Next Article in Journal
Analysis of Wear Resistance of Borided Steel C45
Previous Article in Journal
Achieving Ti-5Al-4Sn-2Zr-1Mo-0.25Si-1Nb Alloys with High Strength and Moderate Ductility through Selective Laser Melting
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of PVP Coating on LiMnBO3 Cathodes for Li-Ion Batteries

1
Hubei Key Laboratory of Mine Environmental Pollution Control & Remediation, School of Materials Science and Engineering, Hubei Polytechnic University, Huangshi 435003, China
2
Institute of Nuclear and New Energy Technology, Tsinghua University, Beijing 100084, China
*
Author to whom correspondence should be addressed.
Materials 2020, 13(23), 5528; https://doi.org/10.3390/ma13235528
Submission received: 28 October 2020 / Revised: 29 November 2020 / Accepted: 30 November 2020 / Published: 3 December 2020

Abstract

:
LiMnBO3 is a potential cathode for Li-ion batteries, but it suffers from a low electrochemical activity. To improve the electrochemical performance of LiMnBO3, the effect of polyvinyl pyrrolidone (PVP) as carbon additive was studied. Monoclinic LiMnBO3/C and LiMnBO3-MnO/C materials were obtained by a solid-state method at 500 °C. The structure, morphology and electrochemical behavior of these materials are characterized and compared. The results show that carbon additives and ball-milling dispersants affect the formation of impurities in the final products, but MnO is beneficial for the performance of LiMnBO3. The sample of LiMnBO3-MnO/C delivered a high capacity of 162.1 mAh g−1 because the synergistic effect of the MnO/C composite and the suppression of the PVP coating on particle growth facilitates charge transfer and lithium–ion diffusion.

1. Introduction

Over the last two decades, polyanionic compounds, such as phosphates, sulfates, silicates and borates, have been investigated extensively as a new class of cathodes for lithium–ion batteries due to their high safety depending on their stable three-dimensional framework [1,2,3]. Among these compounds, LiMnBO3 has attracted much attention owing to its high theoretical capacity and low cost [1,4]. LiMnBO3 exists in two polymorphs, the monoclinic phase and the hexagonal phase. Monoclinic LiMnBO3 (m-LiMnBO3) occurs at low temperatures, which transforms into hexagonal LiMnBO3 (h-LiMnBO3) at high calcination temperatures [5,6,7,8]. The two polymorphs have the same theoretical capacity of 222 mAh g−1, while h-LiMnBO3 has a higher redox voltage and m-LiMnBO3 has better electrochemical activity [5]. More studies focus on the improvement of m-LiMnBO3 [9,10,11,12].
Monoclinic LiMnBO3 polymorph was first synthesized by Bondareva in 1978 [13]. It was discovered to have a reversible capacity in 2001, but only 2% of its theoretical capacity could be delivered, which is attributed to large polarization [14]. To improve the performance of LiMnBO3, several approaches have been proposed, including carbon coating [15,16,17,18,19], particle size reduction [20,21,22,23], and cation doping or substitution [9,24,25,26]. Although the performance of LiMnBO3 can be enhanced by doping and fine particles, it needs to be coated with carbon at the same time [15,16,17,18,19]. Carbon coating has played a key role in improving electrode materials. It could not only enhance the electronic and ionic conductivity of materials, but could also inhibit crystal growth [15,16,17,18,19].
Carbon coating is a surface modification. The carbon content directly affects the performance of LiMnBO3. To obtain high-rate performance, more carbon is required. However, a large amount of carbon will cause the energy density of electrode materials to decrease. To form highly conductive substances in situ on the surface of the materials is an alternative approach [27]. In addition, the effect of carbon coating depends on carbon sources and synthesis methods. Polyvinyl pyrrolidone (PVP) has outstanding wetting properties and easily forms films, making it useful as a coating additive. The pyrolysis of PVP under inert atmosphere can generate a uniform carbon coating layer on the surface of active materials, thereby improving the electrochemical performance of the electrode materials [28,29]. In this work, LiMnBO3/C and LiMnBO3-MnO/C composites were synthesized by a simple solid-state method using PVP as the additive. The effect of PVP on the structure, morphology and electrochemical behavior of LiMnBO3 was studied.

2. Materials and Methods

2.1. Synthesis of Materials

Carbon-coated LiMnBO3 materials were prepared by a solid-state reaction. Stoichiometric amounts of Li2CO3 (AR, Sinopharm Chemical Reagent Co. Ltd, Shanghai, China), MnCO3 (AR, Aladdin, Shanghai, China), and H3BO3 (AR, Aladdin, Shanghai, China) were added into a stainless steel container, in which 20 wt.% of PVP (K30, GR, Sinopharm Chemical Reagent Co. Ltd, Shanghai, China) from the total weight of the raw materials was mixed with 50 mL alcohol. The PVP-containing precursor was ground by ball-milling at 400 rpm for 7 h, and then was dried in an oven at 80 °C. Subsequently, the mixture was heated at 500 °C for 10 h under an argon atmosphere. After cooling in a furnace, the final black powders (marked as LMB-PVP) were obtained. The other samples in which ethanol was replaced by water or acetone, and PVP was replaced by sucrose, starch or oxalic acid, were prepared by the same route.

2.2. Characterization

X-ray powder diffraction (XRD, Ultima IV, Rigaku, Tokyo, Japan) with Cu Kα radiation was used to identify the phases of the samples. The morphology was observed with field-emission scanning electron microscope (FESEM, JSM-7610F Plus, JEOL, Tokyo, Japan) equipped with energy-dispersive X-ray spectroscopy (EDS, Oxford INCA, High Wycombe, UK). The synthesis temperature and carbon contents were estimated by the TGA/DSC1 thermal analyzer of Mettler Toledo (Schwerzenbach, Switzerland).

2.3. Electrochemical Measurements

Electrochemical measurements were performed with 2025 coin-type cells. The cells were assembled using Li metal foil as the counter electrode with a separator (Celgard 2400, Celgard, LLC., Charlotte, NC, USA) and electrolyte containing 1 M LiPF6 in a solvent mixture of ethylene carbonate and diethyl carbonate (1:1 by volume). The working electrodes were made by mixing active material, conductive carbon black (Super P, Timcal, Brussels, Belgium) and polyvinylidene fluoride binder (HSV900, Arkema, Pierre-Bénite, France) in a weight ratio of 8:1:1. These cells were galvanostatically charged to 4.5 V at a rate of 0.05 C (1 C = 220 mA g−1) and then further charged using the constant voltage mode with a cutoff current density of 0.01 C. Next the cells were discharged to 2.0 V or 1.5 V with the same rate of 0.05 C. Electrochemical impedance spectroscopy (EIS) was measured in a frequency range of 0.1–100 kHz with an alternating current signal of 5 mV. All tests were carried out at room temperature.

3. Results and Discussion

In order to obtain highly active LiMnBO3, the reaction temperature was evaluated by thermo-gravimetric analysis (TGA) and differential scanning calorimetry (DSC). LMB-PVP precursor was heated from 25 °C to 900 °C at a rate of 5 °C min−1 under a nitrogen atmosphere. Two weight loss stages and four endothermic peaks are observed from the TG/DSC curves in Figure 1. The first weight loss between 25 °C and 200 °C is attributed to the release of physically adsorbed water, corresponding to an obvious endothermic peak at 68.1 °C. Sharp weight loss appears at 300~500 °C, corresponding to three endothermic peaks at 397.3, 419.4 and 435.3 °C. The peaks at 397.3 and 435.3 °C are related to the melting and decomposition of PVP, so the crystallization temperature of LiMnBO3 is about 420 °C. The weight loss of the LMB-PVP precursor mainly occurs below 500 °C. Thus, 500 °C was chosen as the synthesis temperature.
The crystal phase of the samples synthesized at 500 °C was identified from the XRD experiments. The XRD patterns of the samples that were synthesized using PVP as the carbon additive in different dispersants are shown in Figure 2. All the samples have a similar pattern, which can be indexed into the monoclinic structure with the C2/c space group. Except for the main phase of m-LiMnBO3, MnO impurity phase was found in the sample with ethanol as dispersant (LMB-PVP), and Mn3(BO3)2 impurity phase was found in the sample with water as dispersant (LMB-W). Single phase m-LiMnBO3 was obtained from the sample using acetone as dispersant (LMB-A). A packet around 2θ = 18–25° is observed from sample LMB-A and sample LMB-PVP, implying the presence of an amorphous phase composition. The residual carbon of PVP pyrolyzed in an argon atmosphere was estimated by TG to be 6.86 wt.%, while no carbon diffraction peaks are seen in all samples because the carbon is in an amorphous state.
The electrochemical behavior of the samples was measured using 2025 coin-type cells. Figure 3a shows the typical charge and discharge curves of LiMnBO3 samples at 0.05 C. No noticeable charge/discharge plateau is observed as in the previous report [30], indicating that the process of lithiation and delithiation is a solid solution reaction. The discharge capacities of samples LMB-A, LMB-W and LMB-PVP are 64.8, 80.1 and 110.2 mAh g−1, respectively. Sample LMB-PVP exhibits a better electrochemical performance due to the formation of the LiMnBO3-MnO/C composite. Although MnO is a semiconductor with low conductivity, the combination of MnO and carbon makes the MnO/C composite demonstrate excellent rate performance [31,32,33]. The MnO/C composite forms a good contact on the surface of LiMnBO3, thereby improving the electrochemical performance.
The electrochemical impedance spectra of samples LMB-A and LMB-PVP were measured in the discharged state after 10 full charge–discharge cycles. As shown in Figure 3b, the two samples have similar Nyquist plots composed of a high-to-medium frequency semicircle and a low-frequency straight-line. An electrochemical cell can be considered an equivalent circuit composed of the ohmic resistance (RΩ), the electric double-layer capacitance (Cd), the charge transfer resistance (Rct) and the Warburg impedance (Zw) [34]. At high frequencies, Zw becomes unimportant and the electrode process is controlled by kinetics. As such, the semicircle in the high frequency region is related to the charge transfer process. Sample LMB-PVP exhibits smaller charge transfer resistances. The inclined line in the low-frequency region of the plots is related to lithium–ion diffusion, where the electrode process is dominated by Warburg impedance. The lithium–ion diffusion coefficient (D) can be roughly calculated according to the formula of D = 0.5 R T / A F 2 C σ 2 , where R is the gas constant, T is the absolute temperature, A is the cathode electrode area, F is the Faraday constant, C is the concentration of lithium-ion in the electrode and σ is the Warburg factor. The value of σ can be obtained from the linear fitting of the real part (Zʹ) versus ω−1/2 in the low frequency region [35,36]. The diffusion coefficients of samples LMB-A and LMB-PVP at room temperature are 2.50 × 10−15 cm2 s−1 and 6.69 × 10−15 cm2 s−1, respectively. The above results demonstrate that the introduction of MnO can facilitate charge transfer and lithium–ion diffusion.
Different carbon-coated samples and a no-carbon sample were also synthesized by the same process. Figure 4 shows the XRD patterns of these samples. Strong diffraction peaks of m-LiMnBO3 are observed for all samples, except for the one using oxalic acid as the carbon additive (LMB-OA). No obvious diffraction peaks of h-LiMnBO3 are found in sample LMB-PVP, but they appear in the other four samples. MnO impurities are seen in all samples, implying that there are lithium and boron compounds in these samples due to stoichiometric raw materials. The sample without carbon coating (LMB-NoC) shows that a small amount of Li2CO3 did not react. Because of the strong reducibility of oxalic acid, MnO becomes the main phase of sample LMB-OA, leading to only a bit of LiMnBO3 in the final product. The carbon-coated sample with starch (LMB-ST) and the one with sucrose (LMB-SU) have very similar diffraction profiles. Although no other impurity peaks except MnO are detected from the XRD pattern of sample LMB-PVP, there may be some amorphous lithium and boron compounds owing to stoichiometric amounts of the raw materials.
The surface structure of LiMnBO3 samples was investigated by SEM. Figure 5 shows the SEM images of LiMnBO3 without carbon and with different carbon coatings. The sample without carbon presents an irregular spherical shape with the size ranging from 100 to 400 nm, and carbon-coated samples show smaller particle sizes of around 100 nm. Obviously, the addition of carbon suppressed particle growth of LiMnBO3. It is advantageous to improve the electrochemical performance of electrode materials. As can be seen in Figure 5c,d, the sample LMB-PVP has the smallest particle size, consisting of particles with a size of ~60 nm and uniform distribution. The starch sample has a particle size about 50–150 nm, which shows obvious agglomeration.
Figure 6a shows the second charge and discharge curves of LiMnBO3 samples at 0.05 C rate. The sample without carbon coating delivered a capacity less than 10 mAh g−1 with a large polarization. The addition of carbon reduced polarization and enhanced the capacity significantly. The discharge capacities of LMB-ST, LMB-SU and LMB-PVP are 64.8, 91 and 147.7 mAh g−1, respectively. The improved electrochemical behavior of carbon-coated samples is attributed to the increase in conductivity and the decrease in particle sizes by the added carbon. Sample LMB-PVP exhibits the best capacity among these samples due to it having less impurities and smaller particle sizes, which is consistent with the results of XRD and SEM.
The cyclic performance of LiMnBO3 samples at 0.05 C is shown in Figure 6b. It can be seen clearly that the sample without carbon coating reveals a low capacity, and this may be related to air exposure [7]. The initial discharge capacities of LMB-ST, LMB-SU and LMB-PVP are 71.2, 98.5 and 162.1 mAh g−1, respectively. Compared with the other carbon-coated materials, sample LMB-PVP with low carbon content exhibits a high discharge capacity, as shown in Table 1. The presence of MnO/C composite enhances the capacity of the LiMnBO3 material, and it will not significantly reduce the energy density of LiMnBO3 due to the high density of MnO (5.37 g cm−3) [37]. After 20 cycles, the capacity retention is only 60% for sample PVP. The irreversible cycling behavior is likely due to considerable Li extraction causing an unstable structure in LiMnBO3 [11]. Although the cycle performance of LiMnBO3 needs further improvement, the sample LMB-PVP still has the best electrochemical performance among all the samples.

4. Conclusions

Carbon-coated LiMnBO3 materials were synthesized by a simple solid-state method using different carbon additives in various dispersants. A LiMnBO3-MnO/C composite was obtained with PVP as the carbon source, which exhibits good electrochemical performance because the synergistic effect of the MnO/C composite and the suppression of the PVP coating on the particle growth facilitates charge transfer and lithium-ion diffusion. The formation of an MnO/C composite by the PVP coating can be used as an alternative surface modification approach to avoid a significant decrease in the energy density of LiMnBO3 coated by only carbon. The future research will focus on improving the cycle performance of LiMnBO3 by controlling the content of MnO.

Author Contributions

Conceptualization, B.H. and X.H.; methodology, C.H. and H.Y.; formal analysis, X.H.; investigation, B.H., C.H. and H.Y.; data curation, B.H. and C.H.; writing—original draft preparation, B.H. and C.H.; writing—review and editing, X.H. and H.Y.; supervision, C.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science Foundation of Hubei Province (No. 2015CFA146, No. 2013CFC100) and the Program for Innovation team in Hubei Polytechnic University (No. 13xtz09).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lee, W.; Kim, J.; Yun, S.; Choi, W.; Kim, H.; Yoon, W. Multiscale factors in designing alkali-ion (Li, Na, and K) transition metal inorganic compounds for next-generation rechargeable batteries. Energy Environ. Sci. 2020. [Google Scholar] [CrossRef]
  2. Masquelier, C.; Croguennec, L. Polyanionic (Phosphates, Silicates, Sulfates) Frameworks as Electrode Materials for Rechargeable Li (or Na) Batteries. Chem. Rev. 2013, 113, 6552–6591. [Google Scholar] [CrossRef] [PubMed]
  3. Canepa, P.; Sai Gautam, G.; Hannah, D.C.; Malik, R.; Liu, M.; Gallagher, K.G.; Persson, K.A.; Ceder, G. Odyssey of Multivalent Cathode Materials: Open Questions and Future Challenges. Chem. Rev. 2017, 117, 4287–4341. [Google Scholar] [CrossRef] [PubMed]
  4. Kalantarian, M.M.; Hafizi-Barjini, M.; Momeni, M. Ab Initio Study of AMBO3 (A = Li, Na and M = Mn, Fe, Co, Ni) as Cathode Materials for Li-Ion and Na-Ion Batteries. ACS Omega 2020, 5, 8952–8961. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Kim, J.C.; Moore, C.J.; Kang, B.; Hautier, G.; Jain, A.; Ceder, G. Synthesis and Electrochemical Properties of Monoclinic LiMnBO3 as a Li Intercalation Material. J. Electrochem. Soc. 2011, 158, A309. [Google Scholar] [CrossRef]
  6. Zhao, L.; Li, R.K. Study on a multifunctional crystal LiMnBO3. Mater. Res. Bull. 2013, 48, 277–280. [Google Scholar] [CrossRef]
  7. Yamada, A.; Iwane, N.; Nishimura, S.; Koyama, Y.; Tanaka, I. Synthesis and electrochemistry of monoclinic Li(MnxFe1-x)BO3: A combined experimental and computational study. J. Mater. Chem. 2011, 21, 10690–10696. [Google Scholar] [CrossRef]
  8. Lee, Y.; Lee, H. Improved lithium storage capacities of LiMnBO3/C via simple high-energy milling. Mater. Lett. 2014, 132, 401–404. [Google Scholar] [CrossRef]
  9. Kim, J.C.; Seo, D.; Ceder, G. Theoretical capacity achieved in a LiMn0.5Fe0.4Mg0.1BO3 cathode by using topological disorder. Energy Environ. Sci. 2015, 8, 1790–1798. [Google Scholar] [CrossRef]
  10. Ragupathi, V.; Safiq, M.; Panigrahi, P.; Hussain, T.; Raman, S.; Ahuja, R.; Nagarajan, G.S. Enhanced electrochemical performance of LiMnBO3 with conductive glassy phase: A prospective cathode material for lithium-ion battery. Ionics 2017, 23, 1645–1653. [Google Scholar] [CrossRef]
  11. Le Roux, B.; Bourbon, C.; Lebedev, O.I.; Colin, J.; Pralong, V. Synthesis and Characterization of the LiMnBO3–LiCoBO3 Solid Solution and Its Use as a Lithium-Ion Cathode Material. Inorg. Chem. 2015, 54, 5273–5279. [Google Scholar] [CrossRef] [PubMed]
  12. Kim, J.C.; Li, X.; Moore, C.J.; Bo, S.; Khalifah, P.G.; Grey, C.P.; Ceder, G. Analysis of Charged State Stability for Monoclinic LiMnBO3 Cathode. Chem. Mater. 2014, 26, 4200–4206. [Google Scholar] [CrossRef]
  13. Bondareva, O.S.; Simonov, M.A.; Egorov-Tismenko, Y.K.; Belov, N.V. The Crystal Structures of LiZn[BO3] and LiMn[BO3]. Sov. Phys. Cryst. 1978, 23, 269–271. [Google Scholar]
  14. Legagneur, V.; An, Y.; Mosbah, A.; Portal, R.; Le Gal La Salle, A.; Verbaere, A.; Guyomard, D.; Piffard, Y. LiMBO3 (M=Mn, Fe, Co): Synthesis, crystal structure and lithium deinsertion/insertion properties. Solid State Ion. 2001, 139, 37–46. [Google Scholar] [CrossRef]
  15. Aravindan, V.; Karthikeyan, K.; Amaresh, S.; Lee, Y.S. LiMnBO3/C: A Potential Cathode Material for Lithium Batteries. Bull. Korean Chem. Soc. 2010, 31, 1506–1508. [Google Scholar] [CrossRef] [Green Version]
  16. Tang, A.; He, D.; He, Z.; Xu, G.; Song, H.; Peng, R. Electrochemical performance of LiMnBO3/C composite synthesized by a combination of impregnation and precipitation followed by annealing. J. Power Sources 2015, 275, 888–892. [Google Scholar] [CrossRef]
  17. Chen, W.; Zhang, H.; Zhang, X.; Wu, L.; Liu, J.; Liu, S.; Zhong, S. Synthesis and electrochemical performance of carbon-coated LiMnBO3 as cathode materials for lithium-ion batteries. Ionics 2018, 24, 73–81. [Google Scholar] [CrossRef]
  18. Li, L.; Zheng, H.; Yin, S.; Wang, S.; Feng, C.; Wang, J.; He, P. Synthesis and Electrochemical Properties of LiMnBO3 and LiMnBO3/C Composite. Sci. Adv. Mater. 2016, 8, 980–986. [Google Scholar] [CrossRef]
  19. Zhang, B.; Zhu, Y.; Yu, W.; Zhang, J.; An, C. Facile synthesis of carbon-encapsulated LiMnBO3 composite by the sol-gel method as a lithium-ion battery cathode material. J. Alloy. Compd. 2017, 704, 343–347. [Google Scholar] [CrossRef]
  20. Moradi, M.; Kim, J.C.; Qi, J.; Xu, K.; Li, X.; Ceder, G.; Belcher, A.M. A bio-facilitated synthetic route for nano-structured complex electrode materials. Green Chem. 2016, 18, 2619–2624. [Google Scholar] [CrossRef] [Green Version]
  21. Ragupathi, V.; Srimathi, K.; Panigrahi, P.; Lee, J.W.; Nagarajan, G.S. Electrochemical Performance of Sol-Gel Derived Hexagonal LiMnBO3 Cathode Material for Lithium-Ion Batteries. Nano Hybrids Compos. 2017, 17, 106–112. [Google Scholar] [CrossRef]
  22. Michalski, P.P.; Gołębiewska, A.; Trébosc, J.; Lafon, O.; Pietrzak, T.K.; Ryl, J.; Nowiński, J.L.; Wasiucionek, M.; Garbarczyk, J.E. Properties of LiMnBO3 glasses and nanostructured glass-ceramics. Solid State Ion. 2019, 334, 88–94. [Google Scholar] [CrossRef] [Green Version]
  23. Afyon, S.; Kundu, D.; Krumeich, F.; Nesper, R. Nano LiMnBO3, a high-capacity cathode material for Li-ion batteries. J. Power Sources 2013, 224, 145–151. [Google Scholar] [CrossRef]
  24. Yamane, H.; Kawano, T.; Fukuda, K.; Suehiro, T.; Sato, T. Preparation, crystal structure and photoluminescence of lithium magnesium manganese borate solid solutions, LiMg1xMnxBO3. J. Alloy. Compd. 2012, 512, 223–229. [Google Scholar] [CrossRef]
  25. Stafeeva, V.S.; Panin, R.V.; Lobanov, M.V.; Antipov, E.V. Stabilization of the LiMnBO3 monoclinic polymorph by the isovalent substitution of manganese for zinc. Russ. Chem. 2013, 62, 374–379. [Google Scholar] [CrossRef]
  26. Jarocka, A.; Michalski, P.P.; Ryl, J.; Wasiucionek, M.; Garbarczyk, J.E.; Pietrzak, T.K. Synthesis, thermal, structural and electrical properties of vanadium-doped lithium-manganese-borate glass and nanocomposites. Ionics 2020, 26, 1275–1283. [Google Scholar] [CrossRef] [Green Version]
  27. Chung, S.Y.; Bloking, J.T.; Chiang, Y.M. Electronically conductive phospho-olivines as lithium storage electrodes. Nat. Mater. 2002, 1, 123–128. [Google Scholar] [CrossRef]
  28. Gan, Q.; Qin, N.; Zhu, Y.; Huang, Z.; Zhang, F.; Gu, S.; Xie, J.; Zhang, K.; Lu, L.; Lu, Z. Polyvinylpyrrolidone-Induced Uniform Surface-Conductive Polymer Coating Endows Ni-Rich LiNi0.8Co0.1Mn0.1O2 with Enhanced Cyclability for Lithium-Ion Batteries. ACS Appl. Mater. Interfaces 2019, 11, 12594–12604. [Google Scholar] [CrossRef]
  29. Lee, J.; Kim, M.; Moon, S.; Kim, E.; Shin, Y.; Choi, S.; Kwon, S.; Kim, S.; Kwon, H.; Park, K. Role of polyvinylpyrrolidone in the electrochemical performance of Li2MnO3 cathode for lithium-ion batteries. RSC Adv. 2019, 9, 10297–10304. [Google Scholar] [CrossRef] [Green Version]
  30. Lee, Y.; Lee, H. Structure and electrochemical behavior of LiMnBO3 synthesized at various temperatures. Electron. Mater. Lett. 2014, 10, 253–258. [Google Scholar] [CrossRef]
  31. Bhide, V.G.; Dani, R.H. Electrical conductivity in oxides of manganese and related compounds. Physica 1961, 27, 821–826. [Google Scholar] [CrossRef]
  32. Zhan, D.; Luo, W.; Kraatz, H.; Fehse, M.; Li, Y.; Xiao, Z.; Brougham, D.F.; Simpson, A.J.; Wu, B. Facile Approach for Synthesizing High-Performance MnO/C Electrodes from Rice Husk. ACS Omega 2019, 4, 18908–18917. [Google Scholar] [CrossRef] [Green Version]
  33. Hou, C.; Tai, Z.; Zhao, L.; Zhai, Y.; Hou, Y.; Fan, Y.; Dang, F.; Wang, J.; Liu, H. High performance MnO@C microcages with a hierarchical structure and tunable carbon shell for efficient and durable lithium storage. J. Mater. Chem. A 2018, 6, 9723–9736. [Google Scholar] [CrossRef] [Green Version]
  34. Bard, A.J.; Faulkner, L.R. Electrochemical Methods Fundamentals and Applications, 2nd ed.; John Wiley & Sons, Inc.: New York, NY, USA, 2001; pp. 376–387. [Google Scholar]
  35. Dinh, H.; Mho, S.; Kang, Y.; Yeo, I. Large discharge capacities at high current rates for carbon-coated LiMnPO4 nanocrystalline cathodes. J. Power Sources 2013, 244, 189–195. [Google Scholar] [CrossRef]
  36. Li, S.; Xu, L.; Li, G.; Wang, M.; Zhai, Y. In-situ controllable synthesis and performance investigation of carbon-coated monoclinic and hexagonal LiMnBO3 composites as cathode materials in lithium-ion batteries. J. Power Sources 2013, 236, 54–60. [Google Scholar] [CrossRef]
  37. Dean, J.A. Lange’s Handbook of Chemistry, 15th ed.; McGraw-Hill: New York, NY, USA, 1999; pp. 3–38. [Google Scholar]
Figure 1. TG/DSC curves of LMB-PVP precursor and PVP measured in N2.
Figure 1. TG/DSC curves of LMB-PVP precursor and PVP measured in N2.
Materials 13 05528 g001
Figure 2. XRD patterns of LiMnBO3 prepared by ball-milling the raw materials with PVP in different dispersants.
Figure 2. XRD patterns of LiMnBO3 prepared by ball-milling the raw materials with PVP in different dispersants.
Materials 13 05528 g002
Figure 3. (a) Typical charge/discharge curves and (b) Nyquist plots (inset: the linear fitting of Zʹ versus ω−1/2 relationship) of LiMnBO3 prepared by ball-milling the raw materials with PVP in different dispersants.
Figure 3. (a) Typical charge/discharge curves and (b) Nyquist plots (inset: the linear fitting of Zʹ versus ω−1/2 relationship) of LiMnBO3 prepared by ball-milling the raw materials with PVP in different dispersants.
Materials 13 05528 g003
Figure 4. XRD patterns of LiMnBO3 prepared by ball-milling the raw materials with various carbon additives and without carbon in ethanol.
Figure 4. XRD patterns of LiMnBO3 prepared by ball-milling the raw materials with various carbon additives and without carbon in ethanol.
Materials 13 05528 g004
Figure 5. SEM images of LiMnBO3 prepared by ball-milling the raw materials with (a) no carbon, (b) PVP, (c) sucrose and (d) starch in ethanol.
Figure 5. SEM images of LiMnBO3 prepared by ball-milling the raw materials with (a) no carbon, (b) PVP, (c) sucrose and (d) starch in ethanol.
Materials 13 05528 g005
Figure 6. (a) Charge/discharge curves and (b) cycle performance of LiMnBO3 prepared by ball-milling the raw materials with various carbon additives in ethanol.
Figure 6. (a) Charge/discharge curves and (b) cycle performance of LiMnBO3 prepared by ball-milling the raw materials with various carbon additives in ethanol.
Materials 13 05528 g006
Table 1. Comparison of the discharge capacities of LiMnBO3/C at 0.05 C.
Table 1. Comparison of the discharge capacities of LiMnBO3/C at 0.05 C.
SampleCarbon ResourceResidual Carbon (wt.%)Discharge Capacity (mAh g−1)
This workPVP6.9162
Ref. [7]Ketjen black10~70
Ref. [12]sucrose<10102
Ref. [16]Ketjen black22.6150
Ref. [30]Ketjen black~10~170
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hong, B.; He, X.; Yi, H.; Hu, C. Effect of PVP Coating on LiMnBO3 Cathodes for Li-Ion Batteries. Materials 2020, 13, 5528. https://doi.org/10.3390/ma13235528

AMA Style

Hong B, He X, Yi H, Hu C. Effect of PVP Coating on LiMnBO3 Cathodes for Li-Ion Batteries. Materials. 2020; 13(23):5528. https://doi.org/10.3390/ma13235528

Chicago/Turabian Style

Hong, Bolong, Xiangming He, Huihua Yi, and Chenglin Hu. 2020. "Effect of PVP Coating on LiMnBO3 Cathodes for Li-Ion Batteries" Materials 13, no. 23: 5528. https://doi.org/10.3390/ma13235528

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