Jia Jin*ab,
Jinping Weibc,
Zhen Zhoubc and
Zhaojun Xiebc
aTianjin Guoan Mengguli New Materials Science and Technology Co., Ltd, Tianjin 301800, China. E-mail: 277282047@qq.com
bTianjin Enterprise Key Laboratory of Key Materials and Technology for Solid State Batteries, Tianjin 300182, China
cInstitute of New Energy Material Chemistry, Nankai University, Tianjin 300350, China
First published on 20th April 2023
The 5V spinel LiNi0.5Mn1.5O4 cathode materials with different morphology were prepared by a solid state calcination method and characterized by X-ray diffraction (XRD), inductively coupled plasma (ICP), field emission scanning electron microscope (FE-SEM). Electrochemical properties of cathode material were investigated by electrochemical impedance spectroscopy (EIS), galvanostatic intermittent titration technique (GITT) and electrochemical performance tests. Compared with polycrystalline morphology (PLNMO), LiNi0.5Mn1.5O4 material with single crystalline morphology (SLNMO) proved smaller electrochemical polarization or voltage difference, lower internal resistance, faster lithium-ion diffusivity, arising from higher Mn3+ content. Differential scanning calorimetry (DSC) showed that SLNMO was more stable than PLNMO at full charged state with organic electrolyte, which exhibited initial discharge capacity of 140.2 mA h g−1 at 0.1C, coulombic efficiency of 96.1%, and specific capacity retention of 89.2% after 200 cycles at 2.5C, a little inferior to that of 91.7% for PLNMO.
In this work, we synthesized two different morphologies of LiNi0.5Mn1.5O4 material including single crystalline and polycrystalline morphology through industrialized solid-state reaction method, analyzed Mn3+ content of the samples, assessed the half coin-cell electrochemical performances, and discussed why the voltage difference between charge and discharge exists and why it becomes larger as cycle increases, thereafter we found the single crystalline sample assembled half coin-cell exhibited smaller electrochemical polarization or voltage difference, lower internal resistance, faster lithium-ion diffusivity, and excellent thermal stability at full charged state with organic electrolyte when compared with PLNMO sample.
Ni0.25Mn0.75(OH)2 precursor powder was synthesized via an inert atmosphere coprecipitation reaction from a stoichiometric amount of NiSO4·6H2O, MnSO4·H2O as metallic salt aqueous solution, a specified amount of NaOH as base aqueous solution and a desired amount of NH4·H2O as chelating agent aqueous solution. Then precipitate was filtered, washed with distilled water and dried in a vacuum oven. Precursor with different morphologies could be obtained through carefully controlling parameters such as chelating mole ratio of metal/NH4, the concentration and flow speed of the above solution pumping into the tank reactor, pH, temperature, stirring speed, and so on. (SLNMO precursor parameter: mole ratio of metal/NH4 was 2, concentration of NiSO4·6H2O, MnSO4·H2O, NaOH solu. and NH4·H2O were 0.6 mol L−1, 1.8 mol L−1, 10 mol L−1, 13 mol L−1, corresponding flow speeds were 10 ml min−1, 10 ml min−1, 8 ml min−1, 1 ml min−1, pH = 11.6–11.8, reaction temperature was 50 °C, stirring speed was 1000 rpm. PLNMO precursor parameter: mole ratio of metal/NH4 was 0.5, concentration of NiSO4·6H2O, MnSO4·H2O, NaOH solu. and NH4·H2O were 0.6 mol L−1, 1.8 mol L−1, 10 mol L−1, 13 mol L−1, corresponding flow speeds were 10 ml min−1, 10 ml min−1, 5 ml min−1, 2 ml min−1, pH = 10.8–11.1, reaction temperature was 50 °C, stirring speed was 600 rpm.)
Sample | Li, % | Ni, % | Mn, % | BET, m2 g−1 | D50, μm | Mn3+*, % | a/b/c, Å | Rwp, % |
---|---|---|---|---|---|---|---|---|
a Mn3+* (ref. 7–9) denote: the ratio of the initial discharge capacity between 4.4 V and 3.0 V to the theoretical capacity (147 mA h g−1). | ||||||||
SLNMO | 3.86 | 15.82 | 45.15 | 0.655 | 5.565 | 11.2 | 8.177 | 3.9 |
PLNMO | 3.72 | 16.18 | 45.00 | 0.636 | 6.848 | 4.1 | 8.165 | 3.4 |
The main element contents of Li/Ni/Mn were in good agreement with designed formula from ICP analysis. BET of single crystal material was similar with poly crystal material in Table 1.
Fig. 2 shows FESEM images of single crystal and poly crystal of LiNi0.5Mn1.5O4. Particle size was 5.565 μm for single crystal and 6.848 μm for poly crystal.
The initial galvanostatic charge–discharge curves for Li/LiNi0.5Mn1.5O4 coin cells with different morphologies were shown in Fig. 3a. Both samples show two charge and discharge plateau. The high plateau located at 4.7 V, represent the redox reaction of Ni ion, and another one located at 4.0 V, represent the redox reaction of Mn ion. The potential difference between charge and discharge plateaus for the both samples was similar, which reflects similar process of lithium deposition/dissolution in the first cycle smoothly and reversibly. The initial discharge capacities for SLNMO and PLNMO were 140.2 mA h g−1 and 138.9 mA h g−1 at the current density of 14 mA g−1 (0.1C) respectively. At the current density of 350 mA g−1 (2.5C), the discharge capacity of SLNMO was 125.2 mA h g−1 with capacity retention 89.2%, a little lower than 128.8 mA h g−1 that of PLNMO with capacity retention 91.7% after 200 cycles in Fig. 3b. PLNMO had better cycling performance than SLNMO sample before 140 cycles, however after 140 cycles, the degradation rate of PLNMO was significantly higher than that of SLNMO. Due to limited testing resources, only 200 cycles were tested in coin cells. The average coulombic efficiency of SLNMO cycled from 1st to 200th was 99.34%, similar to that of 99.35% for PLNMO.
Fig. 3 Initial charge/discharge profiles of Li/electrolytes/LiNi0.5Mn1.5O4 cells at 0.1C rate (a) and cycle performance at 2.5C rate at 25 °C (b). |
Other factors affecting electrochemical performances such as particle size and residual alkali were investigated. Different particle size of SLNMO with D50 = 6.804 μm was prepared to verify the effect of particle size on the electrochemical performance of samples with the same morphology. Experiments have shown that SLNMO samples of different particle sizes between D50 = 6.804 μm and D50 = 5.565 μm contributed a little difference to electrochemical performances. When the content of residual lithium salts is high, the influence on the battery performance is huge, especially in the high nickel cathode material (residual lithium salts 0.5–1.2%) and lithium-rich manganese-based cathode material (residual lithium salts 0.5–2.0%). High nickel materials generally need to wash to reduce the residual alkali on the surface of the materials. When the content of residual lithium salts is very low, usually lower than 0.1%, then its effect is negligible, and 5V spinel cathode material belongs to such category, with residual lithium salts 0.01–0.03%.
The cycle voltage differences at 25 °C between charge plateau and discharge plateau for both samples were different. The increase rate of the voltage difference for SLNMO was lower than the rate for PLNMO all over the 200 cycle in Fig. 4.
Why the voltage difference between charge and discharge exists and why it becomes larger and larger as the number of cycles increase? The main reason lies in the overpotential (denote η), which is one of the core issues for battery dynamics. When the overpotential occurs, the chemical electrode potential or battery voltage (denote V) will deviate from the thermodynamic equilibrium potential (denote E), so the process can be expressed by eqn (1). In the charge process, the chemical electrode potential (denote Vc) can be expressed by eqn (2), and similarly in the discharge process, the chemical electrode potential (denote Vd) can be described as eqn (3).
V(x) = E(x) ± η(x)± | (1) |
Vc = εc + ΣIRc(i) | (2) |
Vd = εd − ΣIRd(i) | (3) |
The internal resistance of both samples could be described in detail through Fig. 5, the Nyquist plots of the EIS measurements performed at open circuit voltage before and after room temperature cycle performance at 2.5C with the frequency range of 10 mHz to 100 kHz for Li/LiNi0.5Mn1.5O4 coin cells incorporating two different samples. Both samples exhibit a semicircle in the high-frequency range and an inclined line with angle over 45° before room temperature cycle performance and two semicircle in the high and middle-frequency range and an inclined line close to 45° in the low-frequency range after 200th cycle. The high-frequency semicircle represents the resistance of solid electrolyte interphase (RSEI), i.e., the resistance of the SEI layer on the anode electrode or CEI layer on the cathode electrode.10 The middle-frequency semicircle usually represents the charge transfer resistance (Rct), and the inclined line corresponds to the lithium-ion diffusion process, called Warburg diffusion.
Fig. 5 Nyquist plots of Li/electrolytes/LiNi0.5Mn1.5O4 cells measured at open circuit voltage before (a) and after 25 °C cycle (b) at 2.5C. |
Single crystal is mainly composed of a single primary particle, while polycrystalline is formed by the agglomeration of many nano/micron-sized primary particles into micron-sized secondary particles. The agglomerated polycrystalline materials have shortened diffusion lengths within their primary particles and an increased number of pores, which accelerates Li+ transport, however, single crystalline materials have reduced surface areas, reduced grain boundaries, and more integrated crystal structures, similar to LiCoO2. Before cycling, different crystallinity of the electrode has different press density, porosity, and contacting area with the electrolyte, thus forming different SEI impedance. With electrochemical charging and discharging cycles, grain boundary cracks, voids, metal dissolution, electrolyte decomposition, gas generation, phase transition or other degradation behaviors of the electrode may occur, leading to the changes of solid state electrolyte impedance, charge transfer impedance and lithium ion diffusion kinetics. The EIS measurements were carried out to further understand the impact of these different material design on the kinetic and mass transport behaviors.11 The equivalent circuits for Fig. 5 were fitted by Nova 2.1 software and fitted data were in Table 2. Before cycle (Fig. 5a and Table 2), Li/LiNi0.5Mn1.5O4 coin cell incorporating SLNMO sample showed a little larger RSEI of 47.8 Ω than 36.3 Ω for PLNMO sample. After 200 cycles (Fig. 5b and Table 2): the RSEI decreased a little to 45.9 Ω for the cell with SLNMO, however increased to 46.6 Ω for the cell with PLNMO. A dramatically different appeared that Rct was 48.8 Ω for SLNMO, much lower than that was 120.3 Ω for PLNMO. Because chemical reaction involving intercalation/deintercalation continually occurs in/on the electrodes as the number of cycles increase, the internal resistance of RSEI and Rct become larger, as a result the electrochemical polarization is obvious. The EIS results demonstrated that SLNMO sample could form lower RSEI and Rct than PLNMO sample during cycling, thus decreased the impedance on the electrode/electrolyte interphase, therefore exhibited smaller electrochemical polarization or voltage difference.
The cycle performance depends not only on the interfacial reaction resistance of the redox reaction on both the cathode and anode electrode,12,13 but also on the speed of Li-ion diffusion.
Lithium-ion diffusion coefficient is one of the most crucial parameters determining the kinetics of intercalation compounds. Usually it can be quantitatively evaluated by electrochemical techniques such as galvanostatic intermittent titration technique (GITT), potentiostatic intermittent titration technique (PITT), electrochemical impedance spectroscopy (EIS), or cyclic voltammetry (CV).14,15 In Fig. 6, lithium-ion diffusion coefficient at different state of charge (SOC) before room temperature cycle performance for Li/LiNi0.5Mn1.5O4 coin cells incorporating two different samples was evaluated via GITT method,16 which can be obtained by the equation:
DLi+ = 4/π(I0Vm/FA)2[(dE/dx)/(dE/dt1/2)]2 | (4) |
Fig. 6 GITT curves of Li/electrolytes/LiNi0.5Mn1.5O4 cells (a), and DLi+ at different state of charge (b) and discharge (c) from GITT method. |
The GITT results proved that lithium-ion diffusion behavior is faster for SLNMO sample. During the charge process from x = 0.22–1.0 in Fig. 6b, lithium-ion diffusion coefficient increased initially and decreased afterwards, then increased again and decreased afterwards, resembling closely in the interlayer spacing variation tendency of lattice parameter, just like the capital letter of “M”, and the fastest lithium-ion diffusion occurred at the state of charge (SOC) ∼ 50% (x = ∼0.5), with DLi+ = 9.86 × 10−8 cm2 s−1 for SLNMO sample and DLi+ = 6.16 × 10−8 cm2 s−1 for PLNMO sample. During the discharge process from x = 1.0–0 in Fig. 6c, lithium-ion diffusion coefficient decreased initially and increased afterwards, then decreased again and increased afterwards, just like the capital letter of “W”, and the fastest lithium-ion diffusion occurred at the state of discharge (SOD) ∼ 100% (x = ∼0), with DLi+ = 3.16 × 10−6 cm2 s−1 for SLNMO sample and DLi+ = 1.28 × 10−6 cm2 s−1 for PLNMO sample. Lithium-ion diffusion coefficient was cross-checked by electrochemical impedance spectroscopy (EIS) method, and the fastest lithium-ion diffusion also occurred at the state of charge (SOC) 50% (x = 0.5), with DLi+ = 2.12 × 10−9 cm2 s−1 for SLNMO sample and DLi+ = 1.56 × 10−9 cm2 s−1 for PLNMO sample.
The heat flow (Φ) measured by DSC is composed of two parts, one is the sensible heat flow caused by the temperature rise, the other is the latent heat flow caused by the internal structure change of samples. According to formula (5), the heat flow is proportional to the heating rate. The higher the heating rate, the more significant the thermal effect. In order to make the thermal effect more obvious, DSC instrument experts recommend that our heating rate is 10 °C min−1. The occurrence time of thermal runaway is generally very short and the heating rate is very fast. DSC test also attempts to simulate the scene of thermal runaway of battery. Large volume crucible can hold more samples, which is beneficial to measure the weak thermal effect. Smaller crucible is beneficial to separate overlapping thermal effects due to better thermal conductivity and smaller temperature gradient, so we chose to use 25 μL crucibles.
Φ = m × Cp × β + ΔHp × (dα/dt) | (5) |
Fig. 7 showed the DSC heat flow curves of the two samples' powder at electrochemically delithiated of 100% SOC with the same high voltage organic electrolyte. The decomposition of SLNMO started at higher onset temperature at 218 °C with less heat released than PLNMO at 200 °C. The peak temperature and heat generation of exothermic reaction profile for SLNMO was 227.8 °C and 438.6 J g−1, higher than that of the corresponding peak of PLNMO at 212.7 °C and 412.8 J g−1, thus confirmed the excellent thermal stability of SLNMO against PLNMO. Theoretically, Mn3+ content of SLNMO was higher, so the stable Mn4+ was less, and the thermal stability would be worse.17 Conversely, the better thermal stability of SLNMO could be ascribed to less oxygen release18 accompanied by higher Mn3+, and relatively inactive big primary grain particle relieving the intense reaction with electrolyte at high voltage.
Compared with the obvious thermal effect of 5V spinel materials, the thermal effect of the electrolyte was too weak to show under the heating rate of 10 °C min−1. Because the amount of electrolyte added was too small considering serious thermal runaway are likely to occur if large amount of electrolyte added. We conducted DSC under the heating rate of 10 °C min−1 with bare electrolyte and increasing the amount of electrolyte to 6 μL, and the exothermal peak of electrolyte can be apparent. The reason why the electrolyte with small amount did not have apparent exothermal peak could be attributed to the higher thermal stability of the salt–solvent pairs used in the electrolyte. In addition, fluorinated ether and carbonate solvents could have superior thermal resilience over non-fluorinated carbonate and ether solvents.19 Moreover, solvent additives can help passivate the reactive surface of the charged LNMO cathode, thereby limiting the side reactions between the electrode surfaces and the electrolyte, thus improving thermal stability.
This journal is © The Royal Society of Chemistry 2023 |