Next Article in Journal
NiO/TiO2 p-n Heterojunction Induced by Radiolysis for Photocatalytic Hydrogen Evolution
Previous Article in Journal
Conceptual Design of a Hybrid Composite to Metal Joint for Naval Vessels Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Theoretical Raman Spectra Analysis of the Effect of the Li2S and Li3PS4 Content on the Interface Formation Between (110)Li2S and (100)β-Li3PS4

by
Naiara Leticia Marana
*,
Eleonora Ascrizzi
,
Fabrizio Silveri
,
Mauro Francesco Sgroi
,
Lorenzo Maschio
and
Anna Maria Ferrari
*
Chemistry Department, University of Turin, 10125 Turin, Italy
*
Authors to whom correspondence should be addressed.
Materials 2025, 18(15), 3515; https://doi.org/10.3390/ma18153515
Submission received: 18 June 2025 / Revised: 18 July 2025 / Accepted: 23 July 2025 / Published: 26 July 2025
(This article belongs to the Section Advanced Materials Characterization)

Abstract

In this study, we perform density functional theory (DFT) simulations to investigate the Raman spectra of the bulk and surface phases of β-Li3PS4 (LPS) and Li2S, as well as their interfaces at varying compositional ratios. This analysis is relevant given the widespread application of these materials in Li–S solid-state batteries, where Li2S functions not only as a cathode material but also as a protective layer for the lithium anode. Understanding the interfacial structure and how compositional variations influence its chemical and mechanical stability is therefore crucial. Our results demonstrate that the LPS/Li2S interface remains stable regardless of the compositional ratio. However, when the content of both materials is low, the Raman-active vibrational mode associated with the [PS4]3− tetrahedral cluster dominates the interface spectrum, effectively obscuring the characteristic peaks of Li2S and other interfacial features. Only when sufficient amounts of both LPS and Li2S are present does the coupling between their vibrational modes become sufficiently pronounced to alter the Raman profile and reveal distinct interfacial fingerprints.

Graphical Abstract

1. Introduction

In the past decades, energy shortage problems have become increasingly serious, and the search for clean energy and batteries with high storage capacity, in addition to being economically viable, has become a main research focus [1]. High-energy-density batteries are considered to be one of the most efficient ways to store energy. Lithium–sulfur (Li–S) batteries are one of the most promising candidates for the next generation of batteries due to their characteristics of cost efficiency, natural abundance, environmental friendliness, and a high theoretical energy density of 2500 W h kg−1 (or 2800 W h L−1) [2]. Furthermore, all-solid-state lithium batteries (ASSLBs) have shown superior performance compared to liquid electrolyte batteries, as ASSLBs are more stable and less flammable. However, the interface problem between materials arises, where the formation of Li dendrites or by-products, in addition to interfacial defects, can hinder proper functioning [3].
In this scenario, the choice of solid electrolyte is important given the low fluidity and insufficient wettability of the ASSLBs. No matter how high the ionic conductivity of the solid electrolyte is, the distribution of active materials, solid electrolytes, and electrically conductive materials must be homogeneous in the electrodes to obtain solid-state cells with high levels of performance [4]. Among sulfide-type electrolytes, β-Li3PS4 (LPS) stands out for its high ionic conductivity (3.0 × 10−2 Scm−1 at 573 K) [5]. Unfortunately, the interface between LPS and the Li anode is not electrochemically stable, and the decomposition of the electrolyte into Li2S and Li3P affects Li transport, increasing cell resistivity. A possible solution for this problem is to insert a functional buffer layer between the LPS electrolyte and the Li anode, preserving the cell components and their properties. Li2S seems to be one of the best choices given its electrochemical inertness with respect to Li metal and the stability of the LPS/Li2S and Li2S/Li interfaces [6,7,8].
The Raman spectrum is a useful tool for identifying the chemical composition of materials and determining the degree of purity based on the appearance (disappearance) of the characteristic vibrational mode peaks of the product (reactant) after synthesis is complete. This non-invasive technique allows one to investigate the chemical nature of components in both ex situ and in situ conditions. In ASSLBs, Raman spectroscopy has also been employed to monitor the evolution of interfaces within practical electrode composites after or during electrochemical treatment [9,10,11]. In particular, the LPS/Li2S interface can be characterized through Raman spectroscopy, taking into account the effects of electrochemical cycling. Identifying characteristic peaks in Raman spectra is key to understanding whether precursors have reacted to form the LPS/Li2S interface and for determining whether undesired degradation products or Li2S2 clusters have been formed [12,13,14,15,16].
Therefore, in this paper, we theoretically discuss the Raman spectra for the (100) LPS and (110) Li2S surfaces and the interfaces formed by modulating the Li2S and LPS content in the heterostructure. For the first time, the simulated Raman spectra of β-LPS and Li2S surfaces have been reported. DFT hybrid calculations carried out with the CRYSTAL23 program [17] were applied in order to precisely characterize this stable interface, whose structure has been already described in a previous publication [6]. Specific spectral features will be identified and related to the atomic structure and the local environment (thickness of the slabs).

2. Materials and Methods

The density functional theory (DFT) approach was combined with the hybrid PBE0 functional [18,19] and all-electron basis set 6–11G [20], 86–311G* [21], and 85–211dG [22] for Li, S, and P, respectively, as implemented in the CRYSTAL program.The adopted computational approach, which was already applied in our previous studies [6,23], has provided structures and energies in good accordance with data available in the literature. The accuracy of the truncation criteria for the bi-electronic integrals, Coulomb, and HF exchange series was controlled by a set of five thresholds with values of [8, 8, 8, 8, 16]. The reciprocal space was sampled according to a sublattice with shrinking factor 8, corresponding to 25 independent k-points in the irreducible part of the Brillouin zone. The LPS/Li2S heterostructures are completely relaxed in order to avoid the appearance of imaginary frequencies caused by geometrical constraints.
Adhesion energy ( E a d h ) is a key parameter for evaluating the interaction and stability of interfaces between solid materials, calculated according to the following equation:
E a d h =   E i n t e r f a c e ( E L i 2 S + E L P S ) A
where E i n t e r f a c e   is the total energy of the heterostructure, A is the surface area of the interface, and E L i 2 S and E L P S are the total energies of the two structures relaxed at the lattice parameters defining the interface. All E a d h values have been corrected for basis set superposition error (BSSE) using the counterpoise method [24]. The BSSE correction does not exceed +3.55 meV/Å2 for PBE0 and +2.95 meV/Å2 for MN15. Both components are therefore either stretched or compressed during the interface formation process, and the energy cost for this deformation defines the strain energy (per surface unit), E s t r a i n , which has to be taken into account for a proper estimate of the overall stability of the composite. E s t r a i n was computed as
E s t r a i n ( L i 2 S ; L P S ) = E ( L i 2 S ; L P S ) E ( L i 2 S ; L P S ) ' 2 A
where E ( L i 2 S ; L P S ) is the energy of the fully relaxed LPS or Li2S.
As in previous studies [25,26], to better estimate noncovalent interactions involved in the interface formation, energy estimates have also been made by using the Minnesota hybrid functional (MN15) [27] by computing single-point energy calculations on the PBE0 optimized structures.
The vibrational frequencies at the Γ point were computed within the harmonic approximation by diagonalizing the mass-weighted Hessian matrix of the second derivatives of the total energy per cell with respect to the pair of atomic displacements in the reference cell. Once the vibrational frequencies are calculated, the Raman intensities and the simulated Raman spectra can also be obtained. A temperature of 298 K is considered, along with laser wavelengths selected to recreate experimental conditions, with 632.8 nm for LPS, 488.0 nm for Li2S, and 532.0 nm for the interfaces [28,29,30]. More details on the computational vibrational frequencies calculated with the CRYSTAL program can be found in Ref. [31].

3. Results and Discussion

3.1. Bulk and Isolated Surfaces

Before discussing the modifications in the Raman spectra of the β-LPS/Li2S heterostructures, we first examine the Raman spectra of the pristine bulk and surface structures, previously presented and characterized in Ref. [6]. To our knowledge, this is the first time the Raman spectrum of the (100) surface of β-LPS has been computed.
Figure 1a–c shows the Raman spectra of the bulk and (100) surface β-LPS with four and eight layers (4L and 8L). Both the general shape of the spectra and the position of the characteristic peak of the [PS4]3− vibration at ~425 cm−1 agree well with previous experimental and theoretical studies on bulk LPS [12,15,28,30,32]. This characteristic peak presents similar intensities in both the bulk and surface structures. In the [PS4]3− tetrahedron, the P–S bonds are covalent, and the overall structure is only moderately affected by the surrounding chemical environment. This limited influence is also reflected in the corresponding Raman features, which exhibit minimal variation. In addition to the intense main peak at 425 cm−1, the Raman spectra of both bulk and surface LPS also exhibit several signals in the 500–600 cm−1 range. These features are associated with vibrational modes involving the P–S within the PS4 tetrahedral units. At lower frequencies, additional peaks are observed, which are primarily attributed to the vibrational contributions of Li–S bonds present in the broader framework of the material. The complete description of vibrational modes, symmetry, and intensity are shown in Table S1.
When comparing the calculated Raman spectrum of bulk LPS with the experimental spectrum presented in Figure 1a, it is essential to emphasize that the peaks observed in the theoretical spectrum result from an idealized, perfectly ordered, and defect-free simulated structure. Conversely, the experimental sample exhibits a certain degree of structural disorder and defects (or are related to impurities during the synthesis), which leads to the broadening and attenuation of these peaks, making them less distinct and more challenging to resolve.
The Raman spectrum for the bulk Li2S has also been widely discussed in several articles in the literature [12,33], but to our knowledge, the experimentally acquired Raman spectrum for the (110) Li2S surface has not been reported. In contrast to LPS, Li2S exhibits a more pronounced difference between bulk and surface Raman spectra. The Raman spectrum of the bulk structure is characterized by a single peak at approximately 389 cm−1, which corresponds to Li–S stretching vibrations, as shown in Figure 1c. This well-defined feature reflects the high symmetry and long-range order of the crystalline bulk material. At the surface, however, the situation changes significantly. The reduced symmetry and increased degrees of freedom for the Li–S bonds lead to a splitting of at 389 cm−1 and the appearance of additional low-intensity peaks, particularly at lower and higher frequencies. These spectral changes are a direct consequence of structural rearrangements and local distortions confined to the outermost layers. As the number of layers increases, the Raman spectrum of the surface begins to resemble that of the bulk: the peaks gradually merge and become more defined, indicating a progressive restoration of the bulk-like crystallinity (compare the spectra reported in Figure 1e–h). For example, in the 36L_Li2S system, Figure 1h, the original 389 cm−1; peak is split into three components (at 390, 391, and 392 cm−1), and additional features appear around 344 cm−1 and 520 cm−1. These are attributed to S–Li–S bending, Li–S stretching, and mixed modes activated by surface symmetry breaking and to the Li–S–Li vibrational modes involving mainly the topmost surface layers (see Table S2 for more details).

3.2. LPS/Li2S Heterostructure

After having identified the characteristic Raman peaks of each material and their isolated surfaces, we can proceed to analyze the LPS/Li2S heterostructures. As previously discussed, these materials are widely employed in sulfur-based solid-state lithium batteries. Their interface commonly arises either at the cathode/solid-electrolyte boundary [34] or when Li2S is used as a protective layer for the lithium anode in contact with LPS [6,7,8]. In this context, understanding and identifying the features of the interface is essential to evaluating the behavior and stability of the system. Previous studies [6] conducted by our group have examined the LPS/Li2S interface particularly under conditions where only a limited amount of Li2S was introduced (4L of Li2S was considered). The results showed that the interface is chemically and mechanically stable, indicating a good intrinsic compatibility between the two materials.
Based on this, we decided to analyze the same interface (100)LPS/(110)Li2S, its stability, and its Raman spectrum profile, taking into account different amounts of LPS and Li2S. We simulated the following interfaces:
(1)
8L of LPS with 12L of Li2S (8L_LPS/12L_Li2S);
(2)
4L of LPS with 12L of Li2S (4L_LPS/12L_Li2S);
(3)
4L of LPS with 24L of Li2S (4L_LPS/24L_Li2S);
(4)
4L of LPS with 36L of Li2S (4L_LPS/36L_Li2S);
(5)
8L of LPS with 36L of Li2S (8L_LPS/36L_Li2S).
The optimized interface models are shown in Figure 2 and Figure S1, and their adhesion energies are shown in Table 1. According to our results, the content of Li2S and LPS at the interface has little effect on its chemical and mechanical stability, the Eadh values are between –27 and –31 meV/Å2, and only slight structural changes are observed, mostly confined to the interfacial region, with the non-interfacial layers remaining largely unaffected. The 4L_LPS/36L_Li2S and 8L_LPS/36L_Li2S models present a reduction in ~1 meV/Å2 in Eadh, which may be due to the high structural stability of Li2S that is mostly undisturbed by the presence of LPS and that recovers its bulk structure after the first 12 layers. Although the MN15 functional yields significantly larger energy values (ranging from –47 to –50 meV·Å−2), the predicted stability order of the heterostructures remains consistent with that computed using the PBE0 functional. The stability of the heterostructures can be assed evaluating the work of adhesion [35,36,37], Wadh = –(Eadh + Estrain). Wadh quantifies the energy required to separate the heterostructure into two isolated slabs, accounting for both the adhesion energy and the strain induced by heterostructure formation. In all the heterostructure examined in this work, Estrain is only a fraction of the Eadh, indicating thermodynamic stability. As a result, a positive Wadh must be supplied to separate the interfaces. Although the computed Eadh may appear relatively small, it is important to consider that such heterostructures are likely to form under experimental conditions, as the synthesis and stability of the interface in the device involves the application of external pressure [38,39].
From now on, we will analyze the simulated Raman spectra of all interfaces. Figure 3 contains the Raman spectra for the heterostructures and, for the sake of comparison, the pristine surfaces. As can be seen in Figure 3, the characteristic peak related to [PS4]3− of LPS appears in all the Raman spectra regardless of LPS content. The intensity of this peak is not significantly affected by the formation of the interface, although a slight shift in its position is observed. In general, the vibrational peaks associated with LPS remain identifiable in the Raman spectra of the interface, showing only subtle changes in both intensity and frequency compared to the pristine LPS surface. These variations can be attributed to the relative LPS content within the interface region.
It is important to note that the vibrational modes related to LPS—particularly those associated with the [PS4]3− tetrahedral unit—exhibit much higher intensities than those originating from the Li2S surface. As the Li2S content increases, however, more prominent and partially degenerate Li2S-related peaks begin to emerge. This effect is clearly illustrated by comparing the spectra of the 8L_LPS/4L_Li2S and 8L_LPS/12L_Li2S heterostructures (see Figure S2 and Figure 3b). In the latter, characteristic Li2S peaks become distinguishable at approximately 365 cm−1, 392 cm−1, and 511 cm−1, corresponding to Li–S bond stretching modes. This observation aligns with previous reports highlighting the experimental challenges in detecting Li2S during the synthesis of Li3PS4. In such cases, Raman measurements often fail to capture Li2S-related peaks following the formation of Li3PS4, likely due to their lower intensity and possible overlap with LPS signals [40].
In our case, significant modifications to the Raman spectra are observed for the 8L_LPS/36L_Li2S interface, where both LPS and Li2S are present in high concentrations. Under these conditions, the main LPS peak couples with the stretching vibrational modes of Li2S, resulting in the formation of three distinct peaks characteristic of the interface at 435 cm−1, 456 cm−1, and 468 cm−1. Additionally, typical Li2S vibrations appear at the interface as a split peak at 388 cm−1, a broad band around 300 cm−1 due to the overlap of Li–S stretching modes from both Li2S and LPS, and a distinct peak at 558 cm−1. Notably, the region between 570 cm−1 and 590 cm−1 continues to display LPS-related vibrations, with no significant contribution from Li2S modes. The details of the peaks and their descriptions are reported in Table S3 and discussed in Section S1 of Supplementary Information. The broadening and splitting of the peak around ~450 cm−1 can thus be regarded as a distinctive spectroscopic fingerprint indicative of the LPS/Li2S interface reflecting changes in local bonding environments and structural heterogeneity. The simulated spectrum likely represents the most accurate model of what would be observed experimentally in a well-formed interface with a balanced composition, providing an important reference for interpreting Raman data. This highlights the importance of detailed spectral analysis, which is essential not only for elucidating how the relative amounts of Li2S and LPS influence the Raman spectrum but also for enabling a reliable interpretation of the experimental results.
Li2S may be intentionally introduced to promote the formation of a stable interface with LPS, where it plays a beneficial role as a passivating coating that prevents the decomposition of LPS in contact with a lithium metal anode. However, Li2S can also arise as an unintended residual byproduct from the synthesis of LPS. In this context, residual Li2S impurities have been shown to adversely affect the ionic conductivity of the solid electrolyte, thereby compromising lithium metal battery performance by increasing overpotentials and shortening cycle life.
Therefore, understanding and controlling the interfacial composition through spectroscopic evaluation is essential for optimizing solid-state battery performance. According to Mirmira et al. [30], Raman spectroscopy could detect residual Li2S in LPS samples when the impurity concentration was above 12 mol%. However, for concentrations below this threshold, Raman spectroscopy’s sensitivity diminished, making it challenging to identify smaller amounts of Li2S impurities. They found that until 30 mol% of Li2S in the sample, the peaks related to the Li-S vibration of Li2S (at ~380 cm−1) are evident, and then the intensity decreases with the Li2S concentration due to the higher intensity of LPS vibrations. The same peak was also evident in the Li2S-LPS composite obtained by Jiang and co-workers [39]. The composite structure was evaluated by Raman spectroscopy and XRD diffraction, and the authors observed that when the lithium sulfide in the starting material was excessive, such as a 45% and 62.5% mass fraction of Li2S, the phase LPS was generated with a residual lithium sulfide, and the characteristic peaks related to LPS (at 420 cm−1 related to [PS4]3− cluster) and Li2S (at 380 cm−1 attributed to Li–S–Li vibration) were apparent in the Raman spectra of the composite. Thus, the presence of these signals does not necessarily indicate the formation of a stable interface but may instead correspond to residual impurities and, even more critically, residual Li2S may be present below the detection limit of Raman spectroscopy, making its presence difficult to identify and potentially leading to underestimated impurity levels that can adversely affect battery performance.

4. Conclusions

In this study, we performed a theoretical Raman spectra analysis to investigate the effect of Li2S and Li3PS4 content on the interface formation between (110) Li2S and (100) β-Li3PS4. Our results confirm that the LPS/Li2S interface remains stable regardless of its content. However, we observed that a higher concentration of LPS complicates the spectral identification of the interface, as the intense vibrational peak of the [PS4]3− cluster overshadows the characteristic Li2S peaks and other interface-related signals.
The analysis of heterostructures with different LPS and Li2S contents showed that, while small amounts of Li2S induce minimal modifications in the Raman spectrum, a significant increase in Li2S leads to substantial spectral changes. Notably, the broadening and splitting of the peak at ~450 cm−1, attributed to the coupling of [PS4]3− and Li2S vibrations, can be considered a spectral fingerprint of the formed interface. This feature was only observed at higher amounts of LPS and Li2S, suggesting its importance for identifying the interface experimentally.
Our findings provide valuable insights into the Raman spectral characteristics of LPS/Li2S interfaces, emphasizing the need for careful spectral interpretation, especially when working with low Li2S content. This study contributes to the understanding of interface formation in all-solid-state lithium batteries, which is key for optimizing their performance and stability.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma18153515/s1, Figure S1: (100)LPS/(110)Li2S interfaces with different LPS and Li2S contents: (a) 4L_LPS/24L_Li2S, (b) 4L_LPS/36L_Li2S, and (c) 8L_LPS/36L_Li2S; Figure S2: Raman spectra of the interface 8L_LPS/4L_Li2S and the respective pristine surfaces; Table S1: Vibrational mode descriptions of β-Li3PS4 bulk and its (100) surface; Table S2: Vibrational mode descriptions of Li2S bulk and its (100) surface; Table S3: Vibrational mode descriptions of the interface 8L_LPS/36L_Li2S; Discussion about the Raman spectra of the 8L_LPS/36L_Li2S interface compared with their isolated components.

Author Contributions

Conceptualization, N.L.M. and A.M.F.; methodology, N.L.M. and A.M.F.; validation, N.L.M. and A.M.F.; formal analysis, N.L.M.; investigation, N.L.M. and A.M.F.; resources, A.M.F.; data curation, N.L.M.; writing—original draft preparation, N.L.M. and A.M.F.; writing—review and editing, N.L.M., E.A., F.S., L.M., M.F.S. and A.M.F.; supervision, A.M.F.; project administration, A.M.F.; funding acquisition, A.M.F., L.M. and M.F.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article/Supplementary Materials. Further inquiries can be directed to the corresponding authors.

Acknowledgments

We acknowledge the support from Project CH4.0 under the MUR program ‘‘Dipartimenti di Eccellenza 2023–2027’’ (CUP: D13C22003520001). We also acknowledge the CINECA award under the ISCRA initiative for the availability of high-performance computing resources and support (ISCRAB project no. HP10BYBKEM). LM acknowledges support from the ICSC - Centro Nazionale di Ricerca in High Per- formance Computing, Big Data and Quantum Computing, which was funded by the European Union: Next- Generation EU (CUP grant noGrant No. J93C22000540006, PNRR Investimento Nono. M4.C2.1.4).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Pathak, A.D.; Cha, E.; Choi, W. Towards the Commercialization of Li-S Battery: From Lab to Industry. Energy Storage Mater. 2024, 72, 103711. [Google Scholar] [CrossRef]
  2. Zhong, M.; Guan, J.; Sun, J.; Shu, X.; Ding, H.; Chen, L.; Zhou, N.; Xiao, Z. A Cost- and Energy Density-Competitive Lithium-Sulfur Battery. Energy Storage Mater. 2021, 41, 588–598. [Google Scholar] [CrossRef]
  3. Wang, L.; Riedel, S.; Zhao-Karger, Z. Challenges and Progress in Anode-Electrolyte Interfaces for Rechargeable Divalent Metal Batteries. Adv. Energy Mater. 2024, 14, 2402157. [Google Scholar] [CrossRef]
  4. Liu, J.; Zhou, Y.; Yan, T.; Gao, X.P. Perspectives of High-Performance Li–S Battery Electrolytes. Adv. Funct. Mater. 2024, 34, 2309625. [Google Scholar] [CrossRef]
  5. Homma, K.; Yonemura, M.; Kobayashi, T.; Nagao, M.; Hirayama, M.; Kanno, R. Crystal Structure and Phase Transitions of the Lithium Ionic Conductor Li3PS4. Solid State Ion. 2011, 182, 53–58. [Google Scholar] [CrossRef]
  6. Marana, N.L.; Casassa, S.; Sgroi, M.F.; Maschio, L.; Silveri, F.; D’Amore, M.; Ferrari, A.M. Stability and Formation of the Li3PS4/Li, Li3PS4/Li2S, and Li2S/Li Interfaces: A Theoretical Study. Langmuir 2023, 39, 18797–18806. [Google Scholar] [CrossRef]
  7. Liu, F.; Wang, L.; Zhang, Z.; Shi, P.; Feng, Y.; Yao, Y.; Ye, S.; Wang, H.; Wu, X.; Yu, Y.; et al. A Mixed Lithium-Ion Conductive Li2S/Li2Se Protection Layer for Stable Lithium Metal Anode. Adv. Funct. Mater. 2020, 30, 2001607. [Google Scholar] [CrossRef]
  8. Lai, C.; Shu, C.; Li, W.; Wang, L.; Wang, X.; Zhang, T.; Yin, X.; Ahmad, I.; Li, M.; Tian, X.; et al. Stabilizing a Lithium Metal Battery by an in Situ Li2S-Modified Interfacial Layer via Amorphous-Sulfide Composite Solid Electrolyte. Nano Lett. 2020, 20, 8273–8281. [Google Scholar] [CrossRef]
  9. Zhou, Y.; Doerrer, C.; Kasemchainan, J.; Bruce, P.G.; Pasta, M.; Hardwick, L.J. Observation of Interfacial Degradation of Li6PS5Cl against Lithium Metal and LiCoO2 via In Situ Electrochemical Raman Microscopy. Batter. Supercaps 2020, 3, 647–652. [Google Scholar] [CrossRef]
  10. Otoyama, M.; Ito, Y.; Sakuda, A.; Tatsumisago, M.; Hayashi, A. Reaction Uniformity Visualized by Raman Imaging in the Composite Electrode Layers of All-Solid-State Lithium Batteries. Phys. Chem. Chem. Phys. 2020, 22, 13271–13276. [Google Scholar] [CrossRef]
  11. Otoyama, M.; Ito, Y.; Hayashi, A.; Tatsumisago, M. Raman Imaging for LiCoO2 Composite Positive Electrodes in All-Solid-State Lithium Batteries Using Li2S–P2S5 Solid Electrolytes. J. Power Sources 2016, 302, 419–425. [Google Scholar] [CrossRef]
  12. Jiao, Z.; Chen, L.; Si, J.; Xu, C.; Jiang, Y.; Zhu, Y.; Yang, Y.; Zhao, B. Core-Shell Li2S@Li3PS4 Nanoparticles Incorporated into Graphene Aerogel for Lithium-Sulfur Batteries with Low Potential Barrier and Overpotential. J. Power Sources 2017, 353, 167–175. [Google Scholar] [CrossRef]
  13. Stöffler, H.; Zinkevich, T.; Yavuz, M.; Hansen, A.L.; Knapp, M.; Bednarčík, J.; Randau, S.; Richter, F.H.; Janek, J.; Ehrenberg, H.; et al. Amorphous versus Crystalline Li3PS4: Local Structural Changes during Synthesis and Li Ion Mobility. J. Phys. Chem. C 2019, 123, 10280–10290. [Google Scholar] [CrossRef]
  14. Phuc, N.H.H.; Morikawa, K.; Totani, M.; Muto, H.; Matsuda, A. Chemical Synthesis of Li3PS4 Precursor Suspension by Liquid-Phase Shaking. Solid State Ion. 2016, 285, 2–5. [Google Scholar] [CrossRef]
  15. Lim, J.; Zhou, Y.; Powell, R.H.; Ates, T.; Passerini, S.; Hardwick, L.J. Localised Degradation within Sulfide-Based All-Solid-State Electrodes Visualised by Raman Mapping. Chem. Commun. 2023, 59, 7982–7985. [Google Scholar] [CrossRef]
  16. Partovi-Azar, P.; Kühne, T.D.; Kaghazchi, P. Evidence for the Existence of Li2S2 Clusters in Lithium–Sulfur Batteries: Ab Initio Raman Spectroscopy Simulation. Phys. Chem. Chem. Phys. 2015, 17, 22009–22014. [Google Scholar] [CrossRef]
  17. Erba, A.; Desmarais, J.; Casassa, S.; Civalleri, B.; Donà, L.; Bush, I.; Searle, B.; Maschio, L.; Daga, L.E.; Cossard, A.; et al. CRYSTAL23: A Program for Computational Solid State Physics and Chemistry. J. Chem. Theory Comput. 2022, 19, 6891–6932. [Google Scholar] [CrossRef]
  18. Perdew, J.P.; Wang, Y. Accurate and Simple Analytic Representation of the Electron-Gas Correlation Energy. Phys. Rev. B 1992, 45, 13244. [Google Scholar] [CrossRef]
  19. Adamo, C.; Barone, V. Toward Reliable Density Functional Methods without Adjustable Parameters: The PBE0 Model. J. Chem. Phys. 1999, 110, 6158. [Google Scholar] [CrossRef]
  20. Ojamäe, L.; Hermansson, K.; Pisani, C.; Causà, M.; Roetti, C. Structural, Vibrational and Electronic Properties of a Crystalline Hydrate from Ab Initio Periodic Hartree–Fock Calculations. Acta Crystallogr. Sect. B 1994, 50, 268–279. [Google Scholar] [CrossRef]
  21. Lichanot, A.; Aprà, E.; Dovesi, R. Quantum Mechnical Hartree-Fock Study of the Elastic Properties of Li2S and Na2S. Phys. Status Solidi 1993, 177, 157–163. [Google Scholar] [CrossRef]
  22. Zicovich-Wilson, C.M.; Bert, A.; Roetti, C.; Dovesi, R.; Saunders, V.R. Characterization of the Electronic Structure of Crystalline Compounds through Their Localized Wannier Functions. J. Chem. Phys. 2001, 116, 1120. [Google Scholar] [CrossRef]
  23. Marana, N.L.; Sgroi, M.F.; Maschio, L.; Ferrari, A.M.; D’Amore, M.; Casassa, S. Computational Characterization of β-Li3PS4 Solid Electrolyte: From Bulk and Surfaces to Nanocrystals. Nanomaterials 2022, 12, 2795. [Google Scholar] [CrossRef]
  24. Boys, S.F.; Bernardi, F. The Calculation of Small Molecular Interactions by the Differences of Separate Total Energies. Some Procedures with Reduced Errors. Mol. Phys. 1970, 19, 553–566. [Google Scholar] [CrossRef]
  25. Rocca, R.; Sgroi, M.F.; Camino, B.; D’Amore, M.; Ferrari, A.M. Disordered Rock-Salt Type Li2TiS3 as Novel Cathode for LIBs: A Computational Point of View. Nanomaterials 2022, 12, 1832. [Google Scholar] [CrossRef]
  26. Marana, N.L.; Silveri, F.; de Oliveira Gomes, E.; Donà, L.; D’Amore, M.; Ascrizzi, E.; Sgroi, M.F.; Maschio, L.; Ferrari, A.M. A Computational Study of the Negative LiIn Modified Anode and Its Interaction with β-Li3PS4 Solid–Electrolyte for Battery Applications. Phys. Chem. Chem. Phys. 2024, 26, 15648–15656. [Google Scholar] [CrossRef]
  27. Yu, H.S.; He, X.; Li, S.L.; Truhlar, D.G. MN15: A Kohn–Sham Global-Hybrid Exchange–Correlation Density Functional with Broad Accuracy for Multi-Reference and Single-Reference Systems and Noncovalent Interactions. Chem. Sci. 2016, 7, 5032–5051. [Google Scholar] [CrossRef]
  28. Wu, X.; El Kazzi, M.; Villevieille, C. Surface and Morphological Investigation of the Electrode/Electrolyte Properties in an All-Solid-State Battery Using a Li2S-P2S5 Solid Electrolyte. J. Electroceramics 2017, 38, 207–214. [Google Scholar] [CrossRef]
  29. Bertheville, B.; Bill, H.; Hagemann, H. Experimental Raman Scattering Investigation of Phonon Anharmonicity Effects In Li2S. J. Phys. Condens. Matter 1998, 10, 2155. [Google Scholar] [CrossRef]
  30. Mirmira, P.; Zheng, J.; Ma, P.; Amanchukwu, C. V Importance of Multimodal Characterization and Influence of Residual Li2S Impurity in Amorphous Li3PS4 Inorganic Electrolytes. J. Mater. Chem. A 2021, 9, 19637–19648. [Google Scholar] [CrossRef]
  31. Zicovich-Wilson, C.M.; Pascale, F.; Roetti, C.; Saunders, V.R.; Orlando, R.; Dovesi, R. Calculation of the Vibration Frequencies of α-Quartz: The Effect of Hamiltonian and Basis Set. J. Comput. Chem. 2004, 25, 1873–1881. [Google Scholar] [CrossRef]
  32. Ates, T.; Neumann, A.; Danner, T.; Latz, A.; Zarrabeitia, M.; Stepien, D.; Varzi, A.; Passerini, S.; Ates, T.; Neumann, A.; et al. Elucidating the Role of Microstructure in Thiophosphate Electrolytes—A Combined Experimental and Theoretical Study of β-Li3PS4. Adv. Sci. 2022, 9, 2105234. [Google Scholar] [CrossRef] [PubMed]
  33. Barkalov, O.I.; Naumov, P.G.; Felser, C.; Medvedev, S.A. Pressure-Induced Transition to Ni2In-Type Phase in Lithium Sulfide (Li2S). Solid State Sci. 2016, 61, 220–224. [Google Scholar] [CrossRef]
  34. Jiang, H.; Han, Y.; Wang, H.; Zhu, Y.; Guo, Q.; Jiang, H.; Sun, W.; Zheng, C.; Xie, K. In-Situ Generated Li2S-Based Composite Cathodes with High Mass and Capacity Loading for All-Solid-State Li-S Batteries. J. Alloys Compd. 2021, 874, 159763. [Google Scholar] [CrossRef]
  35. Schönberger, U.; Andersen, O.K.; Methfessel, M. Bonding at Metal-Ceramic Interfaces; AB Initio Density-Functional Calculations for Ti and Ag on MgO. Acta Metall. Mater. 1992, 40, S1–S10. [Google Scholar] [CrossRef]
  36. Finnis, M.W. The Theory of Metal—Ceramic Interfaces. J. Phys. Condens. Matter 1996, 8, 5811. [Google Scholar] [CrossRef]
  37. Mahmoud, A.; Maschio, L.; Sgroi, M.F.; Pullini, D.; Ferrari, A.M. Ab Initio Simulation of ZnO/LaMnO3 Heterojunctions: Insights into Their Structural and Electronic Properties. J. Phys. Chem. C 2017, 121, 25333–25341. [Google Scholar] [CrossRef]
  38. Jiang, H.; Han, Y.; Wang, H.; Guo, Q.; Zhu, Y.; Xie, W.; Zheng, C.; Xie, K. In Situ Generated Li2S-LPS Composite for All-Solid-State Lithium-Sulfur Battery. Ionics 2020, 26, 2335–2342. [Google Scholar] [CrossRef]
  39. Jiang, H.; Han, Y.; Wang, H.; Zhu, Y.; Guo, Q.; Jiang, H.; Zheng, C.; Xie, K. Li2S–Li3PS4 (LPS) Composite Synthesized by Liquid-Phase Shaking for All-Solid-State Lithium–Sulfur Batteries with High Performance. Energy Technol. 2020, 8, 2000023. [Google Scholar] [CrossRef]
  40. Yamamoto, K.; Takahashi, M.; Ohara, K.; Phuc, N.H.H.; Yang, S.; Watanabe, T.; Uchiyama, T.; Sakuda, A.; Hayashi, A.; Tatsumisago, M.; et al. Synthesis of Sulfide Solid Electrolytes through the Liquid Phase: Optimization of the Preparation Conditions. ACS Omega 2020, 5, 26287–26294. [Google Scholar] [CrossRef]
Figure 1. Computed Raman spectra of pristine materials at 298 K. Spectra were obtained using a wavelength of 632.8 nm for LPS and 488.0 nm for Li2S. Panels show (a) LPS bulk; (b) Li2S bulk; (c) and (d) LPS (100) surface with four and eight layers, respectively; (eh) Li2S (110) surface with 4, 12, 24, and 36 layers, respectively. Experimental bulk spectra are reproduced from Refs. [28,29].
Figure 1. Computed Raman spectra of pristine materials at 298 K. Spectra were obtained using a wavelength of 632.8 nm for LPS and 488.0 nm for Li2S. Panels show (a) LPS bulk; (b) Li2S bulk; (c) and (d) LPS (100) surface with four and eight layers, respectively; (eh) Li2S (110) surface with 4, 12, 24, and 36 layers, respectively. Experimental bulk spectra are reproduced from Refs. [28,29].
Materials 18 03515 g001
Figure 2. (100)LPS/(110)Li2S interfaces: (a) 4L_LPS/12L_Li2S and (b) 8L_LPS/12L_Li2S. Lithium, sulfur, and phosphorus atoms are represented as purple, yellow, and orange spheres, respectively.
Figure 2. (100)LPS/(110)Li2S interfaces: (a) 4L_LPS/12L_Li2S and (b) 8L_LPS/12L_Li2S. Lithium, sulfur, and phosphorus atoms are represented as purple, yellow, and orange spheres, respectively.
Materials 18 03515 g002
Figure 3. Raman spectra of the interface and the respective pristine surfaces at 298 K and a wavelength of 632.8 nm for LPS, 488.0 nm for Li2S, and 532.0 nm for the interfaces: (a) 4L_LPS/12L_Li2S, (b) 8L_LPS/12L_Li2S, (c) 4L_LPS/24L_Li2S, (d) 4L_LPS/36L_Li2S, and (e) 8L_LPS/36L_Li2S.
Figure 3. Raman spectra of the interface and the respective pristine surfaces at 298 K and a wavelength of 632.8 nm for LPS, 488.0 nm for Li2S, and 532.0 nm for the interfaces: (a) 4L_LPS/12L_Li2S, (b) 8L_LPS/12L_Li2S, (c) 4L_LPS/24L_Li2S, (d) 4L_LPS/36L_Li2S, and (e) 8L_LPS/36L_Li2S.
Materials 18 03515 g003
Table 1. Eadh and Estrain (meV/Å2) of all interface models computed as in Equations (1) and (2). a and b represent the cell parameters of the surfaces and interfaces.
Table 1. Eadh and Estrain (meV/Å2) of all interface models computed as in Equations (1) and (2). a and b represent the cell parameters of the surfaces and interfaces.
abEstrain (Li2S)Estrain (LPS) E a d h E a d h
PBE0MN15
(110)Li2S5.708.06-- -
(100)LPS6.238.28-- -
4L_LPS/12L_Li2S5.828.041.397.36−29.15−48.32
8L_LPS/12L_Li2S5.888.054.035.29−28.79−47.81
4L_LPS/24L_Li2S5.828.040.337.45−27.76−50.54
4L_LPS/36L_Li2S5.82 8.041.027.50−31.23−48.99
8L_LPS/36L_Li2S5.888.060.745.34−29.12−47.82
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Marana, N.L.; Ascrizzi, E.; Silveri, F.; Sgroi, M.F.; Maschio, L.; Ferrari, A.M. A Theoretical Raman Spectra Analysis of the Effect of the Li2S and Li3PS4 Content on the Interface Formation Between (110)Li2S and (100)β-Li3PS4. Materials 2025, 18, 3515. https://doi.org/10.3390/ma18153515

AMA Style

Marana NL, Ascrizzi E, Silveri F, Sgroi MF, Maschio L, Ferrari AM. A Theoretical Raman Spectra Analysis of the Effect of the Li2S and Li3PS4 Content on the Interface Formation Between (110)Li2S and (100)β-Li3PS4. Materials. 2025; 18(15):3515. https://doi.org/10.3390/ma18153515

Chicago/Turabian Style

Marana, Naiara Leticia, Eleonora Ascrizzi, Fabrizio Silveri, Mauro Francesco Sgroi, Lorenzo Maschio, and Anna Maria Ferrari. 2025. "A Theoretical Raman Spectra Analysis of the Effect of the Li2S and Li3PS4 Content on the Interface Formation Between (110)Li2S and (100)β-Li3PS4" Materials 18, no. 15: 3515. https://doi.org/10.3390/ma18153515

APA Style

Marana, N. L., Ascrizzi, E., Silveri, F., Sgroi, M. F., Maschio, L., & Ferrari, A. M. (2025). A Theoretical Raman Spectra Analysis of the Effect of the Li2S and Li3PS4 Content on the Interface Formation Between (110)Li2S and (100)β-Li3PS4. Materials, 18(15), 3515. https://doi.org/10.3390/ma18153515

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop