Next Article in Journal
Influence of the Kinematic System on the Geometrical and Dimensional Accuracy of Holes in Drilling
Next Article in Special Issue
The Effect of Excessive Sulfate in the Li-Ion Battery Leachate on the Properties of Resynthesized Li[Ni1/3Co1/3Mn1/3]O2
Previous Article in Journal
Temperature Effect on Stability of Clamped–Clamped Composite Annular Plate with Damages
Previous Article in Special Issue
Metallic Material Selection and Prospective Surface Treatments for Proton Exchange Membrane Fuel Cell Bipolar Plates—A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Utilizing the Intrinsic Thermal Instability of Swedenborgite Structured YBaCo4O7+δ as an Opportunity for Material Engineering in Lithium-Ion Batteries by Er and Ga Co-Doping Processes

1
Department of Energy & Mineral Resources Engineering, Sejong University, Seoul 05006, Korea
2
HMC, Department of Nanotechnology and Advanced Materials Engineering, Sejong University, Seoul 05006, Korea
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Materials 2021, 14(16), 4565; https://doi.org/10.3390/ma14164565
Submission received: 12 July 2021 / Revised: 8 August 2021 / Accepted: 11 August 2021 / Published: 14 August 2021
(This article belongs to the Special Issue Advances in Manufacturing and Recycling of Battery Materials)

Abstract

:
We firstly introduce Er and Ga co-doped swedenborgite-structured YBaCo4O7+δ (YBC) as a cathode-active material in lithium-ion batteries (LIBs), aiming at converting the phase instability of YBC at high temperatures into a strategic way of enhancing the structural stability of layered cathode-active materials. Our recent publication reported that Y0.8Er0.2BaCo3.2Ga0.8O7+δ (YEBCG) showed excellent phase stability compared to YBC in a fuel cell operating condition. By contrast, the feasibility of the LiCoO2 (LCO) phase, which is derived from swedenborgite-structured YBC-based materials, as a LIB cathode-active material is investigated and the effects of co-doping with the Er and Ga ions on the structural and electrochemical properties of Li-intercalated YBC are systemically studied. The intrinsic swedenborgite structure of YBC-based materials with tetrahedrally coordinated Co2+/Co3+ are partially transformed into octahedrally coordinated Co3+, resulting in the formation of an LCO layered structure with a space group of R-3m that can work as a Li-ion migration path. Li-intercalated YEBCG (Li[YEBCG]) shows effective suppression of structural phase transition during cycling, leading to the enhancement of LIB performance in Coulombic efficiency, capacity retention, and rate capability. The galvanostatic intermittent titration technique, cyclic voltammetry and electrochemical impedance spectroscopy are performed to elucidate the enhanced phase stability of Li[YEBCG].

1. Introduction

Recent research for cathode-active materials in the field of lithium-ion batteries (LIBs) pursues high energy density with a long cycle life, which has been triggered by the worldwide growth of the electric vehicle (EV) market. Ni-rich layered cathode-active materials, such as LiNi1−x−yMnxCoyO2 (NMC) and LiNi1−x−yCoxAlyO2 (NCA), have received a great deal of attention due to their higher energy density with moderate cycling stability in the wide range of operating conditions compared to other types of cathode-active materials with olivine- (LiFePO4) or spinel- (LiMn2O4) structures [1]. The well-known LiCoO2 (LCO), which has the same layered crystal structure as NMC (space group: R-3m) and has been widely used mainly in portable IT devices and power tools in recent decades, became a minor option as EV battery cathode material owing to its limited practical capacity characteristics. Specifically, LCO undergoes severe structural distortion constraining the practical capacity when cycled above 4.2 V vs. Li/Li+ [2]. Thus, a proper cathode active material for EV requires a high Ni content in its chemical composition to satisfy high energy density despite the detrimental effect of the Ni ions on structural and thermal stability. Accordingly, many researchers have made a variety of efforts to find new dopants or surface coating compounds to suppress the severe structural degradation of Ni-rich layered cathode active materials during cycling, notwithstanding the role of Mn and Al species for structural stabilization in NMC and NCA, respectively [3,4,5,6,7,8,9,10,11,12,13,14]. This is mainly due to a specific phase transition region from H2 to H3 at around 4.2 V vs. Li/Li+, which leads to abrupt lattice contraction causing particle pulverization induced by microcracks [15,16].
On the other hand, Co-containing swedenborgite-structured YBaCo4O7+δ (YBC) materials with significantly large oxygen-storage capacity are considered to be promising catalysts in solid oxide fuel cells (SOFC) working at intermediate temperatures [17,18]. It is well known that the swedenborgite structure (space group: P63mc) of YBC with two layers of CoO4 tetrahedral coordinates suffers from severe phase instability by thermal decomposition at 600–800 °C due to the preference of Co ions for octahedral coordination, which makes its application in SOFC difficult [19,20]. Manthiram et al. reported that YBC doped with the optimum content of Ga, which substitutes for Co sites, can effectively overcome this phase instability at high temperatures of 600–800 °C [21]. Our recent work reported that Er and Ga co-doped YBC oxide (YEBCG) showed excellent phase stability compared to YBC, presenting long-term durability under reversible protonic ceramic cell conditions [22]. Meanwhile, there are a few reports adopting Er or Ga as dopants for cathode-active materials in LIBs. While Er-doped LiFePO4 and LiNi0.5Mn1.5O4 showed improved cycling stability [23,24], Ga-doped layered LiN0.6Co0.2Mn0.2O2 presented enhanced electrochemical performance and thermal stability, which is in line with the Manthiram group’s result based on the YBC material related to the improved phase stability at high temperatures [21,25].
Inspired by these results, we herein introduced Er and Ga co-doped swedenborgite-structured YBC as a cathode-active material in LIBs, aiming at converting the issue of phase instability of YBC at high temperatures into a challenge to explore a way towards enhancing the structural stability of layered cathode-active materials. The high temperature (800 °C) calcination process for Li-ion intercalation into the swedenborgite structure did not decompose the intrinsic swedenborgite structure of YBC and YEBCG, and forms merely a few secondary phases induced by partial thermal decomposition and a layered LCO phase, which locates Co in octahedral coordinates. The feasibility of the LCO phase, which is derived from the swedenborgite-structured YBC-based materials, as a layered cathode-active material for the LIBs was investigated, and the effects of co-doping with the Er and Ga ions on the structural and electrochemical properties of Li-intercalated YBC were studied in detail.

2. Experimental Section

2.1. Synthesis and Characterization of Materials

To prepare YBC and YEBCG powders, stoichiometric Ba(NO3)2 (99.95%, metal basis Alfa Aesar), Er(NO3)3·5H2O (99.9%, Alfa Aesar), Y(NO3)3·6H2O (99.9%, Alfa Aesar), (NH4)2Ce(NO3)6 (99.99%, Alfa Aesar), and Ga(NO3)3·xH2O (99.9%, metal basis, Alfa Aesar) were dissolved in distilled water with glycine as a combusting fuel and then heated on a heating plate (MSH-20D, DAIHAN Scientific Co., Ltd., Wonju, Korea) at 350 °C until the metal nitrates converted into black ashes. After the combustion, the ashes were ground with a mortar and then calcined at 1000 °C for 12 h under an air atmosphere to obtain a single-phased crystalline [22]. In order to identify the crystal structure of the YBC-based materials, an X-ray diffraction (XRD) technique (X’Pert, PANalytical, Cu Kα radiation, Almelo, The Netherlands) was carried out with a step size of 0.026° in a 2θ range from 10 to 80°. Utilizing the XRD data, the FullProf software was used for Rietveld refinement to yield lattice parameters. The morphological characterization of the YBC-based materials was performed using a field emission scanning electron microscope (FE-SEM, SU-8010, Hitachi Ltd., Tokyo, Japan) with energy-dispersive X-ray spectroscopy (EDS). To measure the Brunauer-Emmett-Teller (BET) surface area, nitrogen gas was used for an adsorption/desorption method (BELSORP-max, BEL Inc., Leitchfield, KY, USA) at 77 K to remove residual impurities and moisture with an average mass of 0.54 g for each specimen. For the sake of Li-intercalation, the synthesized YBC and YEBCG were mixed with LiOH‧H2O in a molar ratio of 1:1.1, followed by calcination under air atmosphere for 10 h at 800 °C. The Li-intercalated YBC and YEBCG oxides are referred to as Li[YBC] and Li[YEBCG], respectively.

2.2. Electrochemical Analysis

For a cathode slurry, the Li[YBC] and Li[YEBCG] cathode active materials, carbon black (Super-P) as a conducting material, and polyvinylidene fluoride (KF 1100) as a binder in a weight ratio of 80:10:10 were thoroughly mixed in a N-methylpyrrolidinone (NMP) solution with 30% solid content and casted on Al foil as a current collector. Electrochemical properties were investigated using CR2032-type coin cells, which were fabricated in a moisture-controlled glove box under an argon atmosphere. While 1 M LiPF6 dissolved in a mixture of ethyl methyl carbonate and ethylene carbonate (2:1, v/v) was used as an electrolyte, Li metal foil and polyethylene film were used as anode and separator, respectively. Cycling stability was assessed by cycling the cells at 0.1 C (46 mA g−1 as 1 C) in different potential ranges of 2.0 to 4.5 V for 50 cycles and 2.5 to 4.3 V for 100 cycles, while a rate capability test was conducted in the potential range from 2.0 to 4.5 V in various C-rate conditions using a battery cycler (WBCS3000L, WonAtech Ltd., Seoul, Korea) at 25 °C. For activating the cells, cycling at 0.1 C during initial two cycles was conducted prior to the electrochemical tests as a formation step. Electrochemical impedance spectroscopy (EIS) and cyclic voltammetry (CV) were performed using a potentiostat system (versaSTAT 3, AMETEK Inc., Berwyn, PA, USA) to compare the change in internal cell resistance and the degree of overpotentials during cycling, respectively. A galvanostatic intermittent titration technique (GITT) was conducted to calculate Li-ion diffusion coefficients. All the values of potentials were based on Li/Li+ in this study unless otherwise mentioned.

3. Results and Discussion

The XRD patterns of swedenborgite-structured YBC and YEBCG before and after the Li-ion intercalating calcination process at 800 °C are presented in Figure 1. It is known that BaCoO3−δ and Y2O3 secondary phases can form readily due to oxidative thermal decomposition from the intrinsic YBC structure above 600 °C [26]. This structural degradation of Li[YEBCG] induced by the thermal decomposition was mitigated when Er and Ga were co-doped in the YBC structure. As seen in Figure 1, the peak intensity of secondary phases was reduced for Li[YEBCG], indicating the better structural stability at the oxidative high temperature than Li[YBC]. A LiCoO2 phase allowing the (de)intercalation of Li ions was also newly formed (marked as green-colored symbols in the figure), which might be ascribed to the evolution of Co ions located at octahedral coordinates during the calcination process. Likewise, the Co ions in BaCoO3 (marked as gray-colored symbols in the figure), which is a main decomposition product of YBC, have octahedral coordination [27], whereas the Co ions in YBC have a tetrahedral coordination. Meanwhile, Table 1 lists the crystal lattice parameters of pristine YBC and YEBCG, showing the increased lattice parameters of a- and c-axes with larger lattice volume for YEBCG compared to YBC, which is consistent with the previous literature [22].
To elucidate the relationship between changes in structural phase and morphology before and after the Li-intercalation process, a FE-SEM analysis was performed. Figure 2 shows the surface morphology of YBC, Li[YBC], YEBCG, and Li[YEBCG]. Compared to the pristine particles, the Li-intercalated Li[YBC] and Li[YEBCG] samples exhibited a negligible change in morphology. It can be concluded that the aforementioned phase deformation induced by the oxidative thermal decomposition does not affect the morphological change from their original shapes. Meanwhile, the elemental mapping data of Li[YBC] and Li[YEBCG] are presented, as shown in Figure 3a for Li[YBC] and Figure 3b for Li[YEBCG], respectively, indicating that all of constituent elements were evenly distributed.
Figure 4a,b display the initial charge (delithiation) and discharge (lithiation) curves of Li[YBC] and Li[YEBCG] cathode active materials. Li[YEBCG] showed increased charge/discharge capacities with enhanced Coulombic efficiency compared to Li[YBC]. A differential capacity analysis was performed in order to look into the phase transition behavior during the initial cycle in a potential range of 2.0–4.5 V, as shown in Figure 4c,d, which was derived from the initial charge/discharge curves. Two minor peaks between 4.05 and 4.25 V were observed in Li[YBC] besides the major peaks at around 3.9 V, whereas the phase transition was effectively suppressed in the case of Li[YEBCG], indicating the enhanced structural stability. These electrochemical behaviors of Li[YBC] and Li[YEBCG] are consistent with the typical phase transition behavior of the LCO cathode active materials in LIBs [2]. Because these phase transitions during a repetitive cycling, specifically above 4.2 V, can bring about the deterioration of LIB performance, the mitigation of the phase transition is a crucial issue for layered cathode active materials [15]. Consequently, LIB performance including the cycle life and the rate capability of Li[YEBCG] is expected to outperform Li[YBC] considering the structural and electrochemical features of enlarged lattice volume and suppressed phase stability.
Figure 5 presents the cycle performance of the cathode active materials for 50 cycles at 0.1 C in a potential range of 2.0–4.5 V. The capacity retention of Li[YBC] was drastically aggravated right after the initial cycle and reached less than 10% after about 20 cycles. By contrast, Li[YEBCG] showed superior capacity retention with a relatively stable feature of Coulombic efficiency over the entire cycling process. After 50 cycles, Li[YEBCG] had a capacity retention of 50%, which was about ten times higher than Li[YBC]. The corresponding voltage profiles of Li[YBC] (see Figure 5b) clearly showed rapid capacity fading with increasing overpotentials during cycling, while Li[YEBCG] presented a feature of mitigated capacity fading (see Figure 5c), indicating the positive effects of co-doping with Er and Ga on the suppression of the undesired phase transition as aforementioned in Figure 4. Long-term cycle performance was additionally conducted in the potential range of 2.5–4.3 V in Figure S1, which is closer to the practical operating condition of LIBs including a constant voltage charging step with a moderate charging cut-off potential. The long-term cycle performance showed relatively enhanced cycling stability compared to the 4.5 V cut-off condition resulting from a limited (de)lithiation range, which led to decreased discharge capacities for both samples. In general, a higher upper cut-off potential shows higher capacities at the expense of cycling stability [28].
In order to electrochemically elucidate a specific reason for the improved capacity retention of Li[YEBCG], the EIS was conducted in a frequency range from 0.01 Hz to 1 MHz with an alternating voltage amplitude of 15 mV. The EIS technique in LIBs is considered as one of the most attractive analytical tools that can separate and quantify cell resistance in-situ, avoiding any impact of moisture or oxygen on sensitive samples [29]. Figure 6 shows the resulting Nyquist plots of Li[YBC] and Li[YEBCG] with an increasing cycle number under a pseudo-equilibrium state for charged cells. Li[YBC] has shown extremely larger impedance growth than Li[YEBCG] after the formation step. Noticeably, the first semicircles in a high frequency region that are related to the film resistance induced by the formation of solid electrolyte interphase (SEI) between electrodes and electrolytes gradually shrank with an increased cycle number from 10 to 50 in both samples (see the insets in the figures). It is hard to intuitively interpret the abnormal phenomenon about the gradual decrease of film resistance, which is contrary to other relevant research [30,31]. In the meantime, the second semicircles in a low frequency region regarding charge transfer resistance (Rct) indicated incremental trends in both samples during cycling due to the structural degradation of the cathode active materials. In addition, this might correlate with the intrinsically weak adhesion property of YBC-based materials to current collectors (see Figure S2). Therefore, it can be concluded that the poor capacity retention of Li[YBC] was caused by severe impedance growth in the cell, while the doping with Er and Ga is deemed to be a crucial factor to reinforce the host structure and facilitate the charge transfer process in the Li[YEBCG] structure.
Through a series of EIS and CV measurements under various electrochemical conditions, the superior electrochemical performance of Li[YEBCG] compared to that of Li[YBC] was clearly confirmed again. As seen in Figure 7a, the EIS results obtained at equilibrium states before the following CV tests showed that Li[YBC] has a relatively lowered Rct value than that of Li[YEBCG], which leads to a slightly reduced potential intervals (ΔV) between main anodic and cathodic peaks compared with Li[YEBCG] as shown in Figure 7b. Although the ΔV generally indicates the reversibility of Li-ion (de)intercalation, the obvious feature of the suppressed phase transition of Li[YEBCG] during the cycling at 0.1 mV s−1 positively affected the repetitive CV measurements at 0.5 mV s−1 as shown in Figure 7c,d. Li[YBC] presented a different electrode polarization induced by severe overpotentials after 10 cycles leading to the increase in ΔV, whereas the degree of overpotentials for Li[YEBCG] was slightly changed after 10 cycles representing the enhancement of structural stability of Li[YEBCG] by the Er and Ga co-doping process. It is worth noting that the Rct value of Li[YEBCG] measured after the formation step became smaller than Li[YBC] (see Figure 6).
The EIS and CV measurements were additionally conducted at a different temperature of 60 °C, as displayed in Figure 7e,f, respectively. Compared with the results at 25 °C, the Rct values have drastically decreased in both samples, which might be attributed to the effect of elevated temperature facilitating the migration of Li+ ions in the bulk structure. In the CV data, the overpotentials of both samples were also reduced. Interestingly, the feature of severe phase transition of Li[YBC] during the CV measurement at 0.1 mV s−1 completely disappeared, which could lead to the improvement of structural stability during the cycling. Although the capacity retention of Li[YEBCG] at 60 °C was indeed improved in comparison with that at 25 °C, the drastic aggravation of capacity retention after the initial cycle was still a problematic issue (see Figure S3).
The rate capability tests of Li[YBC] and Li[YEBCG] were performed to look into the kinetic behavior during reversible (de)lithiation reactions at different discharging C-rates with a fixed charging C-rate of 0.1 C, as shown in Figure 8a,b. Whereas the overpotentials of Li[YBC] rapidly increased along with the decrease in operating voltage as the C-rate increases, Li[YEBCG] presented a much better rate performance than Li[YBC]. Moreover, given that the decay tendency in the charge capacity of Li[YEBCG] was apparently mitigated compared with Li[YBC], it can be understood that the intrinsic swedenborgite structure of Li[YEBCG] is effectively stabilized by the co-doping with Er and Ga. Figure 8c shows the relative capability of the samples with respect to discharge capacities at various C-rates compared to the discharge capacity at 0.1 C. Li[YEBCG] had superior relative capability to that of Li[YBC] as the C-rate increases, also suggesting that the Er and Ga co-doped host structure of YEBCG is favorable for Li ions to migrate into the bulk structure.
Lastly, to compare the Li-ion diffusion coefficients (DLi+) between the samples, a GITT analysis was carried out as shown in Figure 9. GITT curves shown in Figure 9a were obtained during the second charging process, and a single titration step was described in Figure 9b, where ∆Es (V) is the voltage change between steady states and ∆Eτ (V) is the total change of cell voltage in a single titration step [32]. Fick’s second law is utilized to calculate the DLi+ as below [33]:
D Li + = 4 π ( m V M S ) 2 ( Δ E S τ ( d E τ d τ ) ) 2   ( τ L 2 D Li + )
As seen in Figure 9c, under the assumption that E vs. τ shows a straight-line behavior during the titration step, Equation (1) can be simplified into Equation (2) [34,35],
D Li + = 4 π τ ( m V M S ) 2 ( Δ E S Δ E τ ) 2 ( τ L 2 D Li + )
where m is the mass (g), V is the molar volume (cm3 mol−1), and M is the molecular weight (g mol−1) of cathode active materials. S is the contact area (cm2) between electrodes and electrolytes, and L is the Li-ion diffusion length (cm).
The DLi+ values of samples calculated using Equation (2) were plotted as a function of cell potential, showing relatively higher DLi+ values of Li[YEBCG] compared to Li[YBC] over the entire potential range (Figure 9d). It can be explained that Li[YEBCG] has better kinetic behavior for the (de)intercalation of Li ions compared to Li[YBC] regardless of C-rate conditions due to the higher DLi+. Therefore, the comprehensively improved electrochemical properties of Li[YEBCG] are clearly attributed to the effect of cationic substitution of Er and Ga that stabilize the host structure from undesired phase transition.

4. Conclusions

We synthesized Er and Ga co-doped swedenborgite-structured YBC, which was utilized as a cathode-active material to investigate the potential enhancement of phase stability. The peak intensity of secondary phases in the XRD patterns was mitigated in the case of Li[YEBCG], indicating the better structural stability at the oxidative high temperature compared with Li[YBC]. A LiCoO2 phase allowing the (de)intercalation of Li ions was also newly formed, which might result from the evolution of Co ions located at octahedral coordinates during the calcination process. The partial thermal deformation resulting from oxidative calcination barely affected the morphologies and their elemental distribution of YBC and YEBCG. Regarding the LIB performance, Li[YEBCG] exhibited superior capacity retention with enhanced charge/discharge capacities compared to Li[YBC], showing the mitigated feature of phase transition at around 4.2 V. The poor capacity retention of Li[YBC] was ascribed to severe impedance growth during cycling, while the doping with Er and Ga was considered to be a crucial factor to reinforce the host structure and facilitate the charge transfer process in the Li[YEBCG] structure, leading to the effective suppression of impedance growth. Whereas the overpotentials of Li[YBC] rapidly increased along with the decrease in operating voltage as the C-rate increased, Li[YEBCG] showed much better rate performance than Li[YBC]. The calculated DLi+ of samples showed the relatively higher DLi+ values of Li[YEBCG] compared to Li[YEBCG] over the entire potential range, explaining that Li[YEBCG] has better kinetic behavior for the (de)intercalation of Li ions than Li[YBC] regardless of C-rate conditions. Consequently, the comprehensively improved electrochemical properties of Li[YEBCG] are clearly attributed to the enhanced phase stability induced by the stabilizing effect of Er and Ga co-doping on the host structure from the undesired phase transition. Future work will include the influence of Er and Ga co-doping in more conventional LIB cathode active materials on the mitigation of structural degradation during cycling.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/ma14164565/s1, Figure S1: Initial charge/discharge curves and long-term cycle performance of Li[YBC] and Li[YEBCG] in a potential range of 2.5–4.3 V, Figure S2: Comparative pictures of (a) Li[YBC] and (b) Li[YEBCG] after the cycling test, Figure S3: Cycle performance of Li[YBC] at different temperatures of 25 and 60 °C.

Author Contributions

Conceptualization, S.P. and K.P.; methodology, S.P. and J.-S.S.; validation, G.K. and W.K.; investigation, G.K. and W.K.; writing—original draft preparation, S.P. and K.P.; writing—review and editing, K.K.; supervision, K.K. and J.-Y.P.; project administration, J.-Y.P. and J.-S.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Basic Science Research Program and the Next-generation Engineering Researchers Development Program (2019H1D8A2106002) through the National Research Foundation of Korea (NRF) funded by the Ministry of Education (2020R1A6A1A03038540, 2021R1I1A1A01044063), the Ministry of Science and ICT (2020R1F1A1053911), and the Korea Ministry of Environment (MOE) as ‘Graduate School specialized in Climate Change’.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, W.; Erickson, E.M.; Manthiram, A. High-Nickel Layered Oxide Cathodes for Lithium-Based Automotive Batteries. Nat. Energy 2020, 5, 26–34. [Google Scholar] [CrossRef]
  2. Xia, H.; Lu, L.; Meng, Y.S.; Ceder, G. Phase Transitions and High-Voltage Electrochemical Behavior of LiCoO2 Thin Films Grown by Pulsed Laser Deposition. J. Electrochem. Soc. 2007, 154, A337–A342. [Google Scholar] [CrossRef]
  3. Weigel, T.; Schipper, F.; Erickson, E.M.; Susai, F.A.; Markovsky, B.; Aurbach, D. Structural and Electrochemical Aspects of LiNi0.8Co0.1Mn0.1O2 Cathode Materials Doped by Various Cations. ACS Energy Lett. 2019, 4, 508–516. [Google Scholar] [CrossRef]
  4. Trease, N.M.; Seymour, I.D.; Radin, M.D.; Liu, H.; Hy, S.; Chernova, N.; Parikh, P.; Devaraj, A.; Wiaderek, K.M.; Chupas, P.J.; et al. Identifying the Distribution of Al3+ in LiNi0.8Co0.15Al0.05O2. Chem. Mater. 2016, 28, 8170–8180. [Google Scholar] [CrossRef] [Green Version]
  5. Sivaprakash, S.; Majumder, S.B. Understanding the Role of Zr4+ Cation in Improving the Cycleability of LiNi0.8Co0.15Zr0.05O2 Cathodes for Li Ion Rechargeable Batteries. J. Alloys Compd. 2009, 479, 561–568. [Google Scholar] [CrossRef]
  6. Chen, T.; Li, X.; Wang, H.; Yan, X.; Wang, L.; Deng, B.; Qu, M. The Effect of Gradient Boracic Polyanion-Doping on Structure, Morphology, and Cycling Performance of Ni-rich LiNi0.8Co0.15Al0.05O2 Cathode Material. J. Power Sources 2018, 374, 1–11. [Google Scholar] [CrossRef]
  7. Woo, S.U.; Park, B.C.; Yoon, C.S.; Myung, S.T.; Prakash, J.; Sun, Y.K. Improvement of Electrochemical Performances of LiNi0.8Co0.1Mn0.1O2 Cathode Materials by Fluorine Substitution. J. Electrochem. Soc. 2007, 154, A649–A655. [Google Scholar] [CrossRef]
  8. Becker, D.; Borner, M.; Nolle, R.; Diehl, M.; Klein, S.; Rodehorst, U.C.; Placke, T. Surface Modification of Ni-rich LiNi0.8Co0.1Mn0.1O2 Cathode Material by Tungsten Oxide Coating for Improved Electrochemical Performance in Lithium Ion Batteries. ACS Appl. Mater. Interfaces 2019, 11, 18404–18414. [Google Scholar] [CrossRef]
  9. Konishi, H.; Yoshikawa, M.; Hirano, T. The Effect of Thermal Stability for High Ni-Content Layer-Structured Cathode Materials, LiNi0.8Mn0.1−xCo0.1MoxO2 (x = 0, 0.02, 0.04). J. Power Sources 2013, 244, 23–28. [Google Scholar] [CrossRef]
  10. Chen, M.; Zhao, E.; Chen, D.; Wu, M.; Han, S.; Huang, Q.; Yang, L.; Xiao, X.; Hu, Z. Decreasing Li/Ni Disorder and Improving the Electrochemical Performances of Ni-rich LiNi0.8Co0.1Mn0.1O2 by Ca Doping. Inorg. Chem. 2017, 56, 8355–8362. [Google Scholar] [CrossRef]
  11. Song, B.; Li, W.; Oh, S.M.; Manthiram, A. Long-Life Nickel-Rich Layered Oxide Cathodes with a Uniform Li2ZrO3 Surface Coating for Lithium-Ion Batteries. ACS Appl. Mater. Interfaces 2017, 9, 9718–9725. [Google Scholar] [CrossRef]
  12. Schipper, F.; Bouzaglo, H.; Dixit, M.; Erickson, E.M.; Weigel, T.; Talianker, M.; Grinblat, J.; Erk, C. From Surface ZrO2 Coating to Bulk Zr Doping by High Temperature Annealing of Nickel-Rich Lithiated Oxides and Their Enhanced Electrochemical Performance in Lithium Ion Batteries. Adv. Energy Mater. 2018, 8, 1701682. [Google Scholar] [CrossRef]
  13. Min, K.; Park, K.; Park, S.Y.; Seo, S.W.; Choi, B.; Cho, E. Improved Electrochemical Properties of LiNi0.91Co0.06Mn0.03O2 Cathode Material via Li-Reactive Coating with Metal Phosphates. Sci. Rep. 2017, 7, 7151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Kim, U.; Myung, S.T.; Yoon, C.S.; Sun, Y.K. Extending the Battery Life Using an Al-Doped Li[Ni0.76Co0.09Mn0.15]O2 Cathode with Concentration Gradients for Lithium Ion Batteries. ACS Energy Lett. 2017, 2, 1848–1854. [Google Scholar] [CrossRef]
  15. Ryu, H.H.; Park, K.J.; Yoon, C.S.; Sun, Y.K. Capacity Fading of Ni-Rich Li[NixCoyMn1−x−y]O2 (0.6 ≤ x ≤ 0.95) Cathodes for High-Energy-Density Lithium-Ion Batteries: Bulk or Surface Degradation. Chem. Mater. 2018, 30, 1155–1163. [Google Scholar] [CrossRef]
  16. Lim, J.M.; Hwang, T.; Kim, D.; Park, M.S.; Cho, K.; Cho, M. Intrinsic Origins of Crack Generation in Ni-Rich LiNi0.8Co0.1Mn0.1O2 Layered Oxide Cathode Material. Sci. Rep. 2017, 7, 39669. [Google Scholar] [CrossRef]
  17. Kim, J.H.; Kim, Y.N.; Cho, S.M.; Wang, H.; Manthiram, A. Electrochemical Characterization of YBaCo3ZnO7 + Gd0.2Ce0.8O1.9 Composite Cathodes for Intermediate Temperature Solid Oxide Fuel Cells. Electrochim. Acta 2010, 55, 5312–5317. [Google Scholar] [CrossRef]
  18. Medvedev, D.; Lyagaeva, J.; Vdovin, G.; Beresnev, S.; Demin, A.; Tsiakaras, P. A Tape Calendering Method as an Effective Way for the Preparation of Proton Ceramic Fuel Cells with Enhanced Performance. Electrochim. Acta 2016, 210, 681–688. [Google Scholar] [CrossRef]
  19. Bhat, M.A.; Zargar, R.A.; Modi, A.; Arora, M.; Gaur, N.K. Structural, Electrical and Magnetic Features of Kagomé YBaCo4O7 System. Mater. Sci. Pol. 2016, 34, 786–793. [Google Scholar] [CrossRef] [Green Version]
  20. Tsvetkov, D.S.; Pralong, V.; Tsvetkova, N.S.; Zuev, A.Y. Oxygen Content and Thermodynamic Stability of YBaCo4O7±δ. Solid State Ion. 2015, 278, 1–4. [Google Scholar] [CrossRef]
  21. Lai, K.Y.; Manthiram, A. Phase Stability, Oxygen-Storage Capability, and Electrocatalytic Activity in Solid Oxide Fuel Cells of (Y, In, Ca) BaCo4–yGayO7+δ. Chem. Mater. 2016, 28, 9077–9087. [Google Scholar] [CrossRef]
  22. Shin, J.-S.; Park, H.; Park, K.; Saqib, M.; Jo, M.; Kim, J.H.; Lim, H.-T.; Kim, M.; Kim, J.; Park, J.-Y. Activity of Layered Swedenborgite Structured Y0.8Er0.2BaCo3.2Ga0.8O7+δ For Oxygen Electrode Reactions in at Intermediate Temperature Reversible Ceramic Cells. J. Mater. Chem. A 2021, 9, 607–621. [Google Scholar] [CrossRef]
  23. Goktepe, H.; Sahan, H.; Ulgen, A.; Patat, S. Synthesis and Electrochemical Properties of Carbon-Mixed LiEr0.02Fe0.98PO4 Cathode Material for Lithium-Ion Batteries. J. Mater. Sci. Technol. 2011, 27, 861–864. [Google Scholar] [CrossRef]
  24. Liu, S.; Zhao, H.; Tan, M.; Hu, Y.; Shu, X.; Zhang, M.; Chen, B.; Liu, X. Er-Doped LiNi0.5Mn1.5O4 Cathode Material with Enhanced Cycling Stability for Lithium-Ion Batteries. Materials 2017, 10, 859. [Google Scholar] [CrossRef] [Green Version]
  25. Liu, Z.; Li, J.; Zhu, M.; Wang, L.; Kang, Y.; Dang, Z.; Yan, J. Enhanced Structural Stability and Electrochemical Performance of LiNi0.6Co0.2Mn0.2O2 Cathode Materials by Ga Doping. Materials 2021, 14, 1816. [Google Scholar] [CrossRef]
  26. Parkkima, O.; Karppinen, M. The YBaCo4O7+δ-Based Functional Oxide Material Family: A Review. Eur. J. Inorg. Chem. 2014, 4056–4067. [Google Scholar] [CrossRef]
  27. Kim, J.-H.; Manthiram, A. Low Thermal Expansion RBa(Co,M)4O7 Cathode Materials Based on Tetrahedral-Site Cobalt Ions for Solid Oxide Fuel Cells. Chem. Mater. 2010, 22, 822–831. [Google Scholar] [CrossRef]
  28. Kim, J.H.; Park, K.J.; Kim, S.J.; Yoon, C.S.; Sun, Y.K. A Method of Increasing the Energy Density of Layered Ni-Rich Li[Ni1−2xCoxMnx]O2 cathodes (x = 0.05, 0.1, 0.2). J. Mater. Chem. A 2019, 7, 2694–2701. [Google Scholar] [CrossRef]
  29. Choi, W.; Shin, H.C.; Kim, J.M.; Choi, J.Y.; Yoon, W.S. Modeling and Applications of Electrochemical Impedance Spectroscopy (EIS) for Lithium-Ion Batteries. J. Electrochem. Sci. Technol. 2020, 11, 1–13. [Google Scholar] [CrossRef] [Green Version]
  30. Beak, M.; Park, S.; Kim, S.; Park, J.; Jeong, S.; Thirumalraj, B.; Jeong, G.; Kim, T.; Kwon, K. Effect of Na from the Leachate of Spent Li-Ion Batteries on the Properties of Resynthesized Li-Ion Battery Cathodes. J. Alloys Compd. 2021, 873, 159808–159816. [Google Scholar] [CrossRef]
  31. Jeong, S.; Park, S.; Beak, M.; Park, J.; Shon, J.-S.; Kwon, K. Effect of Residual Trace Amounts of Fe and Al in Li[Ni1/3Mn1/3Co1/3]O2 Cathode Active Material for the Sustainable Recycling of Lithium-Ion Batteries. Materials 2021, 14, 2464. [Google Scholar] [CrossRef] [PubMed]
  32. Shaju, K.M.; Rao, G.V.S.; Chowdari, B.V.R. Electrochemical Kinetic Studies of Li-ion in O2-Structured Li2/3(Ni1/3Mn2/3)O2 and Li(2/3)+x(Ni1/3Mn2/3)O2 by EIS and GITT. J. Electrochem. Soc. 2003, 150, A1–A13. [Google Scholar] [CrossRef]
  33. Zheng, J.M.; Shi, W.; Gu, M.; Xiao, J.; Zuo, P.J.; Wang, C.M.; Zhang, J.G. Electrochemical Kinetics and Performance of Layered Composite Cathode Material Li[Li0.2Ni0.2Mn0.6]O2. J. Electrochem. Soc. 2013, 160, A2212–A2219. [Google Scholar] [CrossRef]
  34. Peng, F.W.; Mu, D.Y.; Li, R.H.; Liu, Y.L.; Ji, Y.P.; Dai, C.S.; Ding, F. Impurity Removal with Highly Selective and Efficient Methods and the Recycling of Transition Metals from Spent Lithium-Ion Batteries. RSC Adv. 2019, 9, 21922–21930. [Google Scholar] [CrossRef] [Green Version]
  35. Shaju, K.M.; Rao, G.V.S.; Chowdari, B.V.R. EIS and GITT Studies on Oxide Cathodes, O2-Li(2/3)+x(Co0.15Mn0.85)O2 (x = 0 and 1/3). Electrochim. Acta 2003, 48, 2691–2703. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of pristine YBC and YEBCG, and their Li-intercalated forms of Li[YBC] and Li[YEBCG].
Figure 1. XRD patterns of pristine YBC and YEBCG, and their Li-intercalated forms of Li[YBC] and Li[YEBCG].
Materials 14 04565 g001
Figure 2. FE-SEM images of (a) YBC, (b) Li[YBC], (c) YEBCG, and (d) Li[YEBCG].
Figure 2. FE-SEM images of (a) YBC, (b) Li[YBC], (c) YEBCG, and (d) Li[YEBCG].
Materials 14 04565 g002
Figure 3. Elemental mapping images of (a) Li[YBC] and (b) Li[YEBCG] analyzed by EDS.
Figure 3. Elemental mapping images of (a) Li[YBC] and (b) Li[YEBCG] analyzed by EDS.
Materials 14 04565 g003
Figure 4. (a,b) Initial charge/discharge and (c,d) differential capacity curves of Li[YBC] and Li[YEBCG], respectively. A distinct phase transition region (red-dotted circles in (c,d)) in the potential around 4.2 V was displayed with insets of the corresponding voltage profile. (C.E.: Coulombic efficiency).
Figure 4. (a,b) Initial charge/discharge and (c,d) differential capacity curves of Li[YBC] and Li[YEBCG], respectively. A distinct phase transition region (red-dotted circles in (c,d)) in the potential around 4.2 V was displayed with insets of the corresponding voltage profile. (C.E.: Coulombic efficiency).
Materials 14 04565 g004
Figure 5. (a) Capacity retention at 0.1 C with Coulombic efficiency, and corresponding voltage profiles of (b) YBC and (c) YEBCG during 50 cycles.
Figure 5. (a) Capacity retention at 0.1 C with Coulombic efficiency, and corresponding voltage profiles of (b) YBC and (c) YEBCG during 50 cycles.
Materials 14 04565 g005
Figure 6. Nyquist plots of (a) Li[YBC] and (b) Li[YEBCG] with increasing cycle numbers.
Figure 6. Nyquist plots of (a) Li[YBC] and (b) Li[YEBCG] with increasing cycle numbers.
Materials 14 04565 g006
Figure 7. Nyquist plots and CV curves of Li[YBC] and Li[YEBCG] under various experimental conditions. (a,e) EIS data before the CV tests at equilibrium states, (b,f) initial CV curves at 0.1 mV s−1, and (c,d) CV curves during the 1st and 10th cycles at 0.5 mV s−1, where (ad) were tested at 25 °C and (e,f) at 60 °C.
Figure 7. Nyquist plots and CV curves of Li[YBC] and Li[YEBCG] under various experimental conditions. (a,e) EIS data before the CV tests at equilibrium states, (b,f) initial CV curves at 0.1 mV s−1, and (c,d) CV curves during the 1st and 10th cycles at 0.5 mV s−1, where (ad) were tested at 25 °C and (e,f) at 60 °C.
Materials 14 04565 g007
Figure 8. Rate capability of (a) Li[YBC] and (b) Li[YEBCG], and (c) their relative capability at various discharging C-rates from 0.1 to 2 C with a fixed charging C-rate of 0.1 C.
Figure 8. Rate capability of (a) Li[YBC] and (b) Li[YEBCG], and (c) their relative capability at various discharging C-rates from 0.1 to 2 C with a fixed charging C-rate of 0.1 C.
Materials 14 04565 g008
Figure 9. (a) GITT curves of Li[YBC] and Li[YEBCG] during the second charge process at a current density of 4.6 mA g−1. (b) GITT curve at a single titration step of Li[YEBCG] with (c) corresponding linear behavior in the relationship of E vs. τ1/2. (d) Li-ion diffusion coefficients of Li[YBC] and Li[YEBCG] calculated from the GITT curves as a function of the cell potential.
Figure 9. (a) GITT curves of Li[YBC] and Li[YEBCG] during the second charge process at a current density of 4.6 mA g−1. (b) GITT curve at a single titration step of Li[YEBCG] with (c) corresponding linear behavior in the relationship of E vs. τ1/2. (d) Li-ion diffusion coefficients of Li[YBC] and Li[YEBCG] calculated from the GITT curves as a function of the cell potential.
Materials 14 04565 g009
Table 1. Crystallographic parameters of YBC and YEBCG calculated by Rietveld refinement.
Table 1. Crystallographic parameters of YBC and YEBCG calculated by Rietveld refinement.
YBCYEBCG
a-axis (Å)6.29936.3097
c-axis (Å)10.253410.2674
Volume (Å3)352.3540354.0059
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Park, S.; Park, K.; Shin, J.-S.; Ko, G.; Kim, W.; Park, J.-Y.; Kwon, K. Utilizing the Intrinsic Thermal Instability of Swedenborgite Structured YBaCo4O7+δ as an Opportunity for Material Engineering in Lithium-Ion Batteries by Er and Ga Co-Doping Processes. Materials 2021, 14, 4565. https://doi.org/10.3390/ma14164565

AMA Style

Park S, Park K, Shin J-S, Ko G, Kim W, Park J-Y, Kwon K. Utilizing the Intrinsic Thermal Instability of Swedenborgite Structured YBaCo4O7+δ as an Opportunity for Material Engineering in Lithium-Ion Batteries by Er and Ga Co-Doping Processes. Materials. 2021; 14(16):4565. https://doi.org/10.3390/ma14164565

Chicago/Turabian Style

Park, Sanghyuk, Kwangho Park, Ji-Seop Shin, Gyeongbin Ko, Wooseok Kim, Jun-Young Park, and Kyungjung Kwon. 2021. "Utilizing the Intrinsic Thermal Instability of Swedenborgite Structured YBaCo4O7+δ as an Opportunity for Material Engineering in Lithium-Ion Batteries by Er and Ga Co-Doping Processes" Materials 14, no. 16: 4565. https://doi.org/10.3390/ma14164565

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