Next Article in Journal
Environment-Friendly Power Scheduling Based on Deep Contextual Reinforcement Learning
Previous Article in Journal
Performance Investigation on a Double-Slope Passive Solar Desalination System Targeting towards Sustainable Development of Oman
Previous Article in Special Issue
A Review on Battery Thermal Management for New Energy Vehicles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Coprecipitation Synthesis and Impedance Studies on Electrode Interface Characteristics of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 Cathode Material

1
China Automotive Engineering Research Institute Co., Ltd., Chongqing 401122, China
2
Technology Innovation Center of New Energy Vehicle Digital Supervision, Technology and Application for State Market Regulation, Beijing 100028, China
3
Defective Product Administrative Center, State Administration for Market Regulation, Beijing 100088, China
*
Author to whom correspondence should be addressed.
Energies 2023, 16(16), 5919; https://doi.org/10.3390/en16165919
Submission received: 6 June 2023 / Revised: 11 July 2023 / Accepted: 26 July 2023 / Published: 10 August 2023
(This article belongs to the Special Issue Prognostics of Battery Health and Faults)

Abstract

:
The nanoscale 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 Li-manganese-rich electrode material was synthesized by the co-precipitate method, and its electrochemical properties were systematically analyzed, especially the electrochemical impedance spectroscopy. The failure of the electrode interface and the structural transformation of the material at high potential are the main reasons for the deterioration of the Li-manganese-rich electrode, and high temperatures accelerate the deterioration. Based on the systematic analysis of the induced reactance change with electrode polarization potential, it is found that the induced reactance of a Li-manganese-rich electrode is not only related to the degree of delithiation/lithiation but also has a great relationship with the performance of the electrode/electrolyte interface. This conclusion is beneficial for the manufacturing of battery failure analysis by providing a theoretical basis for guidance.

1. Introduction

Energy and environmental problems have gradually become one of the most prominent problems facing mankind. Especially after 2004, due to rapid population growth and rapid socio-economic development, all countries in the world, especially the United States, China, and other major energy-consuming countries, are overly dependent on fossil fuels such as oil, natural gas, and other energy sources, resulting in the rapid consumption of oil, natural gas, and other disposable energy sources, which prompts people to explore and develop new clean energy and energy storage devices.
At present, the main chemical energy storage devices that people focus on are lithium-ion batteries, fuel cells, and super capacitors [1,2,3,4,5,6,7,8]. With the progress of science and technology, the requirements for energy storage devices in electronic products, electric vehicles, medical equipment, aerospace, and other fields have become increasingly higher, and among the many energy storage devices, new energy storage devices are gradually becoming a hot spot for research.
Li-ion batteries [9,10,11] have become the mainstream of digital products [12,13,14], automotive power batteries [15,16,17], and energy storage batteries because of their advantages of high energy density, small size, and long cycle life. Among them, the nickel-cobalt-manganese ternary cathode material has received wide attention and become a hot research topic, while the Li-manganese-rich electrode material also provides a new idea for the further development of Li-ion batteries because of its higher mass-specific capacity [18,19]. However, some adverse factors, such as large first irreversible capacity losses, poor magnification performance, and phase transitions of some materials during the cycle, inhibit the development of commercialization. To solve the above problems, we must first understand more information about the electrochemical behavior of electrode materials during lithium delithiation/lithiation, including the performance of SEI and the charge transfer process at the electrode/electrolyte interface, which play a key role in cycle stability and capacity maintenance. EIS is one of the most effective instruments for researching the interface properties of electrode materials based on the delithiation/lithiation mechanism and has been confirmed before [20,21,22,23,24]. However, there are few studies on the failure mechanism of the electrode interface of Li-manganese-rich electrode materials in systematic EIS testing. An in-depth understanding of the failure mechanism of the electrode interface can provide the correct direction and appropriate method for the improvement of the electrode materials performance, which is of great significance for the further improvement of the electrochemical performance of materials.
This paper focuses on the electrode/electrolyte interface performance of Li-manganese-rich electrode materials and extracts the relevant kinetic parameters of the delithiation/lithiation process of lithium ions in the intercalation compound particles, such as the resistance of lithium ion diffusion and migration through the SEI, the electrode inductance, etc. At the same time, the relationship between these kinetic parameters and electrode polarization potential, temperature, and degree of lithium insertion was explored to understand the reaction mechanism and failure mechanism of Li-manganese-rich electrode materials. Meanwhile, the dependence of the electrode-induced reaction on temperature and electrode polarization potential was tested and analyzed by shortening the slurry mixing process and controlling the low temperature according to the special impedance characteristics of the electrode-induced reaction during EIS testing and analysis.

2. Experiment

2.1. Coprecipitation Synthesis and the Electrode Interface Characteristics Studies of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2

Commercial Li-manganese-rich electrode materials are usually synthesized by the solid-phase method, which has the advantages of being simple, practical, low-cost, and suitable for large-scale production. But the disadvantage is that the material particles are generally large, and the control of the material uniformity is not precise enough. In addition to the solid-phase method in production, the co-precipitation method [25] is also a practical method commonly used in industry for large-scale production of electrode materials. Compared with the solid phase method, the co-precipitation method is more costly, but it has great advantages in the control of material particle size and material purity.
Therefore, in this experiment, 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 cathode material with high nickel content was synthesized by the co-precipitation method, and its electrochemical properties and reaction mechanism at the electrode/electrolyte interface were characterized and investigated.
Figure 1 shows the schematic diagram of the co-precipitation device for the synthesis of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 cathode material.
Firstly, MnSO4·H2O, NiSO4·6H2O, CoSO4·7H2O were weighed according to the molecular ratio of nickel, cobalt, and manganese in 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 cathode material and dissolved in deionized water to produce a total cation concentration of 1 molL−1. Then, a 1 molL−1 LiOH solution was prepared and poured into a large beaker, stirring continuously while heating. The temperature was controlled at 50 °C and the magnetic stirring speed was 200 r/min. Then the peristaltic pump was used to drop the nickel, cobalt, manganese cationic solution, and ammonia water into the LiOH solution of 1 molL−1, and the peristaltic pump was adjusted to control the pH of the reaction solution at about 11. When the cation solution was added, the drip of ammonia water was stopped, and then the water bath heating and magnetic stirring were set for 6 h so that the cationic precipitation in the reaction could be soluted completely.
The resulting suspension was then filtered and cleaned with deionized water; the procedure was repeated until the pH of the supernatant was close to 7. Then, after drying at 50 °C for 12 h, the suspension was thoroughly mixed with LiOH·H2O ball milling with a certain stoichiometric ratio. After pre-sintering at 400 °C for 4 h and then at 900 °C for 4 h, 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 cathode material was obtained after cooling.

2.2. Characterization

X-ray diffraction (XRD) was performed by step scanning over an angular range of 10–80° with a step width of 0.02°. To study the micromorphology of the samples, a scanning electron microscope (SEM) was used.

2.3. Electrochemical Tests

The positive electrode mixture was prepared by the traditional mechanical homogenization method, and the positive electrode was prepared by coating and drying in an extremely dry environment. The electrode was transferred to a glove box. The assembly of the CR2032 (Shenzhen LiYou new energy Technology Co., Ltd., Shenzhen, China) coin cell and three-electrode glass cell shown in Figure 2 was completed in the glove box to systematically analyze the electrochemical properties, especially the interface property of the electrode.

3. Results and Discussion

3.1. XRD and SEM Studies

The XRD pattern of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 electrode material is shown in Figure 3. It has been observed that the XRD characteristic peaks of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 electrode material synthesized by the co-precipitate method are in good agreement with those of commercial Li-manganese-rich electrode materials, and the relatively weak peaks of 21–23° correspond to the structural characteristics of Li2MnO3 in the layered solid solution [27], and the XRD characteristic peaks synthesized by co-precipitate method are sharper with fewer impurity peaks and better crystallization.
From the SEM image of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 electrode material powder sample shown in Figure 4, it can be observed that the lithium-manganese-based cathode material prepared by the coprecipitate method is composed of irregular particles with a size of about 100–200 nm and short rod-like particles with a length of 300–500 nm.

3.2. Electrochemical Study

The CV tests shown in Figure 5 are the results of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 electrode materials at the voltage range of 2.0–5.0 V, respectively, and the scanning speed is 0.1 mVs−1. It can be observed from the figure that in the first cycle of CV forward scanning of commercial lithium manganese-based electrode materials, an oxidation current peak appears around 4.0 V and 4.6 V, respectively, corresponding to the redox between Ni2+/4+ and Co3+/4+ and the controversial deep de-lithium process. There was no obvious oxidation peak after 4.8 V. The oxidation peak of about 4.5 V almost disappeared in the subsequent cycle process, but in the voltage range of 3.0–3.5 V in the reduction process, a new steamed bun peak appeared, and as the CV cycle went on, the peak gradually increased, which was opposite to the change law of the current peak of about 4.5 V.
Figure 6 shows the charge and discharge curve of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 electrode in the first two cycles. Due to the limitation of the high voltage range of the electrolyte, the test voltage range of the battery is 2.5–4.8 V. The cyclic capacity tests were carried out at room temperature and 55 °C under a current density of 50 mAg−1. It was shown that the first cycle charging capacity of the battery at 55 °C is slightly higher than that at room temperature, but the discharge capacity is relatively lower, indicating that the battery at high temperature has a larger irreversible capacity loss of 28.5% in the first cycle than that at room temperature of 22.7%. It can also be observed from the cycle performance curve in Figure 7 that the discharge capacity of the battery tested at room temperature attenuates from 279.4 mAhg−1 to 250 mAhg−1 after 50 charge-discharge cycles and the capacity retention rate reaches 89.5%. The discharge capacity of the battery tested at 55 °C decays from 264 mAhg−1 to 198.3 mAhg−1 after 50 cycles, and the capacity retention rate reaches 77.4%.
Figure 8 shows the magnification performance test results of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 electrode under different charging and discharging current densities (50/100/200/500 mAg−1) at 2.0–4.8 V. It can be observed that the reversible capacity of the battery is close to 270 mAhg−1 with the current density of 50 mAg−1, and it is still 170 mAhg−1 with the current density of 500 mAg−1, with the capacity retention rate reaching 63%.
Although such performance and magnification performance still cannot well meet the increasing requirements of high temperature working environments and fast charging of power batteries, Further improvements can be made to the overall performance of lithium manganese-based electrode materials by controlling the material composition and morphology, coupled with appropriate material modification and appropriate electrolyte. With the development of power battery technology, Li-manganese-rich electrode materials will be more widely used.

3.3. Electrochemical Study

Figure 9 shows the Nyquist diagram of 2.5–4.8 V during the first charge and discharge of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 electrode at room temperature. It can be observed that the Nyquist diagram with an open circuit potential of 3.3 V is presented as a small arc in the high frequency range, an arc in the middle-high frequency range, and a large arc in the middle-low frequency range. As for the two small arcs in the high frequency range, they are respectively attributed to the contact resistance and the transmission resistance of lithium ions in the SEI. As the electrode polarization potential increased from 3.3 V to 3.9 V, the arc of the middle-low frequency range gradually curved and evolved into a semicircle in the middle frequency range and a slash in the low frequency range at 3.9 V. At this point, the Nyquist diagram consists of four parts: two arcs, one semicircle, and a slash. According to previous results, the semicircle in the middle frequency range is attributed to the charge transfer process, while the slant in the low frequency range is attributed to the solid diffusion process of lithium ions within the active material particles of the Li-manganese-rich electrode. As the electrode polarization potential continues to increase, the semicircle in the middle frequency range transforms back into an arc and shifts away from the real axis. Finally, the semicircle in the middle frequency range and the slash in the low frequency range turn into a large arc in the low frequency range at 4.7 V. In the process of the first discharge, the arc in the low frequency range changes towards or away from the real axis as the electrode potential changes, but it does not evolve into a semicircle or slash through the discharge process.
In order to further analyze the mechanism of more rapid decaying 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 electrode at high temperature, shown in Figure 7. The Nyquist diagram of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 electrode at 3.3–4.8 V during the first charge at 55 °C is shown in Figure 10. It can be observed that consistent with the test results at room temperature and 3.3 V, the EIS of the Li-manganese-rich electrode consists of a small semicircle in the high frequency range and a large arc in the middle-low frequency range. With the increase of the polarization potential, the arc in the middle-low frequency range gradually curves towards the X axis and completely transforms into a semicircle in the middle frequency range and a slant in the low frequency range at 3.9 V. Then, with the increase in polarization potential, it gradually transforms into a large arc again in the middle-low frequency range.
As shown in Figure 11, an equivalent circuit is proposed to fit the Nyquist plots of Figure 9. The RC circuit signifies the semicircle or arc in the Nyquist plots. Rs is Ohmic resistance; Rcf is the resistance of Schottky contact; RSEI is the resistance of SEI; and Rct is the resistance of charge transfer. The capacitance of the Schottky contact, SEI resistance, and double layer are represented by Qcf, QSEI, and Qdl, respectively.
Figure 12 shows the relationship between the arc resistance value in the high frequency range and the polarization potential of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 electrode fitted according to the equivalent circuit in Figure 11 under different experimental conditions.
Figure 12a,b show the transformation laws of Rcf and RSEI as a function of electrode polarization potential during the first charge and discharge of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 electrode at 2.5–4.8 V at room temperature. It can be observed that there is a strong dependence between the changes of Rcf and RSEI, but the change of contact impedance is smaller and increases only during the second step of de-lithium.
Figure 12b shows the transformation law of RSEI during the first charge and discharge of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 electrode at room temperature between 2.5 V and 4.8 V. It can be observed that RSEI slowly decreased as the polarization potential increased from 3.3 V to 3.85 V during the first charging process, which was mainly due to the reaction of a small amount of initial SEI with the electrolyte. When the electrode potential increased from 3.85 V to 4.5 V, RSEI increased slowly. Compared with CV test results, it can be found that the first step of the de-lithium process begins and the redox processes of Ni2+/4+ and Co3+/4+ occur, which are mainly attributed to the removal of lithium ions and the oxygenation of electrolyte near the electrode surface caused by Ni4+ with strong catalytic activity to generate a new SEI film [28]. Then, at the potential range above 4.5 V, the impedance of SEI film increased rapidly again with the second step of the de-lithium process, indicating that this step of the de-lithium process leads to the instability of the Li-manganese-rich electrode/electrolyte interface. The oxygen removal process and the transition of Mn4+ to a lower valence state will inevitably lead to changes in the material structure, and the components of SEI film will also be affected. In addition, there are some problems, such as component decomposition in the potential interval above 4.5 V of the EC base electrolyte, resulting in increased instability of the interface and rapid thickening of the SEI film. During the first discharge, RSEI gradually increased with the decrease of the polarization potential from 4.8 V to 2.5 V, which was caused by the continuous thickening of the SEI film caused by the insertion of lithium ions, the irreversibility of the redox process, and the spontaneous decomposition of the electrolyte.
Figure 12c shows the transformation law of RSEI as a function of polarization potential during the first charge of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 electrode between 3.3 V and 4.8 V at 55 °C. It can be seen from the figure that 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 electrode RSEI changed basically the same as at room temperature. Before the second step of the de-lithium process, SEI film is relatively stable. Later, due to the influence of high temperature and the instability of the electrode/electrolyte interface at high potential, RSEI increased rapidly.
Figure 13 shows the relationship between the semi-circular resistance values in the middle frequency range of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 electrode and electrode polarization potential at room temperature and 55 °C, fitted according to the equivalent circuit in Figure 10. According to previous results, the semicircle in the intermediate frequency range is attributed to the charge transfer process impedance of the electrode/electrolyte interface.
As shown in Figure 13a, the variation curve of charge transfer impedance Rct as a function of electrode polarization potential during the first charge of a Li-rich manganese-based electrode at room temperature is 3.8–4.6 V. It can be observed that in the potential range below 4.0 V, the RctS gradually decreased with the increase in polarization potential, which is in good combination with the change law of intercalation electrode RctS and electrode polarization potential, representing the redox processes of Ni2+/4+ and Co3+/4+. Then, with the increasing polarization potential, the Rct basically remained unchanged when the voltage range was lower than 4.5 V, indicating that the lithium removal reaction inside the electrode and the lithium ion transfer process at the electrode/electrolyte interface were relatively stable within this potential range, which was also consistent with the change rule of RSEI. When the potential range exceeds 4.5 V, Rct increases rapidly during the second stage of the de-lithiation process. This is because the oxygen removal process and the transformation of Mn4+ to the low valence state will inevitably lead to the transformation of material structure, which seriously affects the lithium ions transport performance at the electrode/electrolyte interface. Moreover, the changes in Rct and RSEI are strongly dependent on Li-manganese-rich electrodes. In other words, the stability of the electrode/electrolyte interface directly determines the electrochemical activity of the entire electrode, which makes improving the interface performance of materials an important means for the modification of Li-manganese-rich electrodes.

3.4. The Variations of Inductive Reactance for Li-Manganese-Rich Cathode

Figure 14 shows the change rule of inductive reactance with temperature during charge and discharge of Li-manganese-rich electrode. Due to the phenomenon of stripping of Li-manganese-rich electrode in the discharge process, the semicircle in the high frequency range will rapidly increase, resulting in the reactance in the middle frequency range being covered. Therefore, only the EIS data charged to 4.0 V is provided here. It can be observed that, in the EIS of 4.0 V charging, the diameter of the circles associated with inductive reactance in the middle frequency range decreases from large to small as the working temperature increases from −20 °C to 30 °C, and the circles related to inductance disappear at about 10 °C. This phenomenon indicates that lower temperatures weaken the contact between the active material particles of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 electrode and between the active material and the conductive agent, which leads to increased inhomogeneity of lithium ions during embedding and removal processes. The electric field formed by the concentration cell zone resists the delithiation and lithiation processes.
However, the large arc associated with the charge transfer process in the middle-low frequency range increases gradually as the temperature decreases and finally becomes a diagonal line at low temperatures, indicating the low electrochemical activity of the Li-manganese-rich electrode at lower temperatures.
Combined with the previous findings on the change of the induced reactance in the electrochemical impedance spectrum of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 electrode with temperature, in order to better study the change of the inductive reactance in the electrochemical impedance spectrum of Li-manganese-rich electrode with electrode polarization potential during charging, the EIS of Li-Manganese-rich electrode was performed at −20 °C. According to the research results of the variable temperature-induced reactance shown in Figure 14, the inductive reactance can be found more easily at −20 °C for long-term research.
Figure 15 shows the variation of inductive reactance with electrode polarization potential during charging and discharging at −20 °C for 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 electrode. From Figure 15a, it can be seen that at an open-circuit potential of 3.3 V, there is an initial large circle associated with inductive reactance in the middle-high frequency range of the EIS of Li-manganese-rich electrode because there is no delithiation/lithiation process of lithium ions at this time. The induced reactance at this time should be completely attributed to the local concentration cell caused by the inhomogeneity of the conductivity between the active material particles of the electrode and the inhomogeneity of the initial SEI film. With the increase of polarization potential to 3.8 V, it can be seen that the diameter of the ring increases gradually and then decreases gradually with the increase of polarization potential, and the ring moves to the middle frequency range. It can be observed from the CV results that the process of delithium begins after 3.8 V, and a stable SEI film gradually forms with the process of delithium. The inductive reactance caused by the inhomogeneity of the initial SEI film is weakened, and the induced reactance caused by the inhomogeneity of the lithium ion is gradually dominated by the removal of lithium ions, resulting in the movement of the ring to the middle frequency range. Then the rings in the middle frequency range remained until the end of the charging process, while the rings associated with inductive reactance disappeared during the discharge process and did not appear again. As the electrode polarization potential increases, the arc in the middle-low frequency range bends toward the x-axis, eventually forming a semicircle attributed to charge transfer in the middle frequency range and a slash related to solid diffusion in the low frequency range. The radius of the semicircle in the middle frequency range gradually increases until it eventually reverts back to an arc in that range.
According to the experimental results, Figure 16 shows the equivalent circuit fitted by the EIS at −20 °C for a lithium-manganese-based electrode. Where Q stands for constant phase angle element (CPE), Rs is the solution resistance of the electrolyte, RSEI, and Qdl, respectively, represent the semi-circular resistance of lithium-ion transmission in SEI film and SEI film capacitance in the high frequency range, L and RL, respectively, represent inductive reactance and corresponding resistance, and Rct and Qdl represent the middle frequency range and charge transfer resistance and double layer capacitance. QD designates the diffusion impedance.
According to the equivalent circuit given in Figure 16, the EIS data in Figure 15 were fitted, and the obtained inductive reactance of the Li-Mangan-based electrode changing with the polarization potential of the electrode was given in Figure 17. As can be observed from the figure, in the voltage range of 3.3–3.8 V during the first charge, the inductance value of the Li-manganese-rich electrode increases slowly with the increase in electrode polarization potential. There is no lithium-ion removal process in this voltage range, which is only attributed to the local concentration battery caused by the uneven conduction of the electrode and the uneven initial SEI film. The L value decreased rapidly after 3.8 V when the Li-manganese-rich electrode material just started to delithium. The transfer of lithium ions in the SEI film promoted the homogenization of the SEI film and weakened the previous inductance. In the voltage range of 4.1–4.5 V, the first step of the delithium process ends, the second step of the delithium reaction does not begin, and the inductive reactance tends to be stable. However, after 4.5 V, the simple intercalation process basically ends, while the process related to phase transition may cause microscopic changes in the electrode, resulting in the gradual reduction of inductance and finally disappearing in the discharge process.
As for the specific causes of inductive reactance, there are generally two explanations: one is that the formation of uneven SEI film on the surface of electrode materials leads to the uneven transmission of lithium ions, resulting in a local concentration difference of lithium ions on the surface of the electrode, forming an induced electromotive force, which can well explain the mechanism of inductive reactance in the EIS of a Si negative electrode [29]. Second, due to the inhomogeneity of lithium ions in the delithiation/lithiation process, the amount of lithium ion delithiation/lithiation is different among particles of different active materials, resulting in Li-rich and Li-poor ranges and producing concentrated batteries. This explanation has also been well verified in positive electrode materials LiCo2 and LiMn2 O4 [30,31]. However, the electric field generated by the local current between the two electrodes in the battery with the above local concentration differences will oppose the electric field in the process of lithium ion delithiation/lithiation, thus preventing the delithiation/lithiation of lithium ions in the electrode material and negatively affecting the electrochemical activity and performance of the battery’s electrode material. In addition, multiple indicators suggest that the generation mechanisms of inductive reactance in different electrode materials differ. Studying the generation mechanisms of inductive reactance across different electrode materials and analyzing their dependence on temperature and electrode polarization potential can offer theoretical guidance for analyzing battery failures during production.

4. Conclusions

The nanoscale 0.5 Li2 MnO3·0.5 Li(Ni0.44 Mn0.44 Co0.12)O2 Li-manganese-rich electrode material was synthesized by the co-precipitate method, and the electrochemical tests were carried out, particularly using the EIS. The mechanism of Li-rich manganese-based electrode material and its failure at high temperatures were discussed. Results revealed that the failure of the electrode interface and the structural transformation of the material at high potential are the main reasons for the deterioration of the performance of the Li-manganese-rich electrode, and the increasing instability of the Li-manganese-rich electrode/electrolyte interface at high temperature also accelerates the deterioration of the electrochemical performance of the electrode. Based on the systematic analysis of the inductance problem found in the EIS test of Li-manganese-rich electrode, it is found that the inductive reactance is more likely to occur at low temperatures because low temperatures seriously affect the conductivity of materials and the migration rate of lithium ions. The dependence of induced reactance on electrode polarization potential showed that induced reactance in the Li-rich manganese-based electrode is not only related to the degree of delithiation/lithiation but is also closely linked to the electrode/electrolyte interface’s performance.

Author Contributions

Conceptualization, H.D. and X.Z.; methodology, X.Z.; validation, P.W. and Y.W.; formal analysis, P.C.; investigation, X.Z.; resources, P.C.; data curation, P.W.; writing—original draft preparation, X.Z.; writing—review and editing, H.D.; visualization, X.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Presidential Fund of CNIS (282022Y−9463) and the Science, Technology Program of the State Administration for Market Regulation (2022MK183) and the National Key Research and Development Program of China (2021YFF0601100).

Data Availability Statement

The research data is unavailable due to privacy.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Khan, M.A.; Zhao, H.B.; Zou, W.W.; Chen, Z.; Cao, W.; Fang, J.; Xu, J.; Zhang, L.; Zhang, J. Recent progresses in electrocatalysts for water electrolysis. Electrochem. Energy Rev. 2018, 1, 483–530. [Google Scholar] [CrossRef] [Green Version]
  2. Lu, J.; Chen, Z.W.; Pan, F.; Cui, Y.; Amine, K. High-performance anode materials for rechargeable Li-thiumion batteries. Electeochem. Energy Rev. 2018, 1, 35–53. [Google Scholar] [CrossRef]
  3. Tian, N.; Lu, B.; Yang, X.D.; Huang, R.; Jiang, Y.X.; Zhou, Z.Y.; Sun, S.G. Rational design and synthesis of low-temperature fuel cell electrocatalysts. Electeochem. Energy Rev. 2018, 1, 54–83. [Google Scholar] [CrossRef] [Green Version]
  4. Wang, Y.J.; Fang, B.; Zhang, D.; Li, A.; Wilkinson, D.P.; Ignaszak, A.; Zhang, L.; Zhang, J. A review of carbon-composited materials as air-electrode bifunctional electrocatalysts for metal-air batteries. Electeochem. Energy Rev. 2018, 1, 1–34. [Google Scholar] [CrossRef]
  5. Zang, H.M.; Lu, W.J.; Li, X.F. Progress and perspectives of flow battery technologies. Electeochem. Energy Rev. 2019, 2, 492–506. [Google Scholar] [CrossRef]
  6. Steele, B.C.H.; Heinzel, A. Materials for fuel-cell technologies. Nature 2001, 414, 345–352. [Google Scholar] [CrossRef]
  7. Steele, B.C.H. Fuel-cell technology-Running on natural gas. Nature 1999, 400, 619–621. [Google Scholar] [CrossRef]
  8. Demirdoven, N.; Deutch, J. Hybrid cars now, fuel cars later. Science 2004, 305, 974–976. [Google Scholar] [CrossRef] [Green Version]
  9. Scrosati, B.; Garche, J. Lithium batteries: Status, prospects and future. J. Power Sources 2010, 195, 2419–2430. [Google Scholar] [CrossRef]
  10. Winter, M. What are batteries, fuel cells, and supercapacitors. Chem. Rev. 2004, 104, 4245–4269. [Google Scholar] [CrossRef] [Green Version]
  11. Kang, B.; Ceder, G. Battery materials for ultrafast charging and discharging. Nature 2009, 458, 190–193. [Google Scholar] [CrossRef]
  12. Whittingham, M.S. Lithium batteries and cathode materials. Chem. Rev. 2004, 104, 4271–4301. [Google Scholar] [CrossRef]
  13. Xin, S.; Guo, Y.G.; Wan, L.J. Electrode materials for lithium secondary batteries with high energy densities. Sci. China Chim. 2011, 41, 1229–1239. [Google Scholar] [CrossRef]
  14. Nazri, G.A.; Pistoia, G. Lithium Ion Batteries-Science and Technology; Springer: New York, NY, USA, 2009; pp. 1–8. [Google Scholar]
  15. Armand, M.; Tarascon, J.M. Building better batteries. Nature 2008, 451, 652–657. [Google Scholar] [CrossRef]
  16. Tarascon, J.M.; Armand, M. Issues and challenges facing rechargeable lithium batteries. Nature 2001, 414, 359–367. [Google Scholar] [CrossRef]
  17. Aifantis, K.E.; Hackney, S.A.; Kumar, V.R. High Energy Density Lithium Batteries-Materials, Engineering, Applications; Wiley-VCH: Weinheim, Germany, 2010. [Google Scholar]
  18. Boulineau, A.; Croguennec, L.; Delmas, C.; Weill, F. Structure of Li2MnO3 with different degrees of defects. Solid State Ion 2010, 180, 1652–1659. [Google Scholar] [CrossRef]
  19. Zhang, P.P.; Huang, H.; He, Y.P.; Li, X.B.; Guo, Z.C. Recent development of Li-rich manganese cathode material for Li-ion batteries. J. Mater. Eng. 2021, 49, 48–58. [Google Scholar]
  20. Huang, J. Electrochemical impedance spectroscopy for electrocatalytic interfaces and reactions: Classics never die. J. Electrochem. 2020, 26, 3–18. [Google Scholar]
  21. Ezpeleta, I.; Freire, L.; Mateo-Mateo, C. Characterisation of Commercial Li-Ion Batteries Using Electrochemical Impedance Spectroscopy. Chem. Sel. 2022, 7, e202104464–e202104469. [Google Scholar] [CrossRef]
  22. Kasper, M.; Leike, A.; Thielmann, J.; Winkler, C.; R-Smith, N.A.Z.; Kienberger, F. Electrochemical impedance spectroscopy error analysis and round robin on dummy cells and lithium-ion-batteries. J. Power Sources 2022, 536, 231407. [Google Scholar] [CrossRef]
  23. Galatro, D.; da Silva, C.; Romero, D.A.; Gong, Z.; Trescases, O.; Amon, C.H. Battery Health Diagnosis Approach Integrating Physics-Based Modeling with Electrochemical Impedance Spectroscopy. Energy Technol. Gener. Convers. Storage Distrib. 2022, 10, 2100942. [Google Scholar] [CrossRef]
  24. Huang, Q.A.; Li, W.H.; Tang, Z.P.; Zhang, F.Z.; Li, A.J.; Zhang, J.J. Fundamentals of Electrochemical impedance spectroscopy. Chin. J. Nat. 2020, 42, 1–15. [Google Scholar]
  25. Li, J.; Klopsch, R.; Stan, M.C.; Nowak, S.; Kunze, M.; Winter, M.; Passerini, S. Synthesis and electrochemical performance of the high voltage cathode material Li[Li0.2Mn0.56Ni0.16Co0.08]O2 with improved rate capability. J. Power Sources 2011, 196, 4821. [Google Scholar] [CrossRef]
  26. Zhao, X.; Zhuang, Q.C.; Wu, C.; Wu, K.; Xu, J.M.; Zhang, M.Y.; Sun, X.L. Impedance studies on the capacity fading mechanism of Li(Ni0.5Co0.2Mn0.3)O2 cathode with high-voltage and high-temperature. J. Electrochem. Soc. 2015, 162, A2770–A2779. [Google Scholar] [CrossRef]
  27. Ren, C.; Kong, J.Z.; Zhou, F.; Yang, X.; Tang, Z.; Li, J.X. Research Progress on the Li-excess Mn-based Cathode Materialsfor Lithium-ion Battery. Mater. Rev. 2013, 27, 29–49. [Google Scholar]
  28. Pistoia, G.; Antonini, A.; Rosati, R. Storage characteristics of cathodes for Li-ion batteries. Electrochim. Acta 1996, 41, 2683–2689. [Google Scholar] [CrossRef]
  29. Radvanyi, E.; Havenbergh, K.V.; Porcher, W.; Jouanneau, S.; Bridel, J.S.; Put, S.; Franger, S. Study and modeling of the Solid Electrolyte Interphase behavior on nano-silicon anodes by Electrochemical Impedance Spectroscopy. Electrochim. Acta 2014, 137, 751–757. [Google Scholar] [CrossRef]
  30. Zhuang, Q.C.; Wei, T.; Wei, G.Z.; Dong, Q.F.; Sun, S. GElectrochemical Impedance Spectroscopic Studies of Insertion and Deinsertion of Lithium Ion in Spinel LiMn2O4. Acta Chim. Sin. 2009, 67, 2184–2192. [Google Scholar]
  31. Zhuang, Q.C.; Xu, J.M.; Fan, X.Y.; Wei, G.Z. Electrochemical impedance spectroscopy of LiCoO2 electrode/electrolyte interface. Sci. China 2007, 37, 18–24. [Google Scholar]
Figure 1. The scheme for coprecipitation synthesis of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 Cathode Materials Characterization.
Figure 1. The scheme for coprecipitation synthesis of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 Cathode Materials Characterization.
Energies 16 05919 g001
Figure 2. Three-electrode glass cell [26].
Figure 2. Three-electrode glass cell [26].
Energies 16 05919 g002
Figure 3. Comparison of XRD pattern of 0.5Li2MnO3 0.5Li(Ni0.44Mn0.44Co0.12)O2 Cathode Materials.
Figure 3. Comparison of XRD pattern of 0.5Li2MnO3 0.5Li(Ni0.44Mn0.44Co0.12)O2 Cathode Materials.
Energies 16 05919 g003
Figure 4. SEM image of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 Cathode Materials. (a) micrometer scale; (b) nanoscale.
Figure 4. SEM image of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 Cathode Materials. (a) micrometer scale; (b) nanoscale.
Energies 16 05919 g004
Figure 5. The initial three CVs of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 cathode in the voltage ranges of 2.0–5.0 V.
Figure 5. The initial three CVs of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 cathode in the voltage ranges of 2.0–5.0 V.
Energies 16 05919 g005
Figure 6. Charge-discharge characteristics of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 Cathode. (a) Room temperature; (b) 55 °C.
Figure 6. Charge-discharge characteristics of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 Cathode. (a) Room temperature; (b) 55 °C.
Energies 16 05919 g006
Figure 7. Cycle performance of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 cathode under (a) room temperature and (b) 55 °C.
Figure 7. Cycle performance of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 cathode under (a) room temperature and (b) 55 °C.
Energies 16 05919 g007
Figure 8. The rate performances of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 cathode in the voltage range of 2.0–4.8 V.
Figure 8. The rate performances of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 cathode in the voltage range of 2.0–4.8 V.
Energies 16 05919 g008
Figure 9. Nyquist plots of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 electrode during the first charge (ae) and discharge (fh) processes from 2.5 to 4.8 V at room temperature.
Figure 9. Nyquist plots of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 electrode during the first charge (ae) and discharge (fh) processes from 2.5 to 4.8 V at room temperature.
Energies 16 05919 g009
Figure 10. Nyquist plots of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 electrode during the first charge process from 3.3 to 4.8 V under 55 °C. (a) 3.3−3.7 V; (b) 3.8−4.2 V; (c) 4.3−4.5 V; (d) 4.6−4.8 V.
Figure 10. Nyquist plots of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 electrode during the first charge process from 3.3 to 4.8 V under 55 °C. (a) 3.3−3.7 V; (b) 3.8−4.2 V; (c) 4.3−4.5 V; (d) 4.6−4.8 V.
Energies 16 05919 g010
Figure 11. Equivalent circuit proposed for fitting Nyquist plots of Figure 9.
Figure 11. Equivalent circuit proposed for fitting Nyquist plots of Figure 9.
Energies 16 05919 g011
Figure 12. Variations of (a) Rcf under room temperature, (b) RSEI under room temperature, and (c) RSEI under 55 °C.
Figure 12. Variations of (a) Rcf under room temperature, (b) RSEI under room temperature, and (c) RSEI under 55 °C.
Energies 16 05919 g012
Figure 13. Variations of Rct (a) under room temperature and (b) under 55 °C.
Figure 13. Variations of Rct (a) under room temperature and (b) under 55 °C.
Energies 16 05919 g013
Figure 14. Variations of inductive reactance for Li-manganese-rich Cathode with temperature at 4.8 V during charge.
Figure 14. Variations of inductive reactance for Li-manganese-rich Cathode with temperature at 4.8 V during charge.
Energies 16 05919 g014
Figure 15. Variations of inductive reactance for Li-manganese-rich Cathode with electrode potential at −20 °C. (a) 3.3−3.8 V; (b) 3.85−4.05 V; (c) 4.1−4.5 V; (d) 4.6−4.8 V; (e) 4.7−4.3 V; (f) 4.2−3.8 V.
Figure 15. Variations of inductive reactance for Li-manganese-rich Cathode with electrode potential at −20 °C. (a) 3.3−3.8 V; (b) 3.85−4.05 V; (c) 4.1−4.5 V; (d) 4.6−4.8 V; (e) 4.7−4.3 V; (f) 4.2−3.8 V.
Energies 16 05919 g015
Figure 16. Equivalent circuit proposed for fitting impedance spectra of Li-manganese-rich Cathode under −20 °C.
Figure 16. Equivalent circuit proposed for fitting impedance spectra of Li-manganese-rich Cathode under −20 °C.
Energies 16 05919 g016
Figure 17. Variations of L with electrode potential obtained from fitting the experimental impedance spectra of Li-manganese-rich Cathode during the first delithiation/lithiation process.
Figure 17. Variations of L with electrode potential obtained from fitting the experimental impedance spectra of Li-manganese-rich Cathode during the first delithiation/lithiation process.
Energies 16 05919 g017
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

Zhao, X.; Wang, P.; Wang, Y.; Chao, P.; Dong, H. Coprecipitation Synthesis and Impedance Studies on Electrode Interface Characteristics of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 Cathode Material. Energies 2023, 16, 5919. https://doi.org/10.3390/en16165919

AMA Style

Zhao X, Wang P, Wang Y, Chao P, Dong H. Coprecipitation Synthesis and Impedance Studies on Electrode Interface Characteristics of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 Cathode Material. Energies. 2023; 16(16):5919. https://doi.org/10.3390/en16165919

Chicago/Turabian Style

Zhao, Xing, Peng Wang, Yan Wang, Peipei Chao, and Honglei Dong. 2023. "Coprecipitation Synthesis and Impedance Studies on Electrode Interface Characteristics of 0.5Li2MnO3·0.5Li(Ni0.44Mn0.44Co0.12)O2 Cathode Material" Energies 16, no. 16: 5919. https://doi.org/10.3390/en16165919

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