Sven Neuberger,
Neeshma Mathew,
Sheyi Clement Adediwura
,
Hector Javier Cortes Sanchez and
Jörn Schmedt auf der Günne
*
University of Siegen, Faculty IV: School of Science and Technology, Department for Chemistry and Biology, Inorganic Materials Chemistry and Center of Micro- and Nanochemistry and (Bio)Technology (Cμ), Adolf-Reichwein Straße 2, 57076 Siegen, Germany. E-mail: gunnej@chemie.uni-siegen.de
First published on 10th July 2025
Non-oxide chalcogenides are among the fastest Li–ion conductors currently available. Here, a new crystalline lithium selenido-phosphate, Li4P2Se6, is reported. Its structure was solved using a combined approach of quantum-chemical structure prediction, powder X–ray diffraction, and solid-state NMR. Li4P2Se6 crystallizes in an orthorhombic unit cell in the space group Pnma and the lattice parameters a = 13.8707(1) Å, b = 11.2115(1) Å, and c = 6.45445(7) Å, representing a novel structure type. 31P and 77Se magic–angle–spinning NMR spectra are reported and were analyzed for chemical shift and J-couplings. The material's ionic conductivity, characterized by impedance spectroscopy, slightly exceeds that of the non-isostructural crystalline compound Li4P2S6, which contains the homologous complex anion. A computed phase diagram shows that several compounds can be prepared experimentally, which from the computed energy at a temperature of 0 K, feature a positive energy above the convex hull and thus should not exist. This research highlights the influence of configurational and vibrational entropy in stabilizing ionic chalcogenides, emphasizes the complexity of predicting phase stability in related systems by quantum-chemical calculations and contributes to the understanding of non-oxide chalcogenides, including potential fast ionic conductors.
Only a single ternary crystalline phase in the system Li–P–Se is currently known, corresponding to the high–temperature modification of Li7PSe6.2 This compound crystallizes in the well-known cubic argyrodite structure and exhibits fast Li-ion conductivity, similar to its sulfur-based analogue.3 Efforts to enhance the ionic conductivity of this phase through aliovalent substitution with different elements have been reported.4,5 Additionally, a previous study provided strong evidence for another crystalline phase, “Li4P2Se6”, based on solid-state NMR and powder X-ray diffraction data, including a suggested unit cell; however, its structure remained unsolved.6 A key question to be addressed is whether the crystalline “Li4P2Se6” adopts the same structure as its sulfur analogue Li4P2S6.7,8
A previous study9 on aliovalent doping of Li4P2S6 with magnesium demonstrated that several energetic minima can be accessed by only minor compositional changes, suggesting that the structure of Li4P2Se6 could well differ from the sulfur variant. Furthermore, it has been theoretically proposed that “Li3PSe4” may exist in the system, but this has yet to be confirmed experimentally.10
The energetic hypersurface of existing structures can be explored through a combination of quantum chemical energy calculations and search algorithms like Monte Carlo or genetic algorithm.11–13 Such structure prediction techniques provide insights into the stability of different crystalline phases by probing the convex hull of all known and hypothesized crystalline structures. Moreover, they allow identification of competing structures for a given composition and estimation of their relative formation energies.
The purpose of this study is to determine the structure and properties of Li4P2Se6 and to relate them to other crystalline phases within the Li–P–Se system.
Li4P2Se6 was synthesized according to eqn (1). Red phosphorus (4 mmol, 123.9 mg, Alfa Aesar, 99.999%), selenium powder (10 mmol, 790.0 mg, ChemPur, 99.999%) and lithium selenide (4 mmol, 371.0 mg) were thoroughly ground and mixed in an agate mortar and afterwards filled into a graphitized quartz ampule (8 mm outer diameter). The ampule was sealed under vacuum (p = 1.8 × 10−2 mbar) and heated to 650 °C in a tube furnace for seven days.
![]() | (1) |
Chemical shift values are reported on a deshielding scale. The 1H resonance of 1% Si(CH3)4 in CDCl3 served as an external secondary reference using the Ξ values for 31P and 77Se as reported by the IUPAC.14
Magic angle spinning (MAS) was performed using a commercial 3.2 mm triple resonance Bruker probehead within zirconia rotors (3.2 mm, ZrO2). Rotors were packed under an argon atmosphere in a glove box. All spectra were recorded at a sample spinning frequency of νrot = 20 kHz, except for the 77Se and 31P MAS spectra used for the determination of chemical shift tensors, which were recorded at νrot = 15 and 5 kHz, respectively.
For the 31P and 77Se MAS NMR measurements, repetition delays of 600 s and 820 s were used respectively.
For the homonuclear 31P–31P 2D zero-quantum (ZQ) exchange spectroscopy measurement, the super-cycled R662 sequence15 was used with a repetition delay of 32 s, 72 R-elements, an increment of 50 μs and rotor-synchronized data sampling of the indirect dimension accumulating 8 transients per FID. The R-element consisted of a 90° pulse and a 270° pulse.
Rietveld refinements were carried out using TOPAS academic V7.16 The structure model with the space group Pnma (no. 62) was used as the starting model. In the given order, the following parameters were refined: scale factor and background coefficients using a Chebyshev function with 12 free parameters, the peak shape using the fundamental parameter approach, the zero-shift error, the lattice constants, the atom positions and the isotropic atomic displacement parameters. Fit indicators of Rwp, Rexp, and GOF were used to assess the quality of the refined structural models.17
Structure prediction was done with USPEX version 9.4.4, including predefined molecular building elements.23–26 The formula Li4P2Se6 was decomposed into Li and P2Se6 units. Calculations with different numbers of formula units Z ∈ {1, 2, 3, 4} and different population sizes N ∈ {20, 30, 50, 100} were started individually assuming a temperature of 0 K. For each individual structure suggestion, three consecutive calculations were carried out with increasingly stricter energy convergence criteria (0.001, 0.0001 and 0.00001 a.u.) and a finer k-point density (0.14, 0.10 and 0.04 2π per Å), respectively. Quantum chemical calculations within USPEX were carried out with the plane-wave self-consistent field package from Quantum ESPRESSO (version 6.7)18 using Gaussian smearing and a degauss parameter of 0.02 Ry. Calculations used PBE-pseudopotentials from the GBRV library.19
Electrochemical impedance spectroscopy (EIS) measurements were recorded using a NEISYS electrochemical impedance analyzer (Novocontrol Technologies, Montabaur, Germany) in a home-built cell, which was calibrated before actual measurements based on short/load calibration standards with a 100 Ω resistor as the load. The impedance measurements were recorded in potentiostatic mode, with an amplitude of 7.1 mVrms, in a frequency range from 50 mHz to 1 MHz. The temperature was controlled using a variable temperature and flow controller (NMR Service GmbH) with a constant nitrogen gas flow. Each temperature point was held for 20 min to ensure thermal equilibrium with an accuracy of ±0.1 K throughout the EIS measurement. The data analysis was performed using a home-written Python script.27
The 31P chemical shift tensors were assumed to be collinear with the internuclear vector of the dipolar interaction. This assumption is based on the idealized D3d molecular geometry of [P2Se6]4−, in which the principal C3 axis is aligned along the P–P bond. According to pseudosymmetry arguments, the presence of a Cn axis (with n ≥ 3) suggests that the principal axis of the chemical shift tensor should be parallel to the rotation axis. Although the molecular symmetry in the crystal structure deviates from the idealized case, a significant misalignment between the principal axis of the chemical shift tensor and the dipolar vector is not expected.
The results are sensitive to the sign of the J-coupling, as seen in the lineshape and relative intensities of the spinning sidebands. Simulations suggest that a negative 1J(31P,31P) coupling provides a slightly better match to the experimental spectrum, consistent with the literature values for similar systems.28 A comparison of simulations using positive and negative J-coupling values is provided in Fig. S1.† The fitted parameters are summarized in Table 1.
31P NMR | 77Se NMR | |||||
---|---|---|---|---|---|---|
Spin 1 | Spin 2 | Spin 1 | Spin 2 | Spin 3 | Spin 4 | |
δiso/ppm | 46.4 | 55.2 | 83 | 103 | 166 | 105 |
δaniso/ppm | −109 | –109 | −436 | −446 | 557 | 478 |
η | 0.3 | 0 | 0.7 | 1.0 | 0.7 | 1.0 |
δ11/ppm | 115 | 110 | 455 | 550 | 724 | 585 |
δ22/ppm | 85 | 110 | 149 | 104 | 80 | 100 |
δ33/ppm | −62 | –54 | −352 | −342 | −305 | −367 |
1Jiso(31P,77Se) coupling/Hz | — | — | 488 | 532 | 531 | 496 |
1Jiso(31P,31P) coupling/Hz | −227 | −227 | — | — | — | — |
The homonuclear 31P–31P 2D ZQ exchange MAS NMR spectrum of Li4P2Se6 (Fig. 3) exhibits two cross peaks for the 31P doublets, consistent with a J-coupled two-spin system originating from the same crystalline phase. This indicates that the crystal structure of Li4P2Se6 must contain at least two crystallographic orbits for P with equal multiplicities. The presence of homonuclear J-coupling indicates a P–P bond, suggesting the presence of a [P2Se6]4− unit featuring two different P sites. These findings confirm the results of the previous study.6
The 77Se MAS NMR spectrum reveals four 77Se resonances at δiso = 83.8, 103.8, 105.8 and 166.0 ppm, with intensity ratios of 1:
1.1
:
1.9
:
1.8 (Fig. 4). These observations provide valuable constraints for the minimum number and multiplicities of the Se sites in the crystal structure of Li4P2Se6. The chemical shift parameters for all signals of the experimental 77Se MAS NMR spectrum are listed in Table 1.
The algorithm was provided with the composition Li4P2Se6 and the number of formula units Z ∈ {1, 2, 3, 4} as the input. During the search, the candidate structures proposed by the algorithm were optimized using density functional theory (DFT) under periodic boundary conditions. The search space was drastically reduced by using the fragment [P2Se6] and Li as the basic building elements.
The X-ray diffraction patterns for the 20 best structures generated for each value of Z were obtained and visually compared to the experimental powder X-ray diffraction pattern. Among these, only the orthorhombic structure described below provided a good match with the experimental pattern. The structure that matches the experimental data is among the three structures, which are lowest in energy (Table 4). It is less than 2 meV per atom higher in energy than the most stable structure.
Additionally, the structural model was required to reflect the number of sites observed by solid-state NMR; at least two and four sites should be present for P and Se, respectively. The proposed structural model satisfied this condition, with two P-sites both with Wyckoff labels 4c and four Se-sites with Wyckoff labels 8d, 8d, 4c and 4c, consistent with the experimental NMR data.
The main reflections from the powder XRD pattern and the 31P NMR peaks reported in a previous study are in agreement with the proposed structure in this work.6 This indicates that the presented structure corresponds to the same phase, although the previously suggested unit cell was different.
The structural parameters were refined using the Rietveld method (Fig. 5). The refined parameters are listed in Table 2. The refined structure is novel and has not been classified previously in the ICSD database. It consists of hexaselenidohypodiphosphate units, with Li atoms occupying octahedrally coordinated positions (Fig. 6). The bond lengths and angles are within typical ranges (Table 3).
![]() | ||
Fig. 6 The structure of Li4P2Se6. Lithium (gray), phosphorus (black) and selenium (red) positions are shown. The octahedral Li polyhedra are shown in purple shades. |
Predicted structural model | Refined structural model | |||||
---|---|---|---|---|---|---|
a/Å | 14.081 | 13.8707(1) | ||||
b/Å | 11.256 | 11.2115(1) | ||||
c/Å | 6.494 | 6.45445(7) | ||||
Crystal system | Orthorhombic | Orthorhombic | ||||
Space group | Pnma (No. 62) | Pnma (No. 62) | ||||
Z | 4 | 4 | ||||
Atom | Wyckoff site | x | y | z | Occ. | Beq/Å2 |
Li1 | 8d | 0.625(2) | 0.397(2) | 0.0932(49) | 1 | 1 |
Li2 | 8d | 0.127(2) | 0.893(2) | 0.945(4) | 1 | 1 |
P1 | 4c | 0.7867(4) | 3/4 | 0.0912(12) | 1 | 0.72(9) |
P2 | 4c | 0.9401(4) | 3/4 | 0.1101(10) | 1 | 0.72(9) |
Se1 | 8d | 0.9829(1) | 0.9137(2) | 0.2716(3) | 1 | 0.29(2) |
Se2 | 8d | 0.7490(1) | 0.5862(4) | 0.9243(4) | 1 | 0.29(2) |
Se3 | 4c | 0.9957(2) | 3/4 | 0.7882(4) | 1 | 0.29(2) |
Se4 | 4c | 0.7387(2) | 3/4 | 0.4197(7) | 1 | 0.29(2) |
Atom 1 | Atom 2 | Atom3 | Bond length/Å | Bond angle/° |
---|---|---|---|---|
Se1 | P2 | — | 2.193(4) | — |
Se2 | P1 | — | 2.192(6) | — |
Se3 | P2 | — | 2.216(7) | — |
Se4 | P1 | — | 2.222(9) | — |
P1 | P2 | — | 2.131(8) | — |
Se4 | P1 | P2 | — | 104.2(4) |
P2 | P1 | Se2 | — | 105.4(2) |
Se1 | P2 | P1 | — | 107.3(2) |
P1 | P2 | Se3 | — | 107.1(3) |
Se4 | P1 | Se2 | — | 113.4(2) |
P2 | P1 | Se2 | — | 105.4(2) |
Se1 | P2 | Se3 | — | 110.6(2) |
P1 | P2 | Se1 | — | 107.3(2) |
Se2 | P1 | Se2 | — | 113.8(3) |
Se1 | P2 | Se1 | — | 113.7(3) |
The new structure enables the construction of a preliminary phase diagram (Fig. 7), which considers only ordered structures, whether experimentally confirmed or theoretically predicted in the absence of experimental data (Table 4). The set of thermodynamically stable structures, those that form the convex hull based on their quantum-chemical formation energy at 0 K, are separated from metastable phases. Phases above the convex hull decompose into adjacent phases on the diagram and are thus not stable.
![]() | ||
Fig. 7 Phase diagram Li–P–Se based on calculated energies (Table 4). |
Phase | Pearson symbol | Source | Ef (kJ mol−1) | ΔEah/atom/meV | Source | Collection code |
---|---|---|---|---|---|---|
Pblack | oS8 | XRD | 0.0 | 0 | ICSD | 25253 |
Se8 | mP32 | XRD | 0.0 | 0 | ICSD | 2718 |
Li | cI2 | XRD | 0.0 | 0 | ICSD | 44759 |
Li2Se | cF12 | XRD | −370.9 | 0 | ICSD | 60433 |
Li3P | hP8 | XRD | −280.3 | 0 | ICSD | 26880 |
Li3P7 | oP40 | XRD | −383.8 | 0 | ICSD | 60774 |
LiP | mP16 | XRD | −107.7 | 0 | ICSD | 100465 |
LiP5 | oP24 | XRD | −136.0 | 10 | ICSD | 88710 |
LiP7 | tI128 | Unclear | −152.4 | 0 | ICSD | 23621 |
P2Se5 | mP28 | XRD | −20.3 | 23 | ICSD | 74546 |
P4Se3 | oP112 | XRD | −48.6 | 7 | ICSD | 26483 |
P4Se4 | mP32 | XRD | −71.1 | 0 | ICSD | 74878 |
P4Se5 | oP36 | XRD | −65.5 | 6 | ICSD | 16140 |
Li3PSe4 | cP8 | Predicted | −626.7 | 0 | OQMD | 7772844 |
Li7PSe6 | cP56 | Ag7PSe6 | −1309.8 | 43 | ICSD | 54055 |
Li4P2Se6 | oP48 | XRD, predicted | −874.0 | 0.15 | This work | |
Li4P2Se6 | hP12 | Predicted | −866.2 | 6.92 | This work | |
Li4P2Se6 | oP24 | Predicted | −873.8 | 0.30 | This work | |
Li4P2Se6 | mP24 | Predicted | −873.8 | 0.33 | This work | |
Li4P2Se6 | hP36 | Predicted | −874.2 | 0 | This work | |
Li4P2Se6 | oP48′ | Predicted | −874.1 | 0.06 | This work |
At a first glance, it is surprising that certain compounds, such as LiP5, which can be synthesized at high temperatures, are not predicted to be stable according to the calculated energies. This is unexpected, as it would be assumed that thermodynamic minima are easily achieved for such phases. Upon closer inspection, the small excess energies above the convex hull energy stand out (Table 4). These energy differences are minor when compared to the energy contributions from vibrational or configurational entropy.
For instance, at the synthesis temperature of 600 °C, the contribution from configurational entropy for Li7PSe6 amounts to approximately 5 kJ mol−1. This perspective suggests that using energy above the convex hull as a strict criterion for determining new stable structures needs to be applied with caution.31 Thus, structure prediction is not only limited by the limited search space that can be explored in practice but also by uncertainties in free energy. Kinetic control is still another point which may be difficult to consider in computational chemistry but may be less relevant for synthesis at high temperatures.
The measured impedance spectra were fitted with an equivalent circuit (Fig. 8, inset) consisting of a resistor Rb and a constant phase element Qb, depicting the conductive and non-ideal capacitive characteristics of the bulk region, respectively. The Qel depicts the non-ideal double layer formation at the sample/electrode interface. The impedance of a constant-phase element is given by ZQ = 1/Q(iω)α, with pre-factor Q and α < 1 exponent.
The effective capacitance Cb of the bulk region is calculated from the expression Cb = (QR(α−1))1/α,32 which is 55(9) pF at 393 K (Fig. 8).
The conductivity σ of the bulk region was determined from the fitted bulk resistance values Rb at corresponding temperatures using the expression , where d = 0.13 cm and A = 1.33 cm2 denote the sample thickness and the electrode area, respectively.
The activation energy of the macroscopic ionic conductivity was obtained using the given expression σ·T = a0·exp(−EA/(kBT)), where a0 is the pre-exponential factor, EA is the activation energy, kB is Boltzmann's constant, and T is the temperature (Fig. 9).
The activation energy and the corresponding pre-exponential factors are shown in Table 5. The difference of 1% between the conductivities and activation energies of the repeated measurements (ESI Fig. S2†) confirms the reproducibility of the results.
The contribution of the grain boundary could not be distinctly separated from that of the grain itself. Consequently, the measured activation energy of 0.78(1) eV for the ionic motion of Li+ in Li4P2Se6 likely represents an upper limit for the bulk activation energy. This value remains slightly lower than the activation energy determined for Li4P2S6, which is 0.88(2) eV.9 This reduction in activation energy is expected because Se2− has a higher size and polarizability than S2−, and the larger ionic radii of Se2− weaken the attractive force between the ligand and the Li-cation, which enhances the ion mobility.33 Moreover, the lattice softness of the anion sub-lattice has been shown to decrease the activation energy barrier of the cation transport in the crystal system.34 To enhance the ion transport mechanism, vacancies and interstitial lattice sites could be created by aliovalent doping in anionic as well as cationic lattice sites by substituting Se and Li atoms, but it is rather unlikely that conductivities in the mS cm−1 range can be achieved this way.9
In the context of the paddle-wheel mechanism,35 it is interesting to determine the activation energy for the rotation of a PSe3 fragment around the P–P bond. The calculated activation energy for the rotational motion in Li4P2Se6 (Fig. 6) was estimated by nudged elastic band calculations to be 1.29 eV. These results imply that the Li-ion motion is decoupled from the rotational motion of the complex anion. Hence the movement of ions can be treated independently from the rotation of PSe3 fragments.
Footnote |
† Electronic supplementary information (ESI) available. CCDC 2419766. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d5dt00227c |
This journal is © The Royal Society of Chemistry 2025 |