Orientational Relationship Between the Solid-Electrolyte Interphase and Li4Ti5O12 Electrode in Hybrid Aqueous Electrolytes

Article information

J. Electrochem. Sci. Technol. 2024;15(4):476-483
Publication date (electronic) : 2024 May 20
doi : https://doi.org/10.33961/jecst.2024.00416
1Agency for Defense Development, P.O. Box 35, Yuseong, Daejeon 34186, Republic of Korea
2Graduate School of Energy Science and Technology (GEST), Chungnam National University, Daejeon 34134, Republic of Korea
*E-mail address: shinhy@cnu.ac.kr (H. Shin), richpine87@gmail.com (Y. Choi)
†These authors contributed equally to this work.
Received 2024 April 25; Accepted 2024 May 20.

Abstract

Lithium-ion (Li-ion) batteries are essential to modern society, but pose safety risks because of thermal runaway and ignition. This study explores the use of hybrid aqueous electrolytes to enhance the safety and performance of Li-ion batteries, focusing on the solid-electrolyte interface (SEI) formed on lithium titanate (Li4Ti5O12; LTO) electrodes. We employed high-resolution transmission electron microscopy (HRTEM) and density functional theory (DFT) calculations to analyze the microstructure and stability of the SEI layer. Further, we prepared LTO and LiMn2O4 (LMO) electrodes, assembled full cells with hybrid aqueous electrolytes, and carried out electrochemical testing. The HRTEM analysis revealed the epitaxial growth of a LiF SEI layer on the LTO electrode, which has a coherent lattice structure that enhances electrochemical stability. The DFT calculations confirmed the energetic favorability of the LiF–LTO interface, indicating strong adhesion and potential for epitaxial growth. The full cell demonstrated excellent discharge performance, showing a notable improvement in coulombic efficiency after the initial cycle and sustained capacity over 100 cycles. Notably, the formation of a dense, crystalline LiF SEI layer on the LTO electrode is crucial for preventing continuous side reactions and maintaining mechanical stability during cycling. The experimental results, supported by the DFT results, highlight the importance of the orientational relationship between the SEI and the electrode in improving battery performance. The integration of experimental techniques and computational simulations has led to the development of an LTO/LMO full cell with enhanced discharge capabilities and stability. This study provides insights into the growth mechanisms of the SEI layer and its impact on battery performance, demonstrating the potential of hybrid aqueous electrolytes in advancing lithium-ion battery technology. The findings affirm the viability of this approach for optimizing next-generation Li-ion batteries, which can promote the development of safer and more reliable energy storage solutions.

1. Introduction

Lithium-ion (Li-ion) batteries are crucial in the current digital electronic era, powering a vast range of devices, from handheld electronics to electric vehicles and large-scale energy storage systems [13]. Despite their widespread adoption, Li-ion batteries have significant safety risks, notably thermal runaway and ignition [4]. To address these concerns, researchers have focused on the development of safer alternatives, such as batteries with aqueous electrolyte systems. However, these alternatives typically have lower energy densities and operational voltages because of their limited electrochemical windows [5].

Recently, Suo et al. introduced a “water-in-salt” (WIS) electrolyte that substantially widens the electrochemical window of aqueous batteries, enabling higher voltage operations while enhancing safety by reducing flammability [6,7]. Nevertheless, this breakthrough led to challenges in controlling electrolyte reactivity to prevent harmful side reactions, especially at the electrode interface. To resolve this issue, Wang et al. proposed a hybrid electrolyte with aqueous and non-aqueous components that generates a solid-electrolyte interface (SEI) on the anode [8], which is crucial for improving Li-ion battery safety and efficiency by preventing electrode degradation and electrolyte decomposition. The SEI comprises organic and inorganic phases, and LiF is a particularly effective phase for passivating electrodes and boosting cycling stability [1,3,9]. The microstructure and composition of the SEI have key influences on battery performance. For example, a porous SEI can lead to continuous breakdown and repair cycles [10], inducing side reactions and reducing battery life. Conversely, a dense, highly crystalline SEI remains intact during charge–discharge cycles, minimizing side reactions and enhancing cyclic performance [11].

Even though the mechanism of SEI formation, its composition, and its influence on battery performance have been extensively studied [1,11], all the SEI characteristics are not fully understood. Therefore, sophisticated analytical methods are required to examine the complex structure and function of SEI layers.

In this study, we used high-resolution transmission electron microscopy (HRTEM) to examine the microstructure of hybrid aqueous/non-aqueous electrolytes and their role in stabilizing SEI formation on lithium titanate (Li4Ti5O12; LTO). We also performed atomic-scale analysis of the interactions of the LiF SEI layer and its stability on the LTO electrode through density functional theory (DFT) calculations. We believe that this study contributes to the development of safer and more reliable batteries.

2. Experimental

2.1 Electrode and Electrolyte preparation

Cathode electrodes comprising LiMn2O4 (LMO, L&F Co., Ltd.) were produced by blending the active material, carbon black (Super C, Timcal), and polyvinylidene fluoride (PVDF, Solvay) in a 90:5:5 weight ratio with N-methylpyrrolidinone (NMP, 99.5%, Sigma-Aldrich). The mixture was spread onto 20-μm-thick titanium foil and dried in vacuum at 100°C for 12 h. The same procedure was followed to prepare anode electrodes comprising Li4Ti5O12 (LTO, POSCO Future M), except for the use of aluminum foil as the substrate. The hybrid aqueous electrolyte (HT-29) was formulated by combining lithium bis(trifluoromethanesulfonyl)imide (LiTFSI, 99.95%, Sigma-Aldrich), water, and trimethylphosphate (TMP, 99%, Sigma-Aldrich) in a 37:40:23 molar ratio [12].

2.2 Electrolyte characterization

The ionic conductivity of the electrolyte was determined by a conductivity meter (S213-meter, Mettler-Toledo), and the viscosity of the electrolyte was determined using a rotational viscometer (Brookfield DV3TLV). The linear sweep voltammetry (LSV) tests for the electrolyte were performed with a three-electrode system constituting an inactive current collector (titanium foil), activated carbon, and Ag/AgCl as the working, counter, and reference electrodes, respectively. The LSV measurements were conducted on a Solatron 1260 instrument equipped with a 1287 Electrochemical Interface (Solatron Metrology, UK) at a scanning rate of 0.1 mV s−1.

2.3 Cell assembly and Testing

The cathode, anode, aqueous electrolyte, and separator were assembled in a dry room with a controlled relative humidity of <0.2%, as depicted in Fig. 1. A 260-μm-thick Whatman GF/A glass fiber was used as the separator. A standard 2032-coin cell was filled with 100 μL of electrolyte, sealed, and allowed to stand for at least 12 h to ensure thorough impregnation. Charge–discharge testing was conducted in a temperature-controlled environment using a Maccor 4000 battery tester, cycling between 2.8 and 1.5 V at a 0.5C rate (0.5 mA cm−2).

Fig. 1

Schematic of the fabrication of the coin cell with a LiMn2O4 (LMO) cathode, a Li4Ti5O12 (LTO) anode, and a hybrid aqueous electrolyte.

2.4 Microstructural analysis

The post-cycling crystallographic orientation of the SEI and LTO was determined using field-emission transmission electron microscopy (FE-TEM, JEOL, JEM-2100F) at 200 kV. Samples were prepared with a focused ion beam (FIB, JEOL, JIB-4601F). HRTEM images and fast Fourier-transform (FFT) analysis were used to verify the crystal structure of the SEI on the LTO electrode. The chemical composition of the SEI was further analyzed using X-ray photoelectron spectroscopy (XPS, Axis-Supra, Kratos).

2.5 Computational details

Spin-polarized DFT calculations were conducted using the Vienna Ab initio Simulation Package (VASP) [13] using the projector-augmented wave (PAW) method and the Perdew–Burke–Ernzerhof (PBE) [14] exchange-correlation functional within the generalized gradient approximation (GGA). The Grimme D3 dispersion correction was employed to account for van der Waals interactions [15], and calculations were performed with a plane-wave basis set with a cutoff energy of 600 eV for increased accuracy.

For bulk structure optimization, Monkhorst–Pack (6×6×6) k-point meshes for LiF and (3×3×3) for LTO were used. Slab models representing the (200) crystallographic planes of LiF and LTO were derived from these optimized bulk structures, selected based on the results of TEM analysis. The slabs were further optimized with (6×6×1) k-point meshes for LiF and (3×3×1) for LTO. The lowest-energy configurations from various (200) slab models with different surface terminations were chosen for subsequent analyses, as shown in Fig. 2 and 3. The LiF (200)/LTO (200) interface model, constructed from these slabs, was refined using a (6×3×1) k-point meshes, as illustrated in Fig. 4. A 20 Å vacuum layer was introduced perpendicular to the slabs to prevent artificial periodic interactions, and the bottom two layers were fixed to mimic the bulk conditions, allowing the upper layers to relax.

Fig. 2

Optimized structures of Li4Ti5O12: (a) bulk, (b) side view, and (c) top view of the (200) plane surface of Li4Ti5O12. Green, blue, and red spheres represent Li, Ti, and O atoms, respectively. Li+ ions on the 16d sites are encircled; other Li+ ions occupy the 8a sites.

Fig. 3

Optimized structures of LiF: (a) bulk, (b) side view, and (c) top view of the (200) plane surface of LiF. Green and gray spheres represent Li and F atoms, respectively.

Fig. 4

Optimized structures of the LiF(200)/Li4Ti5O12(200) interface: (a) side view, and (b) top view. Green, blue, red, and gray spheres represent Li, Ti, O, and F atoms, respectively.

The chemical stability and chemical bonding at the LiF/LTO interface were assessed by calculating the formation energy (Ef) as follows:

Ef=ELiF/LTO-ELTO-ELiF

where ELiF/LTO, ELTO, and ELiF represent the total energies of the combined interface, LTO layer, and LiF layer, respectively.

3. Results and Discussion

3.1 Cell performance

Table 1 shows the ion conductivity of the hybrid aqueous electrolyte at 23C and its viscosity at 25C and 40C. The ion conductivity of this hybrid aqueous electrolyte was smaller than that of a non-aqueous electrolyte [16]. Its viscosity was higher than that of a traditional non-aqueous electrolyte, owing to the high concentration of LiTFSI salts in the former.

Ion conductivity of the hybrid aqueous electrolyte at 23°C and its viscosity at 25°C and 40°C

The electrochemical stability window (ESW) of the hybrid aqueous electrolyte was determined from the linear sweep voltammetry curves (Fig. 5). The cathodic and anodic limits were 1.76 and 4.66 V, respectively, with an overall ESW of 2.9 V. The redox reaction potentials of LMO and LTO are completely inside the ESW of the hybrid aqueous electrolyte.

Fig. 5

Linear sweep voltammetry curve of the hybrid aqueous electrolyte showing the electrochemical stability window.

An LMO–LTO full cell was constructed to assess the electrochemical performance of the hybrid aqueous electrolyte. The initial voltage profile, recorded at a current density of 0.5C (0.5 mA cm−2), is shown in Fig. 6. The first cycle exhibited a charge voltage plateau at ~2.50 V and a discharge plateau at 2.35 V, indicative of Li+-ion insertion and removal in the LMO–LTO system. The initial discharge capacity (146.1 mAh g−1) was marginally lower than the theoretical capacity (175 mAh g−1). The coulombic efficiency of the first cycle was 80.1%, suggesting electrolyte consumption for SEI formation and H2O decomposition at the anode. In the subsequent cycle, the coulombic efficiency improved to above 90%, suggesting minimal interface formation. After 100 cycles at 0.5C, the cell retained a discharge capacity of 51.9 mAh g−1, which is 35.4% of the initial capacity, and a coulombic efficiency of ~97.8%.

Fig. 6

Charge/discharge cycle performance at 0.5C.

The XPS spectra of the SEI on the anode are shown in Fig. 7. The F 1s spectrum of the LTO surface shows peaks of −CF2 (686.3 eV) of the PVDF binder in the electrode, −CF3 (688.6 eV) of the TFSI salt anion, and the inorganic LiF (685.3 eV) formed by the reduction of [Li]2+[TFSI]. This suggests that the LiF molecules within the SEI effectively inhibit the hydrogen evolution reaction on the anode by creating a protective film on the LTO surface.

Fig. 7

X-ray photoelectron spectra of the Li4Ti5O12 anode cycled 30 times.

Additionally, the C 1s spectrum exhibited peaks of −CF3 (293 eV) from TFSI or its reduction byproducts, −CO3 (290.5 eV) originated from the reaction of CO2 with H2O, alkyl carbonates (289 eV), ether (C–O) (286 eV), and other carbonaceous sp3- or sp2-hybridized elemental carbon species below 284 eV. The −CO3 peak is indicative of the interaction between atmospheric CO2 and H2O [17].

3.2 Microstructure of the SEI

Fig. 8 shows a high-resolution image of the specimen after 30 cycles, in which the SEI layer is discernible by the fringe contrast with the LTO electrode, measuring ~10 nm in thickness, as shown in Fig. 8a. Before FIB sample preparation, a Pt layer was applied to the SEI surface to prevent ion-beam damage [18]. The FFT analysis (Fig. 8b) identified the crystal structure of the SEI, as LiF has a face-centered cubic (FCC) lattice (Fm-3m; a = 4.027 nm). The epitaxial growth of the SEI on the LTO electrode is facilitated by the minimal lattice mismatch and the similarity in the FCC crystal structures of the LTO and LiF phases. This allows the LTO to serve as a substrate for the LiF phase. The results confirm that the LiF SEI layer grew epitaxially on the LTO electrode surface, for which the (200) and (111) planes are parallel between the two phases. HR-TEM observations revealed a semi-coherent interface between the LTO and LiF SEI layers, as indicated by the nearly 2:1 ratio of their lattice constants. This small lattice mismatch suggests that the LTO electrode is an effective nucleation site for the LiF SEI layer. The well-formed SEI layers prevent continuous side reactions and repeated SEI formation, leading to reduced resistance and enhanced stability [11]. The crystalline and dense SEI layer fully passivates the electrode surface, maintaining mechanical stability during cycling. The epitaxially grown LiF SEI on the LTO electrode demonstrates uniformity and crystallinity with minimal defects, effectively isolating the electrode from the reactive aqueous electrolyte and providing a robust surface.

Fig. 8

High-resolution transmission electron microscopy images of (a) the LTO electrode/SEI layer interface and (b) the corresponding fast Fourier-transform image obtained with beams parallel to the [011] direction.

3.3 DFT calculations

DFT calculations were conducted to investigate the atomic-scale interactions and stability of the LiF SEI layer on the LTO electrode. The results indicate an energetically favorable LiF-LTO interface structure with a formation energy of −5.26 eV, suggesting strong adhesion and a thermodynamically stable interface. Specifically, upon the formation of the interface, the bond length of Li–F within the LiF(200) surface decreased from 2.04 to 1.97 Å. Additionally, the bond lengths of Ti–O and Li–O within the LTO (200) surface changed from 1.93 to 2.09 Å and 1.97 to 1.92 Å, respectively. Moreover, as a direct consequence of the interface formation, new Li–O and Ti–F bonds with bond lengths of 1.98 and 2.13 Å, respectively, have emerged (Fig. 4). Furthermore, the lattice mismatch between LiF and LTO is only 2.22%, suggesting the epitaxial growth of the SEI layer on the LTO substrate. Significant charge redistribution was observed at the LiF/LTO interface, which is critical for the stability of the SEI (Fig. 9). The charge density difference (Δρ) for LiF/LTO was calculated as follows:

Fig. 9

Charge density difference of the LiF(200)/Li4Ti5O12(200) interface, visualized using VESTA [19]. Areas of charge accumulation are indicated in yellow, whereas regions of charge depletion are shown in cyan. Isosurface = 0.0015 e Bohr−3. Green, blue, red, and gray spheres represent Li, Ti, O, and F atoms, respectively.

Δρ=ρ(LiF/LTO)-(ρLTO+ρLiF)

where ρ(LiF/LTO), ρLTO, and ρLiF represent the charge densities of the interface, the LTO layer, and the LiF layer, respectively. The results of these calculations indicate electron redistribution and localization because of interfacial interactions. In addition, Bader charge analysis [20] revealed that the LTO layers gain 0.30 electrons, whereas the LiF layers lose 0.30 electrons, which enhanced the ionic character of the bonding and promoted stability at the LiF/LTO interface.

4. Conclusions

In this study, a LTO/LMO full cell was fabricated with a hybrid aqueous electrolyte. The cell showed superior discharge performance during 0.5C cycle tests. TEM revealed the epitaxial growth of the LiF SEI layer on the LTO electrode, in which the (200) and (111) planes were aligned between the two phases. This structural conformance results in lattice coherence at the interface, which significantly improves the electrochemical stability and functionality of the cell. HRTEM revealed a semi-coherent interface, suggesting that the LTO substrate served as a nucleation site for the LiF SEI layer. DFT calculations further confirmed the stability and favorable energetics of the formation of an LiF SEI layer on the LTO substrate which is consistent with the experimental evidence of epitaxial growth. The synergistic use of experimental and computational methods in this study contributes to the development of LTO/LMO full cells, especially for enhancing both discharge capacity and stability. The DFT calculations and HRTEM findings provide an in-depth understanding of the growth mechanisms of the SEI layer and its influence on battery performance. This study highlights the necessity of an integrated approach to refining lithium-ion batteries for the future.

Acknowledgments

This work was supported by the Agency for Defense Development of the Republic of Korea (Grant No. 912941101).

Notes

Declaration of competing interest

There are no conflicts of interest to declare.

References

1. Li B, Chao Y, Li M, Xiao Y, Li R, Yang K, Cui X, Xu G, Li L, Yang C, Yu Y, Wilkinson D. P, Zhang J. Electrochem. Energy Rev 2023;6(1):7.
2. Meda U. S, Lal L, S. M. , Garg P. J. Energy Storage 2022;47:103564.
3. Tan J, Matz J, Dong P, Shen J, Ye M. Adv. Energy Mater 2021;11(16):2100046.
4. Xu K. Chem. Rev 2004;104(10):4303–4417.
5. Kim H, Hong J, Park K.-Y, Kim H, Kim S.-W, Kang K. Chem. Rev 2014;114(23):11788–11827.
6. Suo L, Borodin O, Gao T, Olguin M, Ho J, Fan X, Luo C, Wang C, Xu K. Science 2015;350:938–943.
7. Yang C, Chen J, Qing T, Fan X, Sun W, von Cresce A, Ding M. S, Borodin O, Vatamanu J, Schroeder M. A, Eidson N, Wang C, Xu K. Joule 2017;1(1):122–132.
8. Wang F, Borodin O, Ding M. S, Gobet M, Vatamanu J, Fan X, Gao T, Edison N, Liang Y, Sun W, Greenbaum S, Xu K, Wang C. Joule 2018;2(5):927–937.
9. Zhang L, Zhang K, Shi Z, Zhang S. Langmuir 2017;33(42):11164–11169.
10. Li Y, Li Y, Pei A, Yan K, Sun Y, Wu C.-L, Joubert L.-M, Chin R, Koh A. L, Yu Y, Perrino J, Butz B, Chu S, Cui Y. Science 2017;358:506–510.
11. Wang J, Huang W, Pei A, Li Y, Shi F, Yu X, Cui Y. Nat. Energy 2019;4(8):664–670.
12. Cresce A, Eidson N, Schroeder M, Ma L, Howarth Y, Yang C, Ho J, Dillon R, Ding M, Bassett A, Stanzione J, Tom R, Soundappan T, Wang C, Xu K. J. Power Sources 2020;469:228378.
13. Kresse G, Furthmüller J. Phys. Rev. B 1996;54(16):11169–11186.
14. Perdew J. P, Burke K, Ernzerhof M. Phys. Rev. Lett 1996;77(18):3865–3868.
15. Grimme S, Antony J, Ehrlich S, Krieg H. J. Chem. Phys 2010;132(15):154104.
16. Zhang J, Cui C, Wang P.-F, Li Q, Chen L, Han F, Jin T, Liu S, Choudhary H, Raghavan S. R, Eidson N, von Cresce A, Ma L, Uddin J, Addison D, Yang C, Wang C. Energy Environ. Sci 2020;13:2878–2887.
17. Shang Y, Chen N, Li Y, Chen S, Lai J, Huang Y, Qu W, Wu F, Chen R. Adv. Mater 2020;32(40):2004017.
18. Mayer J, Giannuzzi L. A, Kamino T, Michael J. MRS Bulletin 2007;32:400–407.
19. Momma K, Izumi F. J. Appl. Cryst 2011;44(6):1272–1276.
20. Henkelman G, Arnaldsson A, Jónsson H. Comput. Mater. Sci 2006;36(3):354–360.

Article information Continued

Fig. 1

Schematic of the fabrication of the coin cell with a LiMn2O4 (LMO) cathode, a Li4Ti5O12 (LTO) anode, and a hybrid aqueous electrolyte.

Fig. 2

Optimized structures of Li4Ti5O12: (a) bulk, (b) side view, and (c) top view of the (200) plane surface of Li4Ti5O12. Green, blue, and red spheres represent Li, Ti, and O atoms, respectively. Li+ ions on the 16d sites are encircled; other Li+ ions occupy the 8a sites.

Fig. 3

Optimized structures of LiF: (a) bulk, (b) side view, and (c) top view of the (200) plane surface of LiF. Green and gray spheres represent Li and F atoms, respectively.

Fig. 4

Optimized structures of the LiF(200)/Li4Ti5O12(200) interface: (a) side view, and (b) top view. Green, blue, red, and gray spheres represent Li, Ti, O, and F atoms, respectively.

Fig. 5

Linear sweep voltammetry curve of the hybrid aqueous electrolyte showing the electrochemical stability window.

Fig. 6

Charge/discharge cycle performance at 0.5C.

Fig. 7

X-ray photoelectron spectra of the Li4Ti5O12 anode cycled 30 times.

Fig. 8

High-resolution transmission electron microscopy images of (a) the LTO electrode/SEI layer interface and (b) the corresponding fast Fourier-transform image obtained with beams parallel to the [011] direction.

Fig. 9

Charge density difference of the LiF(200)/Li4Ti5O12(200) interface, visualized using VESTA [19]. Areas of charge accumulation are indicated in yellow, whereas regions of charge depletion are shown in cyan. Isosurface = 0.0015 e Bohr−3. Green, blue, red, and gray spheres represent Li, Ti, O, and F atoms, respectively.

Table 1

Ion conductivity of the hybrid aqueous electrolyte at 23°C and its viscosity at 25°C and 40°C

Ion conductivity (mS cm−1) Viscosity (cP)

23°C 25°C 40°C
Hybrid aqueous electrolyte 0.51 330 130