. At the same time, refining Li2 occupancy does not provide solid evidence for statistically significant deviations from 100%, which indicates that “filling” of the 2a site in the structure by a Li atom may not be favorable.


image file: d5dt01022e-f3.tif
Fig. 3 Structure of Li7−xSn2 (0.3 < x < 0.5). Blue represents Li atoms, Sn atoms are drawn in yellow. Missing/partially occupied Li atom is shown in lighter blue with thin blue lines connecting its closest Li neighbors. The Li-polyhedra around the Sn2 dumbbell and the Sn atoms are also emphasized.
Table 2 Atomic coordinates and equivalent isotropic displacement parameters (Ueq[thin space (1/6-em)]a in Å2) for Li7−xSn2 (0.3 < x < 0.5). The Li atoms are refined isotropically (Ueq = Uiso)
Atom Site x y z Ueq
a Ueq is defined as one third of the trace of the orthogonalized Uij tensor. The anisotropic displacement factor exponent takes the form: −2π2[(ha*)2U11 + … + hka*b*U12].b The position is freely refined as ca. 1/3 occupied without a constraint on Uiso.
Li6.67Sn2
Sn1 4j 0.34610(4) 0 0 0.0107(1)
Sn2 4g 0 0.18646(2) 1/2 0.0108(1)
Li1 8q 0.3116(9) 0.1554(5) 1/2 0.026(2)
Li2 8p 0.149(1) 0.1503(8) 0 0.042(3)
Li3 4h 0.145(1) 0 1/2 0.021(2)
Li4 4i 0 0.3233(8) 0 0.025(3)
Li5 2c 1/2 0 1/2 0.033(4)
Li6 b 2a 0 0 0 0.06(2)
Li6.5Sn2
Sn1 4j 0.34806(7) 0 0 0.0120(2)
Sn2 4g 0 0.18884(5) 1/2 0.0112(2)
Li1 8q 0.311(1) 0.154(1) 1/2 0.033(4)
Li2 8p 0.143(2) 0.145(2) 0 0.062(6)
Li3 4h 0.146(2) 0 1/2 0.037 (6)
Li4 4i 0 0.321(1) 0 0.041(6)
Li5 2c 1/2 0 1/2 0.037(8)


Although the structure of Li17Sn4 is completely different, we note that it has similar issues concerning one of the Li atoms. When the structure of Li17Sn4 was refined from our own single-crystal X-ray diffraction data (Tables S5 and S10 in SI), there was a residual density of 2.8 e Å−3, located at coordinates 0,0,0 and at a distance of approx. 2.5 Å from another Li atom (whose occupancy was checked and verified to be very close to 100%). Refining the residual density as fully occupied Li atom was possible, but resulted in unphysical Uiso parameter. In the final refinement cycles, Uiso for the said Li position, Li13, had to be constrained (N. B. in their combined single-crystal X-ray and powder neutron diffraction work on Li17Sn4, Guloy et al. also used constraints on this very same Li site).10 By comparison, when the structures of Li13Sn5 and Li7Sn3 were refined, there were no Li–Li contacts shorter than 2.75–2.8 Å (Tables S8 and S9 in SI) and we were able to even refine most of the Li atoms with anisotropic displacement parameters.

We also note that there does not seem to be a clear dependence on how the structures are refined with regards to the lithium content in the samples with nominal compositions of 6.7[thin space (1/6-em)]:[thin space (1/6-em)]2, 6.50[thin space (1/6-em)]:[thin space (1/6-em)]2, 6.25[thin space (1/6-em)]:[thin space (1/6-em)]2, and 6.00[thin space (1/6-em)]:[thin space (1/6-em)]2. In all of these cases, the residual density located at coordinates 0,0,0 was less than 0.5–0.6 e Å−3. The highest residual peaks were of about 1–1.5 e Å−3, always less than 1 Å away from other (always different) atoms, whose occupation factors show no evidence for statistically significant deviations from 100%. It is possible though, that some small disorder due to Li mobility in the structure is present, but it remains elusive from the structural data we have at hand. Taken together, all of the above is suggestive of the actual structure being with some Li disorder and with variable composition Li7−xSn2 (0.3 < x < 0.5), not the Li7Sn2 as previously thought.9

Having alluded to some subtle structural features already, let us now look at the whole atomic arrangement in more details. As mentioned earlier, the ideal, disorder-free structure is known as isotypic with Li7Ge2 (space group Cmmm, no. 65; Z = 4; Pearson symbol oS36).30 Within the asymmetric unit, there are eight crystallographically unique atomic positions (Table 2, Fig. 3): two tin atoms located at the 4j and 4g sites and six lithium atoms at the 2a, 2c, 4h, 4i, 8p, and 8q sites. With Li6 at 2a being either partially occupied or missing altogether, the formulation Li7−xSn2 (0.3 < x < 0.5) would be a more accurate representation. This is in line with the reassessment of the Li7Ge2 structure in 1988 as Li6.5Ge2[triple bond, length as m-dash]Li13Ge4, although the exact structural details are not readily available in the literature.15 Our own single-crystal data, provided in Table S2 in the SI section, confirm that the 2a site in the structure is empty and unlikely to host a Li atom due to the unphysically short distances to neighboring atoms, hence the composition Li13Ge4.

The structure of Li7−xSn2 features key structural motifs previously described by Frank et al. (1975).9 Unlike the more lithium-rich phases, such as Li17Sn4[thin space (1/6-em)]10,11 and Li5Sn,12 where all tin atoms are isolated, Li7−xSn2 contains both isolated tin atoms and tin dumbbells (Fig. 3). The refined bond lengths within the Sn1–Sn1 dumbbells vary between 2.9838(13) Å (Li6.5Sn2) and 3.0131(7) Å (Li6.7Sn2). This is additional experimental evidence lending support to the formulation Li7−xSn2: since Sn–Sn distances vary a lot more than 3–4σ, one may reason that slightly varied Li concentration is a contributing factor. Furthermore, one should also notice that although the refinements of Li occupancies are not conclusive, the refinements of Sn–Sn distances show the structure where Sn is formally more reduced (Li6.67Sn2) to exhibit longer Sn1–Sn1 distance. Such variations of Ge–Ge distances are not observed in the 3 independent refinements of Li7−xGe2 (bond lengths in the interval 2.607(1) to 2.615(2) Å, Table S7). This indicates that the germanide is more akin to a line phase Li6.5Ge2[triple bond, length as m-dash]Li13Ge4, confirming the earlier suggested reformulation of Li7Ge2 as Li13Ge4.16

Along these lines, considering other Li–Sn phases with high Li contents and structures featuring smaller molecular-like fragment is also instructive. For example, Sn–Sn bond lengths within the dumbbells in Li5−xSn2+x (0 < x < 0.1) are approx. 2.86 Å and vary from sample to sample.7 For Li13Sn5, variations of Sn–Sn distances are not observed (2.865(1) and 2.8653(6) Å); Sn–Sn distances within the bent [Sn3] moiety in Li7Sn3 are also nearly invariant of the sample (3–5σ deviations with an average value of approx. 2.94 Å [Table S9]). With the Sn atom formally being in its most reduced state in Li6.67Sn2, it is reasonable to expect the Sn–Sn bonding to be stretched farthest. As discussed in the next chapter, based on simple partitioning of the valence electrons following the Zintl–Klemm concept31,32 and based on the electronic structure calculations, it is reasonable to approximate the [Sn–Sn]6− fragment in Li7−xSn2 as isosteric with the I2 molecule (i.e., single covalent bond). The less reduced Sn atoms in Li5−xSn2+x (0 < x < 0.1) and Li13Sn5 ought to share more valence electrons, meaning that the [Sn–Sn]4.5− fragments in those will be more like the super-oxo [O2] species (i.e., intermediate between single and double covalent bond). By analogy, the bent [Sn3]7− in Li7Sn3, the least reduced stannide from the list, will have to be considered as isosteric with the allyl anion, [H2C–CH[double bond, length as m-dash]CH2].

The isolated tin atoms and the tin dumbbells are well separated, with the shortest contacts between Sn1 and Sn2 on the order of 4.9 Å; Sn2 and Sn2 are more than 4.7 Å apart. Sn atoms have 12 nearest Li atoms neighbors (not counting the partially occupied Li) (Fig. 3). The dumbbells and isolated tin atoms are distinctly arranged in a layered manner, with each type located in separate ab-planes. The Sn–Li distances range from 2.79 to 3.23 Å, and Li–Li distances also vary significantly, ranging from as short as 2.5 Å to over 3.5 Å (SI, Table S6).

Electronic structure and chemical bonding

First, let us try to rationalize the chemical bonding from the standpoint of the Zintl–Klemm concept.31,32 Following the discussion from the preceding section, the valence electrons in Li7Sn2[triple bond, length as m-dash]Li14Sn4 and Li13Sn4 can be partitioned in a manner consistent with the tin dumbbells being [Sn–Sn]6− i.e., single-bonded and isosteric with the I2 molecule. We can then express their formulae in the following way: Li14Sn4 = (Li+)14[Sn2]6−(Sn4−)2 and Li13Sn4 = (Li+)13[Sn2]6−(Sn4−)2(h+), where h+ denotes an electron–hole. Similarly, the formula of Li7Sn3 can be written as (Li+)7[Sn3]7− (recognizing the analogy with the allyl anion) or as (Li+)7[Sn3]8−(h+) (treating both Sn–Sn bonds as 2-center 2-electron interactions). Li13Sn5 with the intermediate Sn–Sn bonding can be expressed as (Li+)13([Sn2]5−)2(Sn4−)(h+) or (Li+)13([Sn2]4.5−)2(Sn4−) when further “oxidation” of the dumbbell is allowed. Clearly, simple bonding theories do not capture the details of the chemical bonding in these compounds, as the only formula that can be readily charge-balanced is that of Li14Sn4.

Electronic structure calculations were carried out for two structural models: Li7Sn2[triple bond, length as m-dash]Li14Sn4, where the Li position at Wyckoff site 2a is completely filled, and Li6.5Sn2[triple bond, length as m-dash]Li13Sn4, where the said position is empty. The total and projected densities of states for Li14Sn4 and Li13Sn4 are shown in Fig. 4 and 5. Electronic structure calculations for Li13Sn5 and Li7Sn3 are also done and are shown as figures in the SI.


image file: d5dt01022e-f4.tif
Fig. 4 Calculated total electronic density of states (DOS) and partial densities of states (pDOS) for Li7Sn2 (top). Crystal orbital Hamilton population (COHP) curves for averaged Sn–Sn, Sn–Li, and Li–Li interactions in Li14Sn4 (bottom). The dashed lines correspond to the Integrated Density of State (IDOS). The vertical magenta line to the left of the Fermi level shows the location of a pseudo-gap corresponding to ca. 29.5 valence electrons per unit cell (Li∼6.75Sn2).

image file: d5dt01022e-f5.tif
Fig. 5 Calculated total electronic density of states (DOS) and partial densities of states (pDOS) for Li13Sn4 (top). Crystal orbital Hamilton population (COHP) curves for averaged Sn–Sn, Sn–Li, and Li–Li interactions in Li13Sn4 (bottom). The dashed lines correspond to the Integrated Density of State (IDOS). The vertical magenta line to the right of the Fermi level shows the location of a pseudo-gap corresponding to ca. 29.5 valence electrons per unit cell (Li∼13.5Sn4).

The difficulties with balancing charges are reflected by the plots of the total and projected densities of states. From the calculated position of the Fermi level (EF), it is apparent that both computed model structures do not represent valence-precise compounds as there are no noticeable (pseudo)gaps at EF. Notice that the Li14Sn4 and Li13Sn4 differ by just one electron and the inclusion of the “extra” lithium atom in the calculation has a very small effect on the overall DOS. Lack of gaps in the electronic band structures is also seen in the cases of the other lithium stannides, Li7Sn3 and Li13Sn5 that were re-evaluated in this work (SI Fig. S1–S4). Thus, all stannides discussed above will be expected to be compounds with metallic properties.

However, very close to the Fermi level, pseudogaps are apparent. Notably, within the rigid band approximation, the Fermi level can be shifted to the local minimum when the electron count is set at ca. 29.5 e. This value is intermediate between Li13Sn4 (29 e) and Li14Sn4 (30 e). This observation is an indirect evidence of support of the proposed off-stoichiometry of the Li7−xSn2 phase as neither end member is expected to have greater electronic stability than the phase with partial occupancy of the Li 2a site. As discussed earlier, the lack of refineable electron density at that position does not rule out small possible disordering elsewhere in the structure that would lead to an upended Li content in alignment with this electronic structure analysis. Consequently, we argue that both experimental and computation evidence support the re-assessment of Li7Sn2 as Li7−xSn2 where x will always be less than 0.5.

The corresponding Crystal Orbital Hamilton Population (COHP) curves for selected interactions in Li14Sn4 and Li13Sn4 are very much alike. The strongest bonds, based on the integrated COHP per bond, are the Sn–Sn bonds. However, these bonds are not fully optimized at the Fermi level due to the presence of some antibonding states. In fact, the Sn–Sn COHP at Fermi level in Li14Sn4 has more pronounced antibonding character compared to Li13Sn4, where the COHP is almost nonbonding. Since the Sn–Sn interactions in Li13Sn4 display a lower occupation of antibonding states, one would expect stronger Sn–Sn bonds, which is supported by the experimentally observed shorter Sn–Sn bond lengths in Li6.5Sn2[triple bond, length as m-dash]Li13Sn4 as compared to those in Li6.67Sn2[triple bond, length as m-dash]Li13.3Sn4.

In contrast, the Li–Sn and Li–Li interactions are effectively optimized at EF. It appears that as the difference of a single Li atom alters the coordination environments and orbital overlap significantly enough to strengthen the Sn–Sn interactions and provide the “drive” for the phase to move away for the ideal “7–2” composition.

To compare relative stabilities of binary Li stannides in the vicinity of the Li7−xSn2 composition, we first computed total energies of Li7Sn2 and Li13Sn5. The latter composition has been reported to be in equilibrium with Li7Sn2 at the Li/Sn ratios between 2.6 and 3.5.33 After that, we evaluated the total energies of two ordered Li7−xSn2 models: Li13Sn4 and Li27Sn8. Comparison of the obtained values suggests that Li13Sn4 is about 3.4 meV per atom less stable, while Li27Sn8 is 0.8 meV per atom more stable that the corresponding mixtures of Li7Sn2 and Li13Sn5. In other words, Li27Sn8 is predicted to be on the convex hull in the Li–Sn system. Of course, the limited accuracy of the PBE calculations for estimation of the decomposition enthalpies should not be ignored,34 but in general, our calculations support the experimental observation that Li7−xSn2 can be stabilized for a range of values x > 0. Furthermore, the composition of the idealized Li27Sn8[triple bond, length as m-dash]Li13.5Sn4 corresponds to 29.5 valence electrons per formula unit. The predicted thermodynamic stability of this phase is in line with the optimized chemical bonding at that electron count suggested by the LMTO calculations.

Conclusions

Li7Sn2 has been known for nearly five decades and is now shown to exhibit notable lithium site defects. The present study reveals that one specific Li site—namely, the 2a Wyckoff position is either partially or completely unoccupied. This observation leads to a revised composition of Li7−xSn2, with x ranging between 0.3 and 0.5. Structural analysis across multiple samples shows no clear correlation between nominal Li[thin space (1/6-em)]:[thin space (1/6-em)]Sn ratio and occupational disorder, however Sn–Sn interactions do indicate structural response scaling with Li content. Comparative analysis with related stannides (Li7Sn3 and Li13Sn5) emphasizes consistent trends between bond lengths and bonding strength.

Li7−xSn2, as well as other Li-rich germanides and stannides show metallic characteristics, making it difficult for the valence rules and the Zintl concept to sufficiently account for the observed vacancies at Li sites in these compounds. These limitations underscore the broader challenges in accurately resolving Li positions and occupancies through crystallography and structural analysis.

To address these challenges and gain deeper insights into the site-specific behavior of alkali elements in tetrelides systems, investigation of analogous Na-containing compounds is in order. Given sodium's higher electron counts and expected chemical similarity to Li, Na-based systems may offer improved structural resolution and contribute to a more comprehensive understanding of alkali–tetrel interactions. Furthermore, these studies will also provide valuable knowledge applicable to the development of Na-ion battery technologies.

Author contributions

Salina Rahman: conceptualization, investigation, methodology, formal analysis, visualization, writing – original draft. Kowsik Ghosh: conceptualization, investigation, methodology, formal analysis, visualization, writing – original draft. Alexander Ovchinnikov: electronic structure calculations, formal analysis, chemical bonding analysis, writing – review & editing. Anirudh Nandakumar: investigation, formal analysis, writing – review & editing. Candace K. Chan: formal analysis, project administration, writing – review & editing. Svilen Bobev: conceptualization, supervision, project administration, writing – review & editing.

Conflicts of interest

There are no conflicts to declare.

Data availability

The data supporting this article have been included as part of the SI.

Tables with additional crystallographic data for all refined structures; DOS and COHP plots for Li13Sn5 and Li7Sn3. See DOI: https://doi.org/10.1039/d5dt01022e

CCDC 2444618–2444627, 2444630, 2444633 and 2444635 contain the supplementary crystallographic data for this paper.35a–m

Acknowledgements

This work was supported by the National Science Foundation through collaborative grants DMR-2004579 and DMR-2418368 (UD) and DMR-2004514 and DMR-2418367 (ASU). We acknowledge the use of facilities within the Eyring Materials Center at Arizona State University and assistance with PXRD measurements from R. Zhang, supported in part by NNCI-ECCS-1542160. NSF Major Research Instrumentation Award DMR-2217231: “MRI: Acquisition of a Dual Transmission X-ray Diffractometer (DTXRD) for Studying the Local and Bulk Structure of Soft and Hard Materials under In situ and Operando Conditions” is acknowledged for acquisition of the STOE Stadi P X-ray diffractometer.

References

  1. C.-M. Park, J.-H. Kim, H. Kim and H.-J. Sohn, Li-alloy based anode materials for Li secondary batteries, Chem. Soc. Rev., 2010, 39, 3115–3141 RSC.
  2. C.-Y. Chou, H. Kim and G. S. Hwang, A comparative first-principles study of the structure, energetics, and properties of Li–M (M = Si, Ge, Sn) alloys, J. Phys. Chem. C, 2011, 115, 20018–20026 CrossRef CAS.
  3. D. Hansen and L. Chang, Crystal structure of Li2Sn5, Struct. Sci., 1969, 25, 2392–2395 CAS.
  4. W. Müller and H. Schäfer, Die kristallstruktur der phase LiSn: the crystal structure of LiSn, Z. Naturforsch., B, 1973, 28, 246–248 CrossRef.
  5. W. Müller, Darstellung und Struktur der Phase Li7Sn3, Z. Naturforsch., B, 1974, 29, 304–307 CrossRef.
  6. U. Frank, W. Müller and H. Schäfer, Die Struktur der Phase Li5Sn2, Z. Naturforsch., B, 1975, 30, 1–5 CrossRef.
  7. K. Ghosh, S. Rahman, A. Ovchinnikov and S. Bobev, Applicability of the Zintl Concept to Understanding the Crystal Chemistry of Lithium-Rich Germanides and Stannides, Inorg. Chem., 2024, 63, 20173–20185 CrossRef CAS PubMed.
  8. U. Frank and W. Müller, Darstellung und Struktur der Phase Li13Sn5 und die Strukturelle Verwandtschaft der Phasen in den Systemen Li–Sn und Li–Pb, Z. Naturforsch., B, 1975, 30, 316–322 CrossRef.
  9. U. Frank, W. Müller and H. Schäfer, Die Kristallstruktur der Phase Li7Sn2, Z. Naturforsch., B, 1975, 30, 6–9 CrossRef.
  10. C. Lupu, J.-G. Mao, J. W. Rabalais, A. M. Guloy and J. W. Richardson, X-ray and neutron diffraction studies on “Li4.4Sn”, Inorg. Chem., 2003, 42, 3765–3771 CrossRef CAS PubMed.
  11. G. Goward, N. Taylor, D. Souza and L. Nazar, The true crystal structure of Li17M4 (M = Ge, Sn, Pb) – revised from Li22M5, J. Alloys Compd., 2001, 329, 82–91 CrossRef CAS.
  12. R. U. Stelzer, Y. Ikeda, P. Srinivasan, T. S. Lehmann, B. Grabowski and R. Niewa, Li5Sn, the most lithium-rich binary stannide: a combined experimental and computational study, J. Am. Chem. Soc., 2022, 144, 7096–7110 CrossRef CAS PubMed.
  13. J. Sangster and C. W. Bale, The Li-Sn (lithium-tin) system, J. Phase Equilib., 1998, 19, 70–75 CrossRef CAS.
  14. M. Mayo, J. P. Darby, M. L. Evans, J. R. Nelson and A. J. Morris, Correction to structure prediction of Li–Sn and Li–Sb intermetallics for lithium-ion batteries anodes, Chem. Mater., 2018, 30, 5516–5517 CrossRef CAS.
  15. V. Hopf, W. Müller and H. Schäfer, Die Struktur der Phase Li7Ge2, Z. Naturforsch., B, 1972, 26, 1157–1160 CrossRef.
  16. R. Nesper, Synthese, Struktur und Eigenschaften von Verbindungen im Grenzbereich zwischen Zintl-Phasen und Metallen: experimentelle und quantenmechanischen Beiträge zum Verständnis der chemischen Bindung im Grenzbereich zwischen Halbleitern und Metallen und topologische Konzepte zur Struktursytematik. Habilitation Thesis, Universität Stuttgart, 1988 Search PubMed.
  17. U. Frank, W. Müller and H. Schäfer, Zur Kenntnis der Phase Li13Si4, Z. Naturforsch., B, 1975, 30, 10–13 CrossRef.
  18. M. Zeilinger and T. F. Fässler, Revision of the Li13Si4 structure, Acta Crystallogr., Sect. E:Crystallogr. Commun., 2013, 69, i81–i82 CrossRef CAS PubMed.
  19. A. J. Morris, C. P. Grey and C. J. Packard, Thermodynamically stable lithium silicides and germanides from density functional theory calculations, Phys. Rev. B:Condens. Matter Mater. Phys., 2014, 90, 054111 CrossRef CAS.
  20. O. V. Dolomanov, L. J. Bourhis, R. J. Gildea, J. A. K. Howard and H. Puschmann, OLEX2: a complete structure solution, refinement and analysis program, J. Appl. Crystallogr., 2009, 42, 339–341 CrossRef CAS.
  21. G. M. Sheldrick, Crystal structure refinement with SHELXL, Acta Crystallogr., Sect. C:Struct. Chem., 2015, 71, 3–8 Search PubMed.
  22. G. M. Sheldrick, SHELXT – Integrated space-group and crystal-structure determination, Acta Crystallogr., Sect. A:Found. Adv., 2015, 71, 3–8 CrossRef PubMed.
  23. L. Gelato and E. Parthé, STRUCTURE TIDY – a computer program to standardize crystal structure data, J. Appl. Crystallogr., 1987, 20, 139–143 CrossRef.
  24. O. K. Andersen and O. Jepsen, Explicit, first-principles tight-binding theory, Phys. Rev. Lett., 1984, 53, 2571 CrossRef CAS.
  25. U. von Barth and L. Hedin, A local exchange-correlation potential for the spin polarized case. i, J. Phys. C Suppl., 1972, 5, 1629 CrossRef CAS.
  26. S. Steinberg and R. Dronskowski, The crystal orbital Hamilton population (COHP) method as a tool to visualize and analyze chemical bonding in intermetallic compounds, Crystals, 2018, 8, 225 CrossRef.
  27. P. Giannozzi, O. Andreussi, T. Brumme, O. Bunau, M. Buongiorno Nardelli, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, M. Cococcioni, N. Colonna, I. Carnimeo, A. Dal Corso, S. de Gironcoli, P. Delugas, R. A. DiStasio Jr, A. Ferretti, A. Floris, G. Fratesi, G. Fugallo, R. Gebauer, U. Gerstmann, F. Giustino, T. Gorni, J. Jia, M. Kawamura, H.-Y. Ko, A. Kokalj, E. Küçükbenli, M. Lazzeri, M. Marsili, N. Marzari, F. Mauri, N. L. Nguyen, H.-V. Nguyen, A. Otero-de-la-Roza, L. Paulatto, S. Poncé, D. Rocca, R. Sabatini, B. Santra, M. Schlipf, A. P. Seitsonen, A. Smogunov, I. Timrov, T. Thonhauser, P. Umari, N. Vast, X. Wu and S. Baroni, J. Phys.: Condens. Matter, 2017, 29, 465901 CrossRef CAS PubMed.
  28. J. P. Perdew, K. Burke and M. Ernzerhof, Generalized Gradient Approximation Made Simple, Phys. Rev. Lett., 1996, 77, 3865 CrossRef CAS PubMed.
  29. D. Vanderbilt, Soft self-consistent pseudopotentials in a generalized eigenvalue formalism, Phys. Rev. B:Condens. Matter Mater. Phys., 1990, 41, 7892 CrossRef PubMed.
  30. P. Villars and L. D. Calvert, Pearson's handbook of crystallographic data for intermetallic phases, American Society for Metals, Metals Park, 1986 Search PubMed.
  31. S. M. Kauzlarich, Chemistry, structure, and bonding of Zintl phases and ions, VCH, New York, 1998 Search PubMed.
  32. R. Nesper, The Zintl–Klemm concept – a historical survey, Z. Anorg. Allg. Chem., 2014, 640, 2639–2648 CrossRef CAS.
  33. S. Fürtauer and H. Fladorfer, The Cu-Li-Sn phase diagram: Isothermal sections, J. Alloys. Compd., 2016, 682, 713 CrossRef.
  34. G. Hautier, S. P. Ong, A. Jain, C. J. Moore and G. Ceder, Accuracy of density functional theory in predicting formation energies of ternary oxides from binary oxides and its implication on phase stability, Phys. Rev. B: Condens. Matter Mater. Phys., 2012, 85, 155208 CrossRef.
  35. (a) S. Rahman, K. Ghosh, A. Ovchinnikov, A. Nandakumar, C. K. Chan and S. Bobev, CCDC 2444618: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2n1tn6; (b) S. Rahman, K. Ghosh, A. Ovchinnikov, A. Nandakumar, C. K. Chan and S. Bobev, CCDC 2444619: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2n1tp7; (c) S. Rahman, K. Ghosh, A. Ovchinnikov, A. Nandakumar, C. K. Chan and S. Bobev, CCDC 2444620: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2n1tq8; (d) S. Rahman, K. Ghosh, A. Ovchinnikov, A. Nandakumar, C. K. Chan and S. Bobev, CCDC 2444621: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2n1tr9; (e) S. Rahman, K. Ghosh, A. Ovchinnikov, A. Nandakumar, C. K. Chan and S. Bobev, CCDC 2444622: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2n1tsb; (f) S. Rahman, K. Ghosh, A. Ovchinnikov, A. Nandakumar, C. K. Chan and S. Bobev, CCDC 2444623: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2n1ttc; (g) S. Rahman, K. Ghosh, A. Ovchinnikov, A. Nandakumar, C. K. Chan and S. Bobev, CCDC 2444624: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2n1tvd; (h) S. Rahman, K. Ghosh, A. Ovchinnikov, A. Nandakumar, C. K. Chan and S. Bobev, CCDC 2444625: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2n1twf; (i) S. Rahman, K. Ghosh, A. Ovchinnikov, A. Nandakumar, C. K. Chan and S. Bobev, CCDC 2444626: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2n1txg; (j) S. Rahman, K. Ghosh, A. Ovchinnikov, A. Nandakumar, C. K. Chan and S. Bobev, CCDC 2444627: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2n1tyh; (k) S. Rahman, K. Ghosh, A. Ovchinnikov, A. Nandakumar, C. K. Chan and S. Bobev, CCDC 2444630: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2n1v1m; (l) S. Rahman, K. Ghosh, A. Ovchinnikov, A. Nandakumar, C. K. Chan and S. Bobev, CCDC 2444633: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2n1v4q; (m) S. Rahman, K. Ghosh, A. Ovchinnikov, A. Nandakumar, C. K. Chan and S. Bobev, CCDC 2444635: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2n1v6s.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.