Next Article in Journal
Can Airport and Social Waste Reduction Measures Have a Synergistic Impact on Passenger Behavior?
Previous Article in Journal
Renewable Energy Community: Opportunities and Threats towards Green Transition
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Study and Property Characterization of LiMn2O4 Synthesized from Octahedral Mn3O4

1
School of Materials and Metallurgy, Guizhou University, Guiyang 550025, China
2
Guizhou Provincial Engineering Technology Research Center of Manganese Materials for Batteries, Tongren 554300, China
3
Guizhou Provincial Key Laboratory of Metallurgical Engineering and Energy Saving, Guiyang 550025, China
4
School of Materials and Energy Engineering, Guizhou Institute of Technology, Guiyang 550002, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Sustainability 2023, 15(18), 13858; https://doi.org/10.3390/su151813858
Submission received: 14 August 2023 / Revised: 11 September 2023 / Accepted: 14 September 2023 / Published: 18 September 2023

Abstract

:
The structure of Mn3O4 with an octahedron structure was similar to that of LiMn2O4, and the lithium manganate prepared with it had good electrochemical performance. During the preparation of octahedron Mn3O4, the effects of the pH regulator, temperature, and reaction pH on its morphology, specific surface area, and other properties were studied in this paper. LiMn2O4 was prepared from Octahedron Mn3O4 obtained by using better technology. The effects of calcination time and temperature on the physicochemical and electrochemical properties of LiMn2O4 were studied. The research results indicated that the optimal synthesis conditions for Mn3O4 were as follows: ammonia water was used as a pH regulator and complexing agent, reaction pH was 8, reaction temperature was 80 °C, reaction time was 12 h, and oxygen flow rate was 3 L∙min−1. The LiMn2O4 synthesized had a good octahedron morphology when the calcination temperature was 800 °C and the calcination time was 10 h. The first discharge-specific capacity was 121.9 mAh∙g−1 at a current density of 0.2 C, the discharge-specific capacity was 114.1 mAh∙g−1 after 100 cycles, and the capacity retention rate was 93.6%. Therefore, the lithium manganate prepared by using octahedron manganous oxide had good electrochemical reversibility and a good application prospect.

1. Introduction

Battery-grade Mn3O4 could be used to prepare cathode materials for lithium-ion batteries and sodium-ion batteries. As one of the main materials for lithium-ion positive electrodes, LiMn2O4 has been mainly prepared through the high-temperature calcination of electrolytic MnO2 in industry [1,2,3,4]. More and more studies have shown that LiMn2O4 prepared using Mn3O4 has better electrochemical performance compared to electrolytic MnO2 [5]. Due to the high electrolysis cost, uneven particle size, and poor dispersion of MnO2, the LiMn2O4 crystal size is uneven, resulting in low specific capacity and poor cycling performance. The LiMn2O4 crystal synthesized using Mn3O4 has a small degree of distortion, uniform morphology, and uniform size. A good morphology is conducive to lithium-ion unembedding, reducing the crystal structure changes caused by unembedding and improving the cycling performance of LiMn2O4. Therefore, Mn3O4 has the potential to replace electrolytic MnO2 [6,7,8].
As the most basic cathode material, LiMn2O4 has the advantages of green environmental protection, low cost, and good safety performance [9,10,11,12]. However, due to the disproportionation and Jahn-Teller effect [13], Mn dissolves, which reduces the cycle performance of LiMn2O4 and limits its large-scale production and application [14,15,16]. By controlling the reaction conditions, the morphology can be optimized to achieve the purpose of improving the cycle performance of the cathode material [17]. Cao et al. [18] used MnO2 and LiOH to obtain LiMn2O4 nanowires under high-temperature calcination in the air environment. At a 0.1 C current, the first discharged specific capacity was 132 mAh·g−1, but the cycle capacity decayed rapidly. Niraj et al. [19] prepared one-dimensional LiMn2O4 with a diameter of 10–50 nm using ultrafine α-MnO2 nanorods. At a current of 0.05 C, the initial discharge capacity reached 130 mAh·g−1, and the retention rate was 98% after 100 cycles. The corresponding size and morphology were synthesized to alleviate the capacity decay of LiMn2O4 by controlling the reaction conditions [20]. Octahedral Mn3O4 and LiMn2O4 had the same spinel structure, which could reduce crystal structure changes during high-temperature roasting, facilitate lithium ion disembedding, reduce the effect of lithium-ion disembedding on crystal structure, and improve the high-temperature performance, cycling performance, and gram specific capacity of LiMn2O4 [21,22].
Manganese trioxide (Mn3O4) is a natural black manganese ore with a theoretical manganese content of 72.03%, mainly consisting of reddish-brown and brownish-yellow. Mn3O4 is a type of spinel-structured Mn2+ [Mn3+]O4, where Mn2+ and Mn3+ are distributed at different lattice positions. Therefore, Mn3O4 could be expressed in the form of MnO∙Mn2O3 [23]. As an important industrial raw material, Mn3O4 has extremely wide applications; it is widely used in industries such as soft magnetic materials, supercapacitors, new energy materials, and pigments [24,25]. Mn3O4 has spherical, octahedral, nanorod and, hollow spherical shapes. Mn3O4 has a wide variety of morphologies, different sizes, and uneven dispersion, which have a significant impact on the electrochemical and cycling performance of the synthesized positive electrode material [26,27,28]. At present, according to the classification of raw materials, the preparation methods of Mn3O4 include the metal manganese method, manganese salt method, high valent manganese reduction method, etc. [29,30]. Wang et al. [31] synthesized Manganese(II) carbonate by using the precipitation method and prepared linear single crystal Mn3O4 with a particle size of 40–80 nm through high-temperature calcination. Although MnCO3 calcination could produce various morphologies of Mn3O4, its products would contain impurities such as Mn2O3. A high calcination temperature means high cost and high energy consumption. Wang et al. [32] prepared octahedron hollow Mn3O4 using KMnO4 and formamide as raw materials, the reaction was maintained for 12 h at 140 °C in a high-pressure reactor. Cao et al. [33] mixed KMnO4 and sodium citrate 1:1 and reacted the mixture for 10 h at 180 °C in a high-pressure reactor to produce high-purity octahedron Mn3O4. Although this method could obtain various morphologies of Mn3O4, it had a long reaction time, low production efficiency, high equipment requirements, and certain limitations in mass production. Ramulu et al. [34] used cellulose from waste tissue as a scaffold to deposit manganese precursors and then calcined them at high temperatures to prepare high-pore hollow microtubules of Mn3O4 using morphological methods. Mansournia et al. [35] directly synthesized Mn3O4 powder with excellent catalytic performance at room temperature by using manganese(II) chloride and ammonia as raw materials. Researchers also studied some new methods, but these methods had many shortcomings, such as severe agglomeration, low crystallinity, and complex operating procedures [36,37,38,39,40,41]. In this paper, battery-grade Mn3O4 was prepared through liquid phase oxidation of manganese sulfate solution. The Mn3O4 prepared by using this method had the advantages of low cost, abundant raw materials, easy control of grain morphology, and uniform particle size, which were beneficial for improving the electrochemical performance of LiMn2O4.

2. Experiment

2.1. Raw Materials and Instruments

The main reagents used in this experiment included NH3·H2O, NaOH, and C2H5OH, which were analytically pure and purchased from China, China National Pharmaceutical Group Chemical Reagent Co., Ltd. LiOH·H2O, C, Li, C5H9NO, (CH2CF2)n, LiPF6/(EC + EMC), and diaphragms were purchased from China, Guangdong Zhuguang New Energy Technology Co., Ltd. MnSO4 was produced in the laboratory from China manganese rhodochrosite after purification by acid leaching with sulfuric acid and meets the standard of battery-grade MnSO4.
The main instruments used in this experiment were as follows. The constant-temperature blast drying oven (DHG-101-4B) was produced by China Shanghai Langgan Experimental Equipment Co., Ltd. The digital display constant temperature water bath pot (HH-2J) was produced by China Changzhou Langyue Instrument Manufacturing Co., Ltd. The muffle furnace (SX4-10) was produced by China Tianjin Taist Instrument Co., Ltd. The electric mixer (LC-OES-120SH) was produced by China Shanghai Lichen Instrument Technology Co., Ltd. The oxygen concentrator (CR-P5W) was produced by China Keer Medical Technology Co., Ltd. The glovebox (S175) was produced by China Chengdu Delis Industrial Co., Ltd. The battery testing system (CT-4008T-5V) was produced by China Shenzhen Xinweier Electronics Co., Ltd. The button-type battery sealing machine (MSK-110) and manual slicer (MSK-T10) were all produced by China Shenzhen Kejing Zhida Technology Co., Ltd. The X-ray powder diffractometer (D8 ADVANCE) was produced by Brooke Company in Germany; The field emission scanning electron microscope (SU8020) was produced by Hitachi, Japan. The X-ray electron spectrometer (Thermo escalab 250Xi) was produced by Thermo Fisher & Company in the United States Massachusetts City. The laser particle size analyzer (Mastersizer 2000) was produced by Marvin Instruments Limited in the UK Marvin City.

2.2. Experimental Process

The experiment mainly included the following two steps:
The experimental process for preparing battery-grade Mn3O4 was as follows. The 1LMnSO4 solution was loaded into a beaker and placed in a constant-temperature water bath. When the solution was heated to a certain temperature, a pH regulator was added until the preset reaction pH was reached, and an oxidant was continuously introduced. The pH regulators were slowly added to maintain a stable solution pH value during the reaction process. After a certain period of reaction, filtration and washing were carried out. The filter residue was repeatedly washed with 60 °C distilled water until no white precipitate was generated when BaCl2 was added to the washing solution. The filter cake was dried at 105 °C in a constant temperature blast drying box for 12 h, removed, and ground for later use.
The preparation and electrochemical detection of lithium manganese oxide cathode material were as follows: Mn3O4 and LiOH with a certain lithium manganese molar ratio were weighed and ground in an agate mortar for 30 min before being loaded into a crucible. It was roasted in a muffle furnace in an air atmosphere according to the preset heating rate, roasting temperature, and insulation time. After the reaction ended, it was cooled in the furnace to obtain lithium manganese oxide powder. Finally, the physicochemical and electrochemical properties of lithium manganese oxide were tested.

2.3. Performance Characterization

(1)
X-ray diffraction analysis (XRD)
In this experiment, the XRD diffractometer (D8ADVANCE of Brooke Company) was used to detect and analyze the phase of the product. The radiation source was a Cu target, the scanning rate was 6°∙min−1, and the scanning range of 2θ was 10–90°. The diffraction peak of the sample was decided by comparing it with the PDF card of the known substance.
(2)
Scanning electron microscope (SEM)
The scanning electron microscope (SU8020 of Hitachi, Japan) was used to characterize the surface morphology of the samples at 15 kV with different magnifications, and SEM pictures were obtained.
(3)
X-ray photoelectron spectroscopy (XPS) analysis
The Thermo escalab (250Xi electron spectrometer of Massachusetts United States City) was used to analyze the chemical valence state of Mn3O4.
(4)
Particle size analysis
The Mastersizer 2000 laser particle size analyzer produced by Marvin Company in the UK was used to perform particle size detection and analysis on materials dispersed in the aqueous phase.
(5)
Electrochemical performance testing
Preparation of electrode slice. Firstly, polyvinylidene fluoride was dissolved in N-Methyl-2-pyrrolidone to prepare a binder with a concentration of 0.03 g·mL−1. After grinding the positive electrode material, acetylene black, and adhesive evenly in a mass ratio of 8:1:1, they were then coated on carbon-coated aluminum foil. Secondly, they were dried in a constant-temperature blast drying oven for 30 min and then dried in a 110 °C constant-temperature vacuum drying oven for 12 h. Next, the aluminum foil was taken out and made into the positive electrode plate using a manual slicing machine with a diameter of 12 mm punch. Finally, they were weighed and put into the glove box to assemble the battery for testing.
Button battery assembly. The half-cell was assembled in the order of negative case–lithium sheet–electrolyte–diaphragm–electrolyte–positive sheet–gasket–spring sheet–positive case in the glove box filled with argon (water and oxygen contents were less than 0.01 ppm). They were hydraulically sealed using a button-type battery sealing machine and were taken out and left to stand for 12 h before undergoing electrochemical performance testing.
Constant current charging and discharging test. The battery testing system (CT-4008T-5V) manufactured by China Shenzhen Newell Electronics Co., Ltd. was used to conduct constant current charging and discharging tests and rate performance tests on LiMn2O4 positive electrode material. The current was calculated based on 148 mAh·g−1 of the LiMn2O4 standard specific capacity. The voltage range of the constant current charge–discharge test was 3–4.3 V, and the rate range of the rate performance test was 0.2 C, 0.5 C, 1 C, 2 C, 5 C, and 10 C.

3. Results and Discussion

3.1. Thermodynamic Analysis of Mn3O4 Preparation from MnSO4

Based on relevant thermodynamic data, the E-pH diagram of the Mn-NH3-SO42−-H2O system was drawn, as shown in Figure 1.
From Figure 1, it can be seen that when ammonia was added to a complex with Mn2+, it could be directly oxidized in the pH range of 6.7–10.5. Because Mn2+ could form stable complex regions with different amounts of ammonia, the pH range of Mn2+ oxidation became wider. This was more conducive to controlling the reaction. In the Mn-NH3-SO42−-H2O system at 25 °C, there existed a stable thermodynamic region for Mn3O4. Mn2+ could be oxidized to Mn3O4 in three ways.
(1)
Direct oxidation. When the pH was between 6.76 and 8.24, the voltage (E) was between 0.254 and 0.433, Mn2+ could be directly oxidized to Mn3O4, and the reaction formula was as follows.
6 Mn 2 + + 6 H 2 O + O 2 2 Mn 3 O 4 + 12 H +
(2)
Complex oxidation. When pH was between 8.24 and 10.57, the coordination number of the ammonia manganese complex was different, but the principle of their oxidation to Mn3O4 was the same. The reaction formula could be expressed as follows.
6 [ Mn NH 3 x ] 2 + + O 2 + 6 H 2 O = 2 Mn 3 O 4 + x NH 3 + 12 H +
The oxidation potential range was as follows when x = 1, −0.045 < E < 0.343. The oxidation potential range was as follows when x = 2, 0.122 < E < 0.343. The oxidation potential range was as follows when x = 3, −0.086 < E < 0.343. It could be directly oxidized to Mn3O4 within the corresponding oxidation potential.
(3)
Precipitation oxidation. When pH was between 10.57 and 14, Mn2+ could form Mn(OH)2 precipitate, the oxidation potential range was between −0.327 and 0.203, and Mn(OH)2 could be oxidized to Mn3O4. The reaction formula was as follows.
Mn 2 + + 2 OH Mn ( OH ) 2
6 Mn ( OH ) 2 + O 2 2 Mn 3 O 4 + 6 H 2 O
Among the three oxidation methods, the direct oxidation method had a narrow reaction pH range. After the addition of ammonia and Mn2+ complexation, the reaction pH range of the complexation oxidation method became wide, and the oxidation rate of Mn2+ slowed down, which was conducted to the nucleation and growth of Mn3O4 grains. The precipitation oxidation method needed to filter, wash, and adjust the Mn(OH)2 slurry, and the reaction process needed to isolate the air to avoid the Mn(OH)2 on the surface of the solution being oxidized to Mn3O4 first.
From Figure 1, it can be seen that the byproducts such as Mn2O3 and MnO2 could also be generated under the corresponding redox potential in the reaction system. Therefore, when using ammonia as a pH regulator and complexing agent to control the oxidation rate of Mn2+ in this experiment, to prepare high-purity and high-performance battery-grade Mn3O4, the reaction conditions should be strictly controlled to suppress the generation of side reactants such as MnO2, Mn2O3, and Mn2(OH)2SO4.

3.2. Effect of pH Regulator on Mn3O4

When the content of Mn2+ was 80 g∙L−1, the reaction pH was 8, the reaction temperature was 70 °C, the reaction time was 12 h, the oxygen flow rate was 2 L∙min−1, and the effects of pH regulators such as NaOH (120 g∙L−1) and ammonia (12%) on the physicochemical properties of Mn3O4 were explored.
The microstructure of Mn3O4 prepared with NaOH and ammonia as pH regulators is shown in Figure 2. As shown in Figure 2, pH regulators had a significant impact on the morphology of Mn3O4. The Mn3O4 grains prepared by NaOH were mostly sheet-like, with large voids, no fixed morphology, and poor crystallinity. The Mn3O4 grains prepared by using ammonia were complete, most of which were octahedron-shaped, with uniform size and small voids. When ammonia was as a pH regulator, on the one hand, the complex was formed between NH3 and Mn2+; they could slow down the oxidation rate of Mn2+ and control the nucleation rate of Mn3O4. On the other hand, ammonia manganese complexes existed in the form of ion clusters and had a certain spinel structure, resulting in a regular and uniform grain shape and size of Mn3O4.
The manganese content, D50, and surface areas of manganese tetroxide prepared with different pH regulators are shown in Table 1.
According to Table 1, the manganese content in Mn3O4 prepared with NaOH and ammonia water was 71.08% and 71.32%, respectively. When NaOH served as a pH regulator, Mn2+ was converted into Mn(OH)2 and then oxidized to Mn3O4. The generation and oxidation of Mn(OH)2 occurred simultaneously, resulting in some Mn(OH)2 being tightly encapsulated by the Mn3O4 product layer. The diffusion of oxygen to the surface of unreacted Mn(OH)2 became a limiting link, resulting in a lower Mn content in the product. When ammonia was used as a pH regulator, the ammonia manganese complex was a soluble complex, which was in full contact with oxygen, so the Mn content in the product was higher. From Table 1, it could be seen that the D50 particle sizes of Mn3O4 prepared with NaOH and ammonia water were similar, and they were 3.68 μm and 4.67 μm, respectively. However, there was a significant difference in specific surface area; the specific surface areas were 22.18 m2∙g−1 and 3.49 m2∙g−1, respectively. Using ammonia as a pH regulator to form a complex with Mn2+ not only slowed down the nucleation rate of Mn3O4 but also helped to oxidize Mn2+ into a single, regular morphology. The sheet-like Mn3O4 synthesized by using NaOH was stacked and disordered, resulting in a larger specific surface area. A smaller specific surface area was beneficial for improving the activity of the positive electrode material and improving the cycling and rate performance of the battery. Therefore, it was better to choose ammonia water as a pH regulator to prepare Mn3O4.

3.3. Effect of Reaction pH on Mn3O4

When the content of Mn2+ was 80 g∙L−1, the ammonia concentration was 8%, the reaction temperature was 70 °C, the reaction time was 12 h, and the oxygen flow rate was 3 L∙min−1, The effects of reaction pH values of 7, 7.5, 8, 8.5, and 9 on the physicochemical properties of Mn3O4 were investigated.
The microstructure of Mn3O4 prepared under different reaction pH conditions is shown in Figure 3. It could be seen from Figure 3a–e that the microscopic morphology of Mn3O4 was still composed of octahedron particles under different reaction pH, but their particle size and uniformity changed. The aggregation phenomenon of Mn3O4 microparticles was severed and they had a few voids and small particles, all around 100 nm when the reaction pH was 7. The particles became larger, the dispersity increased, the octahedron shape was completed, and the average particle size was about 250 nm when the reaction pH was 7.5. The octahedron particles were the largest, the average particle size was about 280 nm, the particle size was uniform, and the boundary between particles was cleared when the reaction pH was 8. When the pH was 8.5 and 9, the agglomeration phenomenon increased, particle boundaries became blurred, voids and particles became smaller, the particle sizes were around 200 nm, and many small particles appeared. When the reaction pH was low, the oxygen oxidation potential was high, and the oxidation difficulty of the ammonia manganese complex increased, slowing down the oxidation rate of the ammonia manganese complex. However, it reduced the precipitation of components for crystal nucleus growth, leading to hindrance of crystal nucleus growth. Therefore, severe aggregation and narrow gaps occurred. When the reaction pH was low, the crystal particles were small. When the reaction pH was high, the formation rate of the ammonia manganese complex was fast, the oxidation potential was low, and the ammonia manganese complex was easy to oxidize, which increased the supersaturation rate of Mn3O4 solute and was conducive to the formation of the crystal core. The nucleation rate of Mn3O4 was much higher than the growth rate of crystal nuclei, the crystals were dispersed, and a large number of crystal nuclei were formed, making it difficult to grow. Small particles adhered to the surface of large particles. Finally, the nucleation rate and growth rate showed mutual fluctuation.
The analysis results of Mn content, D50 particle size, and specific surface area of Mn3O4 prepared with different reaction pH are shown in Figure 4. The Mn content decreased with the increase in reaction pH. When the reaction pH rose, it was necessary to increase the acceleration of ammonia drops to maintain the stability of the reaction pH, and the concentration of ammonium and Hydroxide ions in the solution increased accordingly. On the one hand, some impurity ions generated insoluble hydroxides and entered the impurity phase. On the other hand, it promoted the formation of Mg (NH4)2(SO4)2 and Mn2(OH)2SO4 through side reactions. The byproducts were tightly wrapped and entered the precipitation, leading to an increase in sulfur content and a decrease in Mn content.
From Figure 4b, the D50 particle size of Mn3O4 gradually decreased, and the specific surface area gradually increased with the increase in reaction pH. When the reaction pH was low, the supersaturation and nucleation rate were also low, the agglomeration sites were few, and the grains attracted and agglomerated with each other to form large particles. When the reaction pH was too high, the solute supersaturation increased, the grains were more dispersed, the particle size was finer, a large number of Mn3O4 crystal nuclei were formed, and the agglomeration points increased, resulting in a large number of agglomerated particles and small particle size, due to which the particle size of Mn3O4 decreased slightly and the specific surface area increased. Therefore, the nucleation rate of Mn3O4 could be controlled, the number of crystal nuclei and aggregation points could be reduced, and the agglomeration of grains into large particles could be promoted by controlling the reaction pH. A comprehensive evaluation of each performance indicated that a pH of 8 was more appropriate.

3.4. Effect of Reaction Temperature on Mn3O4

When Mn2+ was 80 g∙L−1, the ammonia concentration was 8%, the reaction pH was 8, the reaction time was 12 h, and the oxygen flow rate was 3 L∙min−1. The effects of reaction temperatures of 25 °C, 50 °C, 60 °C, 70 °C, and 80 °C on the physicochemical properties of Mn3O4 were investigated.
The microstructure of Mn3O4 prepared at different temperatures is shown in Figure 5. The microstructure of Mn3O4 underwent significant changes with different reaction temperatures. When the reacting temperature was at a room temperature of 25 °C, the majority of Mn3O4 grains were spherical and rod-shaped small particles, at only a few nanometers in size, the agglomeration was serious, and there were large voids within the structure. When the temperature rose to 50 °C and 60 °C, the crystal nucleus grew fast and the grain size became larger. Their particle sizes were about 160 nm and 200 nm, respectively, but the grain size was uneven, a large number of particles were agglomerated into small particles, and there were large gaps between particles. When the temperature rose to 70 °C and 80 °C, the microstructure was all octahedral grains, the grain size was about 280 nm and 300 nm, the grain surface was smooth, they agglomerated into large particles, and the gap between particles was small. When the temperature was low, the nucleation rate was small, the crystal nucleus grew slowly, it was difficult to grow, and a large number of small grains agglomerated. With the increase in temperature, the oxidation potential of oxygen decreased, and the ammonia manganese complex was more easily oxidized. The growth rate of the crystal nucleus was accelerated, the grain shape was more regular, the size was uniform, the thermal movement of particles was intensified, and the particles could be agglomerated into compact large particles.
Some physicochemical properties of Mn3O4 prepared at different reaction temperatures are shown in Figure 6. As shown in Figure 6a, the Mn content increased with increasing temperature. On the one hand, as the temperature increased, the Gibbs free energy of the system reaction decreased, promoting the conversion of Mn2+ to ammonia manganese complex, increasing the oxidation rate of the complex, and reducing the generation of byproduct basic MnSO4. On the other hand, as the temperature increased, the oxygen oxidation potential decreased, which could oxidize some basic MnSO4 and weaken the peroxidation of oxygen, reducing the generation of high-valent manganese oxides. Therefore, the Mn content gradually increased with the increase in temperature.
From Figure 6b, as the reaction temperature increased, the D50 particle size gradually increased and the specific surface area gradually decreased. When the temperature was low, the crystal growth rate was slowed and small grains were formed, so the particle size of the agglomerated particles was smaller. The supersaturation and nucleation rate increased with increasing temperature, which promoted the growth of grains. At the same time, with the increase in temperature, the collision probability of grains increased due to the intensification of thermal motion, and the grains were agglomerated to form large particles. As a result, the particle size increased and the specific surface area decreased.

3.5. Characterization Analysis of Mn3O4 Product

According to the requirements of particle size and specific surface area of Mn3O4 for lithium batteries, Mn3O4 with regular morphology, large particle size, and small specific surface area was selected for characterization and analysis. Based on the above research results, the preparation of Mn3O4 under the following conditions was suitable. The Mn2+ content was 80 g∙L−1, ammonia water was the regulator, the reaction pH was 8, the reaction temperature was 80 °C, the reaction time was 12 h, and the oxygen flow rate was 3 L∙min−1.

3.5.1. Phase Analysis

X-ray diffraction analysis was performed on the product Mn3O4, and its spectrum is shown in Figure 7.
It can be seen from Figure 7 that the product Mn3O4 was a solid powder, the baseline of the diffraction peak was flat, the peak pattern was narrow, and the peak intensity was high, so the crystallinity of the product was good. The cell size of Mn3O4 was a = b = 0.576 nm, c = 0.945 nm, and the crystal volume was 0.314 nm3. The grain size was around 290 nm according to the Scherrer formula [42], which is consistent with the result of the previous electron microscopy scans.

3.5.2. Chemical Element Analysis

The chemical elemental analysis of Mn3O4 was performed, the Mn content was detected by using the chemical titration method, and the remaining impurity elements were detected according to the method of the Mn3O4 standard for lithium batteries (YB T4736-2019). The results are shown in Table 2.
From Table 2, the Mn content and impurity element content could meet the product standard of Mn3O4 for lithium batteries.

3.5.3. Valence Analysis

The average valence of Mn3O4 was +2.7, and there were two cases. When it existed in the form of MnO∙Mn2O3, it exhibited +2 and +3 valence states. When it existed in the form of 2MnO·MnO2, it exhibited +2 and +4 valence states. XPS detection was performed to analyze the valence states, and the results are as shown in Figure 8.
The characteristic peaks of Mn 2p and Mn 3s appeared in the full spectrum of the XPS map. The characteristic peaks of Mn 2p and Mn 3s were fitted. The fitting curves of Mn 2p showed the peak of Mn2+ 2p1/2, Mn2+ 2p3/2, Mn3+ 2p1/2, and Mn3+ 2p3/2. The fitting curves of Mn 3s showed the peak of Mn2+ 3s and Mn3+ 3s. ∆E = 5.9 eV for Mn2+ 3s and ∆E = 5.6 eV for Mn3+ 3s. According to the relevant literature reports [43,44], it could be judged that the prepared Mn3O4 could be expressed as MnO·Mn2O3.

3.5.4. Particle Size Distribution

The particle size of the product was analyzed, and the result is shown in Figure 9.
As shown in Figure 9, the particle size of Mn3O4 was normally distributed. D10, D50, and D90 were 3.669 μm, 6.17 μm, and 9.969 μm, respectively. The particle size distribution of Mn3O4 was good, indicating that the particle size was uniform.

3.6. Effect of Roasting Time on LiMn2O4

Mn3O4 prepared under the above optimal conditions was a manganese source. The molar ratio of lithium to manganese was 0.5, and they were mixed evenly and roasted in a muffle furnace at 800 °C for a certain time. After the reaction was completed, the sample naturally cooled in the furnace. The effects of roasting times of 4 h, 6 h, 8 h, 10 h, and 12 h on the cathode material LiMn2O4 were investigated.

3.6.1. The Roasting Time Phase and Morphology

The X-ray diffraction patterns of LiMn2O4 prepared under different roasting times are shown in Figure 10.
From Figure 10, the characteristic peaks of the five samples corresponded to the characteristic peaks of the standard card of LiMn2O4, basically indicating that these samples were all spinel LiMn2O4 with an Fd-3m space group. The diffraction peaks of LiMn2O4 prepared at different times were sharp, the baseline was stable, and no impurity peaks appeared, indicating that the product had good crystallinity and high purity. With the increase in roasting time, the characteristic peak became more and more sharp, and the peak value increased, indicating that the crystallinity of the product became better and better.
The crystal structure analysis of LiMn2O4 samples was performed, and the results are shown in Table 3.
As could be seen from Table 3, the lattice constants and cell volume first decreased and then increased with increasing roasting time, and the R-values changed in the same trend. When the reaction time was short and the reaction was incomplete, lithium and manganese could not fully react to form a new phase, and some reactants did not fully react. The reaction was completely carried out until 8 h, the new phase was completely generated with good crystallinity, and the degree of distortion was small. The reaction time was prolonged, small particles fused to grow, and the distortion increased.
The microstructure of LiMn2O4 prepared at different calcination times is shown in Figure 11.
From Figure 11, the LiMn2O4 particles formed and grew gradually with the increase in roasting time. When the roasting time was 4 h, there were a small number of particles that did not participate in the reaction, so the sample showed small particles and irregular shapes. As the reaction time increased, when it reached 6 h, 8 h, and 10 h, the particles were completed, the size was uniform, the gaps of grains were small, their density was high, and the octahedron had sharp edges and corners. The increase in time made the crystalline phase transformation complete, and the crystallinity increased. When the reaction time reached 12 h, the grains continued to grow after the crystalline phase transformed, and some of the products deoxygenated, which led to the lack of stability inside the material and the increase in distortion. Therefore, the roasting time of 10 h was more suitable.

3.6.2. Electrochemical Properties

The LiMn2O4 prepared with a lithium manganese ratio of 0.5 and different roasting times was used as the cathode material to prepare the CR2032 button half-cell. Cycling performance was tested at 0.2 C, multiplication performance was tested at different current densities from 0.2 C to 10 C, and the results are shown in Figure 12.
From Figure 12a, the charging and discharging curves both showed a voltage platform around 3.95 V and 4.1 V at a current density of 0.2 C, indicating the existence of two reversible redox reactions for the process. With the increase in roasting time, the first discharge-specific capacity gradually increased. The first discharge-specific capacity of LiMn2O4 roasted for 4 h was the lowest at 101.1 mAh∙g−1, while that of LiMn2O4 roasted for 10 h was the highest at 121.1 mAh∙g−1.
As shown in Figure 12b, the first charge/discharge-specific capacities of the five samples were 101.1mAh∙g−1, 110.5 mAh∙g−1, 117.5 mAh∙g−1, 121.9 mAh∙g−1, and 115.2 mAh∙g−1 at a 0.2 C current density, and the retention rates of specific capacity were 72.4%, 92%, 92.5%, 93.6%, and 91.5% after 100 cycles, respectively. Among them, the specific capacity decay of the sample was the fastest after roasting for 4 h, and the cycling performance was the best after roasting for 10 h. From Figure 12c, after being charged and discharged at different current densities, the specific capacitance values of the five samples could still return to the specific capacitance values at the corresponding current density of 0.2 C. The sample roasted for 10 h had a specific capacity of about 87 mAh∙g−1 after charging and discharging at a current density of 10 C. When the current density returned to 0.2 C, the specific capacity became 110 mAh∙g−1. It was shown that the LiMn2O4 obtained after 10 h roasting had good multiplicative properties and good electrochemical reversibility.
When the roasting time was too short, the reaction of the reactant was not sufficient, the crystallinity of the product was poor, and the crystal cell structure was unstable. Therefore, the first charge–discharge capacity of the product was low, its cyclic decay was fast, and its rate performance was poor. When the roasting time was extended, the reaction was sufficient, the crystal phase transformation was complete, the product particle size was uniform, the structure was stable, it was conducive to the removal of lithium ions, and the high current had little damage to the crystal structure. Therefore, the first charge–discharge had a higher specific capacity, slower capacity decay, and good rate performance. When the reaction time continued to extend, on the one hand, it caused oxygen deficiency in LiMn2O4, and on the other hand, small particles melted and grew into large particles, leading to an increase in distortion and structural instability. As a result, specific capacity decreased gradually. In summary, LiMn2O4 had good electrochemical properties after roasting for 10 h.

3.7. Effect of Roasting Temperature on LiMn2O4

The molar ratio of lithium to manganese was 0.5, and the sample was roasted at a certain temperature for 10 h and then cooled in the furnace. The effects of roasting temperatures of 500 °C, 600 °C, 700 °C, 800 °C, and 850 °C on LiMn2O4 were investigated.

3.7.1. The Roasting Temperature Phase and Morphology

X-ray diffraction analysis was performed on LiMn2O4 prepared at different calcination temperatures, and the results are shown in Figure 13.
As could be seen from Figure 13, the diffraction peaks of the products synthesized at different temperatures were basically the same, and their characteristic peaks were consistent with those of the spinel structure LiMn2O4 standard card, all of which had the Fd-3m space group. There were a few Mn3O4 spurious peaks at 500 °C and 600 °C, indicating that some Mn3O4 failed to react at lower temperatures. When the temperature reached 850 °C, some impurity peaks also appeared, which was caused by the deoxygenation reaction of LiMn2O4 to generate LiMn2O4−x. When the temperature was 700 °C and 800 °C, no impurity peak appeared, the baseline was stable, the characteristic peak was sharp, and the peak value of the characteristic peak increased. It was indicated that the material LiMn2O4 had high purity and good crystallinity.
The crystal structure analysis of the lithium manganate samples was performed, and the results are shown in Table 4.
As can be seen from Table 4, the lattice constant and cell volume of LiMn2O4 first decreased and then increased with the increase in temperature, both of which were smaller than the lattice constant of the standard LiMn2O4 (the standard lattice constant of LiMn2O4 is a = 0.8245 nm, V = 0.5609 nm3). The ratios of (311) to (400) changed in the same pattern. At a low temperature of 500 °C, it failed to melt completely, resulting in intergranular adhesion, larger lattice constants, and a large degree of distortion. There was enough Mn3O4 to react with LiOH at 600 °C, so the intergranular boundaries were cleared, and the lattice constants became smaller. As the reaction temperature increased, the energy of reactants increased, the percentage of activated molecules increased, and the number of effective collisions of reacting molecules increased, so the grains kept growing, and the lattice constants and cell volume kept increasing. When the temperature was at 800 °C, the cell volume was small, the R-value was suitable at 1.016, the distortion was small, the cell structure was stable, which was favorable for the deintercalation of lithium-ions, and the sample had a better electrochemical basis.
The microstructure of LiMn2O4 prepared at different roasting temperatures is shown in Figure 14.
From Figure 14, it can be seen that the LiMn2O4 grains formed and grew gradually with the increase in temperature. The boundary between the grains was blurred, and many edges and corners had not yet grown completely when the temperature was 500 °C. When the temperature rose to 600 °C and 700 °C, the grains were formed, and they were small and evenly distributed, but there were many voids in the structure. When the temperature rose to 800 °C, the morphology of the particles was a regular octahedron, the grains grew up, the size of the particles was uniform, the boundaries were obvious, the voids became fewer, the particles were closely agglomerated, and the average particle size was about 400 nm. When the temperature was increased to 850 °C, the grains continued to grow, but the grains exhibited a polyhedral growth trend, which was probably due to structural changes caused by the deoxygenation of LiMn2O4 at this temperature. The increase in temperature resulted in higher crystallinity and larger grains of LiMn2O4, closer aggregation between particles, and a more stable structure. This was beneficial for Li+ deintercalation, reducing the impact of deintercalation on grain structure and improving the cycling performance of LiMn2O4. However, if the temperature was too high, some oxygen atoms would detach and the crystal structure would collapse, affecting its performance.

3.7.2. The Electrochemical Properties

The LiMn2O4 prepared with the lithium manganese ratio of 0.5 and different roasting temperatures was used as the cathode material to prepare the CR2032 button half-cell. The cycling performance was tested at a 0.2 C current, the multiplier performance was tested at different current densities from 0.2 C to 10 C, and the results are shown in Figure 15.
From Figure 15a, when charging and discharging were at a current density of 0.2 C, LiMn2O4 prepared at different temperatures exhibited two voltage platforms, indicating two reversible oxidation and reduction processes during the charging and discharging process of the material. The first discharge-specific capacity of LiMn2O4 first increased and then decreased with the increase in roasting temperature. Its first discharge capacity was 102.1 mAh∙g−1, 108.4 mAh∙g−1, 113.8 mAh∙g−1, 121.9 mAh∙g−1, and 116.7 mAh∙g−1, and the first charge–discharge coulomb efficiency was above 90%. Due to the increase in temperature, the purity and crystallinity of the material were improved, promoting a more stable structure, reducing the degree of structural distortion, and reducing the impact of lithium-ion intercalation on the structure of lithium manganese.
According to Figure 15b, the capacity retention rates of the five samples after 100 cycles were 86%, 87.5%, 89%, 93%, and 91.3%. The cycling performance of this material first increased and then decreased with temperature, and then it decreased when it exceeded a certain temperature value. Increasing the reaction temperature was beneficial for increasing the crystallinity of particles, stabilizing the structure, and thus improving cycling performance and capacity retention rate. When the temperature was too high, the distortion of the structure increased as the particles became larger, which increased the diffusion distance of lithium ions and led to a decrease in capacity and cycling retention rate.
The rate curves of the LiMn2O4 material are shown in Figure 15c. Overall, LiMn2O4 could still return to a higher discharge-specific capacity after being charged and discharged at different current densities of 0.2C 10C 0.2C. After charging and discharging at a current density of 10 C, the discharge-specific capacity of the five samples was 79.6, 81.3, 82.9, 86, and 85.09 mAh∙g−1. The discharge-specific capacity of LiMn2O4 prepared at 800 °C was 110 mAh∙g−1 when the current density returned to 0.2 C. The above results indicated that the LiMn2O4 cathode material obtained by roasting at 800 °C had good electrochemical performance.

4. Conclusions

Battery-grade Mn3O4 could be prepared through liquid-phase complex oxidation of MnSO4. The optimal synthesis conditions were as follows. Ammonia water was used as a pH regulator and complexing agent in the experiment. The reaction was carried out at a pH of 8, a temperature of 80 °C, and a duration of 12 h. The oxygen flow rate during the reaction was 3 L∙min−1. The impurity elements in the sample met the standard for manganese tetroxide in lithium batteries. The D50 particle size was approximately 6.17 μm, and the chemical formula could be represented as MnO∙Mn2O3. The cathode material was prepared using octahedron Mn3O4 as the manganese source and LiOH as the lithium source, employing the high-temperature solid-state method. The product, LiMn2O4, was obtained under the following conditions: a lithium manganese molar ratio of 0.5, a roasting temperature of 800 °C, and a roasting time of 10 h. The resulting LiMn2O4 exhibited an Fd-3m spinel structure with good crystallinity, high purity, and a small cell size. When subjected to charge and discharge at a current density of 0.2 C, the first charge–discharge-specific capacity was 125.7 mAh∙g−1 and 121.9 mAh∙g−1, respectively, with a first charge–discharge coulombic efficiency of 96.7%. After 100 cycles, the specific discharge capacity was 114.1 mAh∙g−1, and the capacity retention rate was 93.6%. In the multiplier mode, the discharge-specific capacity was about 87mAh∙g−1 at a 10 C current density and 110 mAh∙g−1 when it returned to 0.2 C. The LiMn2O4 had good electrochemical performance.

Author Contributions

Conceptualization, H.W. (Hao Wang) and J.W.; methodology, H.W. (Hao Wang); software, S.W.; validation, H.W. (Hao Wang), J.W. and H.W. (Haifeng Wang); formal analysis, H.W. (Hao Wang); investigation, S.W.; resources, H.W. (Hao Wang); data curation, X.D., W.H. and J.L.; writing—original draft preparation, H.W. (Hao Wang) and S.W.; writing—review and editing, H.W. (Hao Wang), J.W. and H.W. (Haifeng Wang); visualization, S.W., X.D., W.H. and J.L.; supervision, F.L.; project administration, H.W. (Hao Wang) and J.W.; funding acquisition, H.W. (Haifeng Wang). All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by [Science and technology in Guizhou Province] [(2022)key020], Projects supported by [Science and technology in Guizhou Province] [(2023)243], [Major special projects in Guizhou Province] [(2022)003], and the [Tongren Science and Technology Plan Project] [(2021)13].

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Informed consent was obtained from all subjects involved in the study.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cui, X.; Du, S.; Zhu, K.; Geng, S.; Zhao, D.; Li, X.; Tang, F.; Li, S. Elevated electrochemical property of LiMn2O4 originated from nano-sized Mn3O4. Ionics 2018, 24, 697–706. [Google Scholar] [CrossRef]
  2. Park, S.H.; Ahn, H.S.; Park, G.J.; Kim, J.; Lee, Y.S. Cycle mechanism and electrochemical properties of lithium manganese oxide prepared using different Mn sources. Mater. Chem. Phys. 2008, 112, 696–701. [Google Scholar] [CrossRef]
  3. Chen, F.; Lin, J.; Chen, Y.; Dong, B.; Yin, C.; Tian, S.; Sun, D.; Xie, J.; Zhang, Z.; Li, H.; et al. Enhancement of electrochemical performance in lithium-ion battery via tantalum oxide coated nickel-rich cathode materials. Chin. Phys. B 2022, 31, 58101–58108. [Google Scholar] [CrossRef]
  4. An, Q.; Liu, Q.; Wang, S.; Liu, L.; Wang, H.; Sun, Y.; Duan, L.; Zhao, G.; Guo, H. Oxygen vacancies with localized electrons direct a functionalized separator toward dendrite-free and high loading LiFePO4 for lithium metal batteries. J. Energy Chem. 2022, 75, 38–45. [Google Scholar] [CrossRef]
  5. Zawrah, M.; El Fadaly, E.A.; Khattab, R.; Aly, M.; El Shafei, H. Synthesis and characterization of nano Mn3O4 and LiMn2O4 spinel from manganese ore and pure materials. Ceram. Int. 2020, 46, 17514–17522. [Google Scholar] [CrossRef]
  6. Jiang, J.; Du, K.; Cao, Y.; Peng, Z.; Hu, G.; Duan, J. Syntheses of spherical LiMn2O4 with Mn3O4 and its electrochemistry performance. J. Alloys Compd. 2013, 577, 138–142. [Google Scholar] [CrossRef]
  7. Guo, D.; Chang, Z.; Tang, H.; Li, B.; Xu, X.; Yuan, X.Z.; Wang, H. Electrochemical performance of solid sphere spinel LiMn2O4 with high tap density synthesized by porous spherical Mn3O4. Electrochim. Acta 2014, 123, 254–259. [Google Scholar] [CrossRef]
  8. Feng, T.T.; Jian, Y.A.N.G.; Dai, S.Y.; Wang, J.C.; Wu, M.Q. Microemulsion synthesis of ZnMn2O4/Mn3O4 sub-microrods for Li-ion batteries and their conversion reaction mechanism. Trans. Nonferrous Met. Soc. China 2021, 31, 265–276. [Google Scholar] [CrossRef]
  9. Cai, C.; Koenig, G.M. Investigating dopants to improve sintered LiMn2O4 spinel electrode electrochemical cycling limitations. Electrochim. Acta 2022, 401, 139484. [Google Scholar] [CrossRef]
  10. Ren, Y.; Zhang, J.; Yang, Y.; Liu, F. Polyvinyl Butyral/SiO2 Nanoparticles Composite Coating on Poly(vinylidene fluoride) Separators for Lithium-Ion Batteries. J. Macromol. Sci. Part B 2022, 61, 194–205. [Google Scholar] [CrossRef]
  11. Georgios, N.; Mérièm, A. “Less is More”: Ultra Low LiPF6 Concentrated Electrolyte for Efficient Li-Ion Batteries. Batter. Supercaps 2021, 4, 1708–1719. [Google Scholar]
  12. Cholant, C.M.; Rodrigues, M.P.; Balboni, R.D.; Krüger, L.U.; Lemos, R.M.; Lopes, D.F.; Pawlicka, A.; Avellaneda, C.O. Study of the effect of LiClO4 concentration on the ionic transport of solid polymer electrolyte based on poly(vinyl alcohol)/gum Arabic. Ionics 2022, 28, 2715–2729. [Google Scholar] [CrossRef]
  13. Ouyang, C.Y.; Shi, S.Q.; Lei, M.S. Jahn-Teller distortion and electronic structure of LiMn2O4. J. Alloys Compd. 2009, 474, 370–374. [Google Scholar] [CrossRef]
  14. Yao, L.; Xi, Y.; Han, H.; Li, W.; Wang, C.; Feng, Y. LiMn2O4 prepared from waste lithium ion batteries through sol-gel process. J. Alloys Compd. 2021, 868, 159222. [Google Scholar] [CrossRef]
  15. Yin, T.; Lin, X.; Han, T.; Zhou, T.; Li, J.; Liu, J. A novel coral-like LiMn2O4 nanostructure as Li-ion battery cathode displaying stable energy-storage performance. J. Electroanal. Chem. 2023, 928, 117090. [Google Scholar] [CrossRef]
  16. Patil, A.; Patil, V.; Shin, D.W.; Choi, J.-W.; Paik, D.-S.; Yoon, S.-J. Issue and challenges facing rechargeable thin film lithium batteries. Mater. Res. Bull. 2007, 43, 1913–1942. [Google Scholar] [CrossRef]
  17. Chen, Z.L.; Gu, Y.J.; Huo, Y.L.; Ma, X.Y.; Wu, F.Z. Enhanced electrochemical performance of manganese-based metal organic frameworks-derived spinel LiMn2O4 cathode materials by improving the Mn3+ content and oxygen vacancies. J. Alloys Compd. 2022, 917, 165485. [Google Scholar] [CrossRef]
  18. Cao, J.H.; Guo, S.W.; Yan, R.Y. Carbon-coated single-crystalline LiMn2O4 nanowires synthesized by high-temperature solid-state reaction with high capacity for Li-ion battery. J. Alloys Compd. 2018, 741, 1–6. [Google Scholar] [CrossRef]
  19. Kumar, N.; Rodriguez, J.R.; Pol, V.G.; Sen, A. Facile synthesis of 2D graphene oxide sheet enveloping ultrafine 1D LiMn2O4 as interconnected framework to enhance cathodic property for Li-ion battery. Appl. Surf. Sci. 2018, 463, 132–140. [Google Scholar] [CrossRef]
  20. Nakajima, K.; Souza, F.L.; Freitas, A.L.; Thron, A.; Castro, R.H. Improving Thermodynamic Stability of nano-LiMn2O4 for Li-Ion Battery Cathode. Chem. Mater. 2021, 33, 3915–3925. [Google Scholar] [CrossRef]
  21. Yuan, D.; Su, M.Y.; Liu, Q.B. Effects of AgNPs-coating on the electrochemical performance of LiMn2O4 cathode material for lithium-ion batteries. J. Solid State Electrochem. 2022, 26, 2469–2478. [Google Scholar] [CrossRef]
  22. Guo, H.; Li, M.; Li, F.; Zhu, Q.; Zhao, Y.; Wang, F.; Qin, Z. Enhanced Wettability of a PTFE Porous Membrane for a High-Temperature Stable Lithium-Ion Battery Separator. Chem. Eng. Technol. 2022, 45, 737–744. [Google Scholar] [CrossRef]
  23. Liu, Y.; Yao, J.; Jiang, J.; Li, Y.; Zhu, Q. Exploring the effect of different additives on the preparation of α-Mn2O3/Mn3O4 composites and their zinc ion storage performances. Ionics 2023, 29, 1469–1478. [Google Scholar] [CrossRef]
  24. Lett, J.A.; Alshahateet, S.F.; Fatimah, I.; Sivasankaran, R.P.; Sibhatu, A.K.; Le, M.-V.; Sagadevan, S. Hydrothermal Synthesis and Photocatalytic Activity of Mn3O4 Nanoparticles. Top. Catal. 2023, 66, 126–138. [Google Scholar] [CrossRef]
  25. Zhu, M.; Fu, R. Synthesis and Analysis of Mn3O4@Graphene Nanocomposites for Supercapacitors. Adv. Compos. Lett. 2017, 26, 313–318. [Google Scholar] [CrossRef]
  26. Du, K.; Hu, G.R.; Peng, Z.D.; Qi, L. Synthesis of spinel LiMn2O4 with manganese carbonate prepared by micro-emulsion method. Electrochim. Acta 2010, 55, 1733–1739. [Google Scholar] [CrossRef]
  27. Gabrisch, H.; Wilcox, J.; Doeff, M.M. TEM Study of Fracturing in Spherical and Plate-like LiFePO4 Particles. Electrochem. Solid-State Lett. 2007, 11, A25–A29. [Google Scholar] [CrossRef]
  28. Zhu, H.L.; Chen, Z.Y.; Ji, S.; Linkov, V. Influence of different morphologies on electrochemical performance of spinel LiMn2O4. Solid State Ion. 2008, 179, 1788–1793. [Google Scholar] [CrossRef]
  29. Li, L.; Liang, J.; Kang, H.; Fang, J.; Luo, M.; Jin, X. TEA-assisted synthesis of single-crystalline Mn3O4 octahedrons and their magnetic properties. Appl. Surf. Sci. 2012, 261, 717–721. [Google Scholar] [CrossRef]
  30. Park, H.W.; Lee, H.J.; Park, S.-M.; Roh, K.C. Synthesis and electrochemical properties of Mn3O4 nanocrystals with controlled morphologies grown from compact ion layers. RSC Adv. 2016, 18, 191–197. [Google Scholar]
  31. Wang, W.Z.; Xu, C.K.; Wang, G.H.; Liu, Y.K.; Zheng, C.L. Preparation of Smooth Single-Crystal Mn3O4 Nanowires. Adv. Mater. 2002, 14, 837–840. [Google Scholar] [CrossRef]
  32. Wang, W.; Yang, T.; Yan, G.; Li, H. Synthesis of Mn3O4 hollow octahedrons and their possible growth mechanism. Mater. Lett. 2012, 82, 237–239. [Google Scholar] [CrossRef]
  33. Cao, S.; Han, T.; Peng, L.; Liu, B. Hydrothermal preparation, formation mechanism and gas-sensing properties of novel Mn3O4 nano-octahedrons. Mater. Lett. 2019, 246, 210–213. [Google Scholar] [CrossRef]
  34. Ramulu, B.; Nagaraju, G.; Sekhar, S.C.; Yu, J.S. Waste tissue papers templated highly porous Mn3O4 hollow microtubes prepared via biomorphic method for pseudocapacitor applications. J. Alloys Compd. 2019, 772, 925–935. [Google Scholar] [CrossRef]
  35. Mansournia, M.; Azizi, F.; Rakhshan, N. A novel ammonia-assisted method for the direct synthesis of Mn3O4 nanoparticles at room temperature and their catalytic activity during the rapid degradation of azo dyes. J. Phys. Chem. Solids 2015, 80, 91–97. [Google Scholar] [CrossRef]
  36. Yi, L.X.; Hu, G.X.; Huang, H. Morphology evolution of exfoliated trimanganese tetroxide nanosheets and mass transfer model of growth kinetics in supercritical N,N-dimethylformamide. Powder Technol. 2014, 259, 109–116. [Google Scholar] [CrossRef]
  37. Wang, M.; Huang, Y.; Zhang, N.; Wang, K.; Chen, X.; Ding, X. A facile synthesis of controlled Mn3O4 hollow polyhedron for high-performance lithium-ion battery anodes. Chem. Eng. J. 2018, 334, 2383–2391. [Google Scholar] [CrossRef]
  38. Zhang, H.; Liang, C.; Tian, Z.; Wang, G.; Cai, W. Single Phase Mn3O4 Nanoparticles Obtained by Pulsed Laser Ablation in Liquid and Their Application in Rapid Removal of Trace Pentachlorophenol. J. Phys. Chem. C 2010, 114, 12524–12528. [Google Scholar] [CrossRef]
  39. Chen, P.; Zheng, G.; Li, S.; Wang, Z.; Guo, G.; Tang, J.; Wen, Z.; Ji, S.; Cui, J.; Sun, J. Hydrothermal synthesis of core-shell Mn3O4@C microspheres as superior anode materials for high-performance lithium-ion storage. Solid State Ion. 2019, 338, 121–126. [Google Scholar] [CrossRef]
  40. Apte, S.K.; Naik, S.D.; Sonawane, R.S.; Kale, B.B.; Pavaskar, N.; Mandale, A.B.; Das, B.K. Nanosize Mn3O4 (Hausmannite) by microwave irradiation method. Mater. Res. Bull. 2006, 41, 647–654. [Google Scholar] [CrossRef]
  41. Ozkaya, T.; Baykal, A.; Kavas, H.; Köseoğlu, Y.; Toprak, M. A novel synthetic route to Mn3O4 nanoparticles and their magnetic evaluation. Phys. B 2008, 403, 3760–3764. [Google Scholar] [CrossRef]
  42. He, K.; Chen, N.; Wang, C.; Wei, L.; Chen, J. Method for Determining Crystal Grain Size by X-ray Diffraction. Cryst. Res. Technol. 2018, 53, 1700157. [Google Scholar] [CrossRef]
  43. Allen, G.C.; Harris, S.J.; Jutson, J.A.; Dyke, J.M. A study of a number of mixed transition metal oxide spinels using X-ray photoelectron spectroscopy. Appl. Surf. Sci. 1989, 37, 111–134. [Google Scholar] [CrossRef]
  44. Castro, V.D.; Polzonetti, G. XPS study of MnO oxidation. J. Electron Spectrosc. Relat. Phenom. 1989, 48, 117–123. [Google Scholar] [CrossRef]
Figure 1. E-pH diagram of Mn-NH3-SO42−-H2O system at 25 °C.
Figure 1. E-pH diagram of Mn-NH3-SO42−-H2O system at 25 °C.
Sustainability 15 13858 g001
Figure 2. Effect of pH regulators on the morphology of Mn3O4, (a) NaOH, (b) NH3∙H2O.
Figure 2. Effect of pH regulators on the morphology of Mn3O4, (a) NaOH, (b) NH3∙H2O.
Sustainability 15 13858 g002
Figure 3. Effect of reaction pH on the morphology of Mn3O4. (a) pH = 7, (b) pH = 7.5, (c) pH = 8, (d) pH = 8.5, (e) pH = 9.
Figure 3. Effect of reaction pH on the morphology of Mn3O4. (a) pH = 7, (b) pH = 7.5, (c) pH = 8, (d) pH = 8.5, (e) pH = 9.
Sustainability 15 13858 g003
Figure 4. Effect of reaction pH on physicochemical parameters of Mn3O4. (a) Mn content; (b) D50 and specific surface area.
Figure 4. Effect of reaction pH on physicochemical parameters of Mn3O4. (a) Mn content; (b) D50 and specific surface area.
Sustainability 15 13858 g004
Figure 5. Effect of reaction temperature on the morphology of Mn3O4. (a) 25 °C, (b) 50 °C, (c) 60 °C, (d) 70 °C, (e) 80 °C.
Figure 5. Effect of reaction temperature on the morphology of Mn3O4. (a) 25 °C, (b) 50 °C, (c) 60 °C, (d) 70 °C, (e) 80 °C.
Sustainability 15 13858 g005
Figure 6. Effect of reaction temperature on the physicochemical parameters of Mn3O4. (a) Mn content; (b) D50 and specific surface area.
Figure 6. Effect of reaction temperature on the physicochemical parameters of Mn3O4. (a) Mn content; (b) D50 and specific surface area.
Sustainability 15 13858 g006
Figure 7. XRD pattern of Mn3O4.
Figure 7. XRD pattern of Mn3O4.
Sustainability 15 13858 g007
Figure 8. XPS spectra of Mn3O4. (a) Full spectrum; (b) Fitting curves of Mn 2p, (c) Fitting curves of Mn 3s.
Figure 8. XPS spectra of Mn3O4. (a) Full spectrum; (b) Fitting curves of Mn 2p, (c) Fitting curves of Mn 3s.
Sustainability 15 13858 g008
Figure 9. The particle size distribution curve of Mn3O4.
Figure 9. The particle size distribution curve of Mn3O4.
Sustainability 15 13858 g009
Figure 10. The XRD patterns of LiMn2O4 under different roasting times.
Figure 10. The XRD patterns of LiMn2O4 under different roasting times.
Sustainability 15 13858 g010
Figure 11. The SEM images of LiMn2O4 prepared at different roasting times of (a) 4 h, (b) 6 h, (c) 8 h, (d) 10 h, (e) 12 h.
Figure 11. The SEM images of LiMn2O4 prepared at different roasting times of (a) 4 h, (b) 6 h, (c) 8 h, (d) 10 h, (e) 12 h.
Sustainability 15 13858 g011
Figure 12. Electrochemical performance curves of LiMn2O4 under different roasting times. (a) Initial charge–discharge curves, (b) Cycle performance curves, (c) Magnification performance curves.
Figure 12. Electrochemical performance curves of LiMn2O4 under different roasting times. (a) Initial charge–discharge curves, (b) Cycle performance curves, (c) Magnification performance curves.
Sustainability 15 13858 g012
Figure 13. The XRD patterns of LiMn2O4 under different roasting temperatures.
Figure 13. The XRD patterns of LiMn2O4 under different roasting temperatures.
Sustainability 15 13858 g013
Figure 14. The SEM images of LiMn2O4 under different roasting temperatures of (a) 500 °C, (b) 600 °C, (c) 700 °C, (d) 800 °C, (e) 850 °C.
Figure 14. The SEM images of LiMn2O4 under different roasting temperatures of (a) 500 °C, (b) 600 °C, (c) 700 °C, (d) 800 °C, (e) 850 °C.
Sustainability 15 13858 g014
Figure 15. Electrochemical performance curves of LiMn2O4 under different temperatures. (a) Initial charge–discharge curve, (b) Cycle performance curve, (c) Magnification performance curve.
Figure 15. Electrochemical performance curves of LiMn2O4 under different temperatures. (a) Initial charge–discharge curve, (b) Cycle performance curve, (c) Magnification performance curve.
Sustainability 15 13858 g015
Table 1. Physicochemical parameters of Mn3O4 prepared with different pH regulators.
Table 1. Physicochemical parameters of Mn3O4 prepared with different pH regulators.
pH RegulatorMn Content (%)D50 (μm)Specific Surface Area (m2∙g−1)
NaOH71.083.6822.18
NH3∙H2O71.324.673.49
Table 2. Results of chemical element analysis.
Table 2. Results of chemical element analysis.
ElementContent (%)ElementElement (%)
Mn71.45Cu0.0001
K0.0012Zn0.016
Na0.0011Si0.019
Ca0.015Fe0.0031
Mg0.013S0.031
Co0.0061Cl0.027
Ni0.0059H2O0.3
Table 3. Lattice parameters of LiMn2O4 prepared at different roasting times.
Table 3. Lattice parameters of LiMn2O4 prepared at different roasting times.
Time/hA (Å)V (Å3)R = I311/I400
48.236558.71.020
68.234558.31.016
88.233558.11.014
108.231557.61.013
128.234558.31.021
Table 4. Lattice parameters of LiMn2O4 under different roasting temperatures.
Table 4. Lattice parameters of LiMn2O4 under different roasting temperatures.
Temperature (°C)A (Å)V (Å3)R = I311/I400
5008.246560.71.051
6008.222555.81.010
7008.229557.21.011
8008.231557.61.013
8508.237558.91.023
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

Wang, H.; Wang, J.; Wang, H.; Wang, S.; Dong, X.; Hao, W.; Lu, J.; Lu, F. Study and Property Characterization of LiMn2O4 Synthesized from Octahedral Mn3O4. Sustainability 2023, 15, 13858. https://doi.org/10.3390/su151813858

AMA Style

Wang H, Wang J, Wang H, Wang S, Dong X, Hao W, Lu J, Lu F. Study and Property Characterization of LiMn2O4 Synthesized from Octahedral Mn3O4. Sustainability. 2023; 15(18):13858. https://doi.org/10.3390/su151813858

Chicago/Turabian Style

Wang, Hao, Jiawei Wang, Haifeng Wang, Song Wang, Xinyu Dong, Wenhao Hao, Ju Lu, and Fanghai Lu. 2023. "Study and Property Characterization of LiMn2O4 Synthesized from Octahedral Mn3O4" Sustainability 15, no. 18: 13858. https://doi.org/10.3390/su151813858

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