Skip to content
BY 4.0 license Open Access Published by De Gruyter December 31, 2020

High-efficiency method for recycling lithium from spent LiFePO4 cathode

  • Tingting Yan , Shengwen Zhong EMAIL logo , Miaomiao Zhou , Xiaoming Guo , Jingwei Hu , Fangfang Wang , Fantao Zeng and Sicheng Zuo
From the journal Nanotechnology Reviews

Abstract

The extraction of Li from the spent LiFePO4 cathode is enhanced by the selective removal using interactions between HCl and NaClO to dissolve the Li+ ion while Fe and P are retained in the structure. Several parameters, including the effects of dosage and drop acceleration of HCl and NaClO, reaction time, reaction temperature, and solid–liquid ratio on lithium leaching, were tested. The Total yields of lithium can achieve 97% after extraction process that lithium is extracted from the precipitated mother liquor, using an appropriate extraction agent that is a mixture of P507 and TBP and NF. The method also significantly reduced the use of acid and alkali, and the economic benefit of recycling is improved. Changes in composition, morphology, and structure of the material in the dissolution process are characterized by inductively coupled plasma optical emission spectrometry, scanning electron microscope, X-ray diffraction, particle size distribution instrument, and moisture analysis.

1 Introduction

LiFePO4 is one of the main cathode materials in rechargeable lithium-ion batteries in China [1]. However, one of the values of LiFePO4, compared with other cathode materials such as Li2CoO2 [2], is that it does not contain high recovery value elements such as Mn, Ni, and Co [3]. Given this, China will have 9,400 tons of scrap LiFePO4 batteries by 2021 [4]. The lithium content accounts for about 200 tons, and projections of possible lithium scarcity point to clear environmental and economic benefits from lithium recycling in the future [5].

Currently, there are several methods for treating waste cathode materials [6], including the fire method, the wet method, and a combined method [7]. Many companies used high-temperature method to recycle lithium battery cathode material, such as Toxco Inc. (USA), SONY Corp (Japan), and Umicore (Belgium). These methods are rather inefficient for recovery as mostly lithium is lost in metallurgical slag during refinement and is not recycled [8,9]. These high-temperature processes have high energy consumption and produce waste gas and waste residue pollution and have pushed methods toward wet extraction of lithium in the industry and academic research [10].

Through a direct regeneration process, Li et al. [11] obtained high purity cathode material mixture (LiFePO4 þ acetylene black), anode material mixture (graphite þ acetylene black), and other by-products (shell, Al foil, Cu foil, electrolyte solvent, etc.) from scrapped LiFePO4 batteries. Subsequently, recycled cathode material mixture without acid leaching is further regenerated with Li2CO3 directly. Batteries using recycled LiFePO4 confronted many problems, such as safety and cycling stability. Based on the concept of zero waste, Dutta et al. [12] proposed a complete process for the recycling of LIBs, in which the metals and materials as value added products were obtained. Although the leaching rate of Li+ ions were 99.9%, the recovery rate of lithium indeed was not high, because the consumption of oxidant H2O2 is relatively high, except that the battery grade Li2CO3 cannot be obtained. Zhang et al. [13] studied the selective recovery of lithium from spent LiFePO4 batteries. With the help of sodium persulfate (Na2S2O8), the LiFePO4 were oxidized to FePO4, and more than 99% of Li in the material can be selectively leached in 20 min at room temperature. But the sodium persulfate (Na2S2O8) that has the chemical properties of highly oxidizable and toxic is expensive, and this method is difficult to realize industrial production.

Yang et al. [14] also proposed a method to extract lithium from spent LiFePO4 cathode; the wet extraction of Li from LiFePO4 cathode typically involves H2SO4 and H2O2 to dissolve the cathode material, adjusting pH to precipitate Fe3+, filtering the precipitate, and precipitating Li with Na2CO3 solution. This approach is marred by relatively low lithium recovery, excessive use of acid and alkali, and high energy consumption with relatively poor Li2CO3 purity. Economically, the market value is limited without a more efficient method for the recovery of lithium [15].

The method shown herein is based on the wet process [16]. The refined approach studies the structural change of LiFePO4 by hydrometallurgy to achieve preferred dissolution of lithium, thereby minimizing the reaction volumes and retaining the solid structure of FePO4 [17].

Careful control of HCl and NaClO additions establishes a chemical equilibrium, and lithium is efficiently extracted from the structure with a higher yield [18]. This dramatically reduces the consumption of acid and alkali [19,20]. Recycling lithium of the mother liquor by extraction technology reduces energy consumption [21]. The unique reverse sedimentation process prepares battery-grade Li2CO3 for the increased economic benefit [22,23]. Keeping in view of stringent environmental regulations, limited natural resources, and energy crisis, adopting recycling will not only protect the environment and pacify the gap between demand and supply but also conserve the natural resources [24,25].

2 Materials and methods

2.1 Materials

The waste batteries used in this experiment were provided by the Lithium Battery Power Research Institute of Jiangxi University of Science and Technology. All batteries were discharged in saturated NaSO4 solution and disassembled by hand [26]. The removed positive electrode was washed with DMC (dimethyl carbonate) and was crushed mechanically with a grinder and screened to obtain LiFePO4 material, with a mass fraction of lithium in the LiFePO4 positive electrode material of 3.96% [27]. All the chemicals used in the experiment were analytically pure, and the solutions were prepared with ultrapure water.

2.2 Lithium preferential extraction

LiFePO4 was mixed with water, adding HCl and NaClO slowly over time with heated stirring and careful pH monitoring. Controlled drip rates for HCl and NaClO prevent the production of Cl2 gas [28,29].

2.3 High-efficiency lithium recovery

2.3.1 Purification

The pH of lithium-containing liquid for a certain volume was adjusted to a range of 9–10 with alkali, stirred at 85℃ for 1 h to remove impurities of Fe, Cu, Al, and others in the solution [30]. After filtration, the lithium-containing liquid was purified further by the addition of a sodium salt at a mass ratio of 0.3% and EDTA at a mass ratio of 0.1% as a complexing agent to adjust Ca2+ to 60 μg mL−1 [31].

2.3.2 Precipitation of lithium

The lithium solution was heated to 70–80℃, and Na2CO3 solution was added to the reaction dropwise to produce sediments of Li2CO3 [32]. This precipitate was vacuum filtered, washed with high purity water several times, and dried under vacuum to obtain high purity battery grade Li2CO3 [33].

2.3.3 Extraction

The solubility of lithium is 1.2–2.0 g L−1 in the precipitated mother liquor and accounts for 20–25% of the total lithium content [34]. Lithium extraction was performed using a mixed extraction agent (P507:TBP:NF = 10:2:1) in this experiment. The residual lithium was less than 50 μg L−1 in the final remaining liquid, giving a recovery rate of 98% [35,36].

2.4 Analytical equipment and characterization of materials

The solid samples were digested in concentrated hydrochloric acid for analysis using inductively coupled plasma optical emission spectrometry (ICP-OES, Thermo Fisher ICAP 7400). X-ray diffraction (XRD, D8 Advance, Bruker, Germany) was used to characterize the relative crystal phases during the experimental process and identify any structural changes in the samples in the reaction [37]. Scanning electron microscope (SEM) (Flex SEM 1000, Hitachi, Japan) was used to image the morphology of the products [38,39]. Particle size and the water content of the lithium products were measured by a particle size analyzer and moisture meter to determine if the Li2CO3 product meets the standards for cathode material synthesis [40,41].

3 Results and discussion

Figure 1 shows the influence of different parameters on the leaching efficiency. Meaning of each picture: (a) acid and oxidant molar ratios to Li and the impact on leaching efficiency; (b) HCl:Li ratio effect on unwanted Fe3+ leaching; (c) four temperature conditions are plotted against time to measure efficiency showing 60 min is optimal for leaching; (d) leaching efficiency against solid–liquid ratio and shows a minimal effect; (e) plot of pH against the HCl/Li molar ratio; (f) the relationship between pH of the solution and lithium leaching efficiency.

Figure 1 
               (a) Acid and oxidant molar ratios to Li and the impact on leaching efficiency; (b) HCl : Li ratio effect on unwanted Fe3+ leaching; (c) Four (4) temperature condition are plotted against time to measure efficiency showing 60 minutes is optimal for leaching; (d) Leaching efficiency against solid–liquid ratio and shows a minimal effect, (e) Plot of pH against the HCl/Li molar ratio; (f) the relationship between pH of the solution and lithium leaching efficiency.
Figure 1

(a) Acid and oxidant molar ratios to Li and the impact on leaching efficiency; (b) HCl : Li ratio effect on unwanted Fe3+ leaching; (c) Four (4) temperature condition are plotted against time to measure efficiency showing 60 minutes is optimal for leaching; (d) Leaching efficiency against solid–liquid ratio and shows a minimal effect, (e) Plot of pH against the HCl/Li molar ratio; (f) the relationship between pH of the solution and lithium leaching efficiency.

The addition of NaClO oxidizes Fe2+ → Fe3+ and facilitates the extraction of lithium from its structure. From simple stoichiometry, the moles of NaClO should equal the moles of lithium extracted, and the moles of HCl should be less than the moles of lithium if Fe3+ remains were intact. The rate of addition for HCl and NaClO should maintain a pH of 1.5–2; otherwise, Fe3+ will leach into the solution. Figure 1(a) shows the influence that the amount of NaClO and HCl has on the lithium dissolution rate, and “oxidant/Li and acid/Li molar ratio” expresses the ratio of NaClO and HCl to the molar amount of lithium, respectively. The concentration of Li+ ion in solution was measured by ICP. The calculated recovery rate of lithium was 99% using excess NaClO. The recovery rate was 100% with excess HCl; however, unwanted Fe3+ also dissolved. Figure 1(b) clearly shows that when the molar ratio of HCl and Li is greater than 0.8 (pH ∼ 0.5), the solution rate of Fe3+ increases linearly and further introduces impurities during the extraction of lithium. “Fe leaching efficiency” refers to the mass of the dissolved Fe ion to total iron in LiFePO4, again measured in solution by ICP. Figure 1(c) shows the influence of temperature and time on extraction. Exceeding 60 minutes, the temperature has no effect and the leaching effect of lithium is the best. Figure 1(d) shows that with optimal conditions, the size of the solid–liquid ratio has little influence on extraction above a 1:1 ratio. Figure 1(e) shows the influence of pH during the experiment by HCl addition and the effect on lithium yield. It can be seen from Figure 1(f) that when pH ≤ 2.0 (the value measured with a pH strip and related to the amount of HCl added), marking a small amount of HCl, the lithium yield is greater than 99% [40].

Figure 2 shows the diffraction patterns phase after the extraction of lithium. LFP refers to LiFePO4, LFP-1 refers to the slag that NaClO is used only 70% in the process of extracting lithium, LFP-2 and LFP-3 refers to the slag that the addition amount of NaClO and HCl is optimal. The two reflections (121) and (131) are intense peaks seen in LFP, whereas the leached samples show a significant reduction or absence. The outlined sections A and B of complete patterns correspond to Figure 2(a and b). There are new peaks near the (131) peaks in the XRD pattern of LFP-1 and still retains intensity at (131). However, there are no (131) peaks in the XRD pattern of LFP-2/3.

Figure 2 
               XRD patterns for different NaClO, and the outlined sections A and B of complete patterns correspond to Figure 2a and 2b.
Figure 2

XRD patterns for different NaClO, and the outlined sections A and B of complete patterns correspond to Figure 2a and 2b.

These patterns indicate that a small amount of LiFePO4 remains in the LFP-1 but is completely nonexistent in the LFP-2/3 samples. It can be seen that the structure of LiFePO4 changed in the process of preferential extraction of lithium in the XRD pattern. The main peak (011) and (111) of LiFePO4 did not change with the dissolution of lithium, indicating that only part of the raw material structure changed in the optimal solution process to realize the preferential extraction of lithium.

Figure 3(a–f) shows the morphology of the LiFePO4 cathode material and FePO4 residues. It comprises many primary and secondary particles (Figure 3a), and magnified images in Figure 3(b and c) show spherical agglomerates and much smaller spherical primary particles. Figure 3d shows the particle morphology of post-extraction FePO4 residues. There is an absence of the previously seen secondary agglomerated particles. In Figure 3(e and f), the primary particles are much smaller irregular shapes with some agglomeration. The extraction caused the secondary particles of LiFePO4 to break apart, possibly explained by the structural stress as Li+ ion is extracted and the crystal structure has volumetric changes from LiFePO4 → FePO4.

Figure 3 
               SEM Graph: LiFePO4 materials, (b) the primary particles of LiFePO4 materials, (c) the secondary particles of LiFePO4 materials, (d) the acid slag, (e) the secondary particles of the acid slag are absent, (f) the primary particles of the acid slag broken apart into smaller irregular shapes.
Figure 3

SEM Graph: LiFePO4 materials, (b) the primary particles of LiFePO4 materials, (c) the secondary particles of LiFePO4 materials, (d) the acid slag, (e) the secondary particles of the acid slag are absent, (f) the primary particles of the acid slag broken apart into smaller irregular shapes.

Figure 4 shows the XRD diagrams of battery-grade Li2CO3 products; compared with the standard card of Li2CO3, each peak of the two products is perfectly consistent with the standard peak, indicating that Li2CO3-1 and Li2CO3-2 are both Li2CO3 products.

Figure 4 
               XRD diagrams of battery-grade Li2CO3 products.
Figure 4

XRD diagrams of battery-grade Li2CO3 products.

Figure 5 shows the morphology of battery-grade Li2CO3-1 and Li2CO3-2 of Figure 4. The lithium-extracted solution was reacted with NaCO3 after impurity removal. The dropwise addition of the Na2CO3 solution permits particle ripening over nucleation to produce larger robust particles. This can be seen from the two batches of battery-grade Li2CO3 products in Figure 5 (Li2CO3-1 and Li2CO3-2).

Figure 5 
               Micrographs of battery-grade Li2CO3-1 and Li2CO3-2 products.
Figure 5

Micrographs of battery-grade Li2CO3-1 and Li2CO3-2 products.

Figure 6 plots the impurity fractions in the Li2CO3 products measured by the ICP analysis. Line A represents the impurity standard for battery-grade Li2CO3, and Line B and Line C represent the impurity content of two batches of Li2CO3 products respectively.

Figure 6 
               Comparison of the Li2CO3 battery-grade standard and product impurity content for several common ionic impurities.
Figure 6

Comparison of the Li2CO3 battery-grade standard and product impurity content for several common ionic impurities.

Table 1 presents the purity, moisture content, and particle size ranges for battery-grade Li2CO3 compared with the products from our Li extraction and precipitation methods. The products meet or exceed the standards for Li2CO3 in battery material synthesis.

Table 1

Purity, particle size, and water content of the Li2CO3 standards and the products

Name Li2CO3 content (%) (no less than) H2O(%) Size distribution (μm)
D10 D50 D90
Battery level standard 99.5 0.25 ≥0.5 3.0–8.0 9.0–15.0
Li2CO3-1 99.89 0.01 2 5.8 12
Li2CO3-2 99.93 0.015 1.5 6.2 11.8

The lithium concentration ranged from 1.2 to 2.0 g/L in the mother liquid, and this concentration accounts for 20–25% of the total lithium. The extraction of lithium from the mother liquid was performed using a mixed extraction agent (P507:TBP:NF = 10:2:1). Several mother solutions (numbers 1–6) with different lithium content were selected [41], and the residual lithium after extraction was measured with ICP, as presented in Table 2. The remaining lithium content was less than 50 µg/mL in the mother liquor, yielding a 98% recovery [42]. Combining these results with a leaching efficiency near 99% for lithium from LiFePO4, the total yields for lithium recycling averaged near 97% [43].

Table 2

The yield of lithium during extraction and lithium total yield

Extraction Number Lithium content Li content in raffinate (μg mL−1) Extraction yield of Li (%) Total recovery (%)
P507 + TBP + NF 1 1.24 35.5 99.093 98.102
2 1.31 41.3 98.945 97.956
3 1.58 49.1 98.619 97.633
4 1.67 44 98.876 97.887
5 1.86 24.5 99.374 98.380
6 1.95 28.7 99.267 98.274

4 Conclusion

A refined method has been tested for lithium extraction from the peridotite structure of LiFePO4 by controlling the addition of NaClO and HCl. Mole ratios of NaClO and HCl to moles of lithium at 1 and 0.8 produced the best leaching with minimized impurities. Compared with the traditional total solution process, the total acid and alkali needed are reduced to improve the economic and environ mental factors in the recycling process. A pH of 1.5–2 gave an optimal dissolution recovery rate (99%) and limited the dissolution of Fe3+ to below 2%.

High-quality battery-grade Li2CO3 was prepared by extraction using a mixed extraction agent (P507:TBP:NF = 10:2:1) with a recovery that gave a total yield of lithium above 97%. The economic benefits such as the efficiency of energy savings and consumption reduction also increased. This modified method showed an obvious advantage compared to traditional lithium extraction processes. It was helpful to establish recycled lithium as a viable pipeline for future resource and energy demands.

Acknowledgments

The authors would like to thank all the reviewers who participated in the review and MJ Editor for its linguistic assistance during the preparation of this manuscript. This work was supported by the National Natural Science Foundation of China (51874151), the Scientific Research Foundation for universities from the Education Bureau of Jiangxi Province (GJJ170510), the Natural Science Foundation of Jiangxi Province (20151BBE50106), and the Jiangxi University of Science and Technology (NSFJ2014-G13, Jxxjbs12005).

  1. Conflict of interest: The authors declare no conflict of interest regarding the publication of this paper.

References

[1] Ciez RE, Whitacre JF. Examining different recycling processes for lithium-ion batteries. Nat Sustain. 2019;2(2):148–56.10.1038/s41893-019-0222-5Search in Google Scholar

[2] Mosstafa K, Massoud G, Ali M. Cobalt ferrite nanoparticles (CoFe2O4 MNPs) as catalyst and support: magnetically recoverable nanocatalysts in organic synthesis. Nanotechnol Rev. 2018;7(1):43–68.10.1515/ntrev-2017-0138Search in Google Scholar

[3] Eckard H, Weiss M. Advances and critical aspects in the life-cycle assessment of battery electric cars. Energy Emiss Control Technol. 2017;5:1–18.10.2147/EECT.S60408Search in Google Scholar

[4] Nayaka GP, Yingjie Z, Peng D. Effective and environmentally friendly recycling process designed for LiCoO2 cathode powders of spent Li–ion batteries using mixture of mild organic acids. Waste Manag. 2018;78(AUG):51–7.10.1016/j.wasman.2018.05.030Search in Google Scholar PubMed

[5] Yunqi L, Danping X, Yuwei L, Mingtao L, Jinzhang L, Chen L, et al. Correlation between electrochemical performance degradation and catalyst structural parameters on polymer electrolyte membrane fuel cell. Nanotechnol Rev. 2019;8(1):493–502.10.1515/ntrev-2019-0044Search in Google Scholar

[6] Kumari N, Patel SR, Gohel JV. Current progress and future prospective of perovskite solar cells: a comprehensive review. Rev Adv Mater Sci. 2018;53:2.Search in Google Scholar

[7] Zhong RP, Xu JJ, Hui D, Bhosale SS, Hong R. Pyrolytic preparation and modification of carbon black recovered from waste tyres. Waste Manag Res. 2019;38(1):1–8.10.1177/0734242X19869987Search in Google Scholar PubMed

[8] He L, Sun S, Song X. Leaching process for recovering valuable metals from the LiNi1/3Co1/3Mn1/3O2 cathode of lithium-ion batteries. Waste Manag. 2017;64:171–81.10.1016/j.wasman.2017.02.011Search in Google Scholar PubMed

[9] Larcher D, Tarascon JM. Towards greener and more sustainable batteries for electrical energy storage. Nat Chem. 2015;7:19–29.10.1038/nchem.2085Search in Google Scholar PubMed

[10] Yan TT, Guo XM. Research progress of recycle techniques of spent lithium-ion batteries. Jiangxi Chem Ind. 2018;6:38–40 (in Chinese).Search in Google Scholar

[11] Li X, Zhang J, Song D, Song J, Zhang L. Direct regeneration of recycled cathode material mixture from scrapped LiFePO4 batteries. J Power Sources. 2017;345:78–84.10.1016/j.jpowsour.2017.01.118Search in Google Scholar

[12] Dutta D, Kumari A, Panda R, Jha S, Gupta D, Goel S, et al. Close loop separation process for the recovery of Co, Cu, Mn, Fe and Li from spent lithium-ion batteries. Sep Purif Technol. 2018;200:327–34.10.1016/j.seppur.2018.02.022Search in Google Scholar

[13] Zhang J, Hu J, Liu Y, Jing Q, Yang C, Chen Y, et al. Sustainable and facile method for the selective recovery of lithium from cathode scrap of spent LiFePO4 batteries. ACS Sustain Chem Eng. 2019;7:5626–31.10.1021/acssuschemeng.9b00404Search in Google Scholar

[14] Yang D, Zhaodong X, Dong H, Ning W. Theoretical-molar Fe3+ recovering lithium from spent LiFePO4 batteries: an acid-free, efficient, and selective process. J Hazard Mater. 2020;396:122707.10.1016/j.jhazmat.2020.122707Search in Google Scholar PubMed

[15] Liu Z, Kendall K, Yan X. China progress on renewable energy vehicles: fuel cells, hydrogen and battery hybrid vehicles. Energies. 2019;12:54–9.10.3390/en12010054Search in Google Scholar

[16] Li H, Xing S, Liu Y, Li F, Guo H, Kuang G. Recovery of lithium, iron, and phosphorus from spent LiFePO4batteries using stoichiometric sulfuric acid leaching system. ACS Sustain Chem Eng. 2017;5:8017–24.10.1021/acssuschemeng.7b01594Search in Google Scholar

[17] Nikolovski D, Jeremic M, Paunovic J, Vucevic D, Radosavljevic T, Radojević-Škodri S, et al. Application of iron oxide nanoparticles in contemporary experimental physiology and cell biology. Rev Adv Mater Sci. 2018;53:74–8.10.1515/rams-2018-0005Search in Google Scholar

[18] Yue Y, Xu S, He Y. Lithium recycling and cathode material regeneration from acid leach liquor of spent lithium-ion battery via facile co-extraction and co-precipitation processes. Waste Manag. 2017;64:219–27.10.1016/j.wasman.2017.03.018Search in Google Scholar PubMed

[19] Nirugaa N, Scharf M, Scharf P, Scharf P. Scenarios for the return of lithium-ion batteries out of electric cars for recycling. Procedia CIRP. 2015;29:740–5.10.1016/j.procir.2015.02.170Search in Google Scholar

[20] Liu CW, Lin J, Cao HB, Yi Z, Zhi S. Recycling of spent lithium-ion batteries in view of lithium recovery: a critical review. J Clean Prod. 2019;228:801–13.10.1016/j.jclepro.2019.04.304Search in Google Scholar

[21] Recycle spent batteries. Nat Energy. 2019;10:1038.Search in Google Scholar

[22] Tran MK, Rodrigues MTF, Kato K, Babu G, AjayanP M, Ganguli B, et al. Deep eutectic solvents for cathode recycling of Li–ion batteries. Nat Energy. 2019;10:1038–52.10.1038/s41560-019-0368-4Search in Google Scholar

[23] OrdoñezJ E, Jadraque G, Aymeric G. Processes and technologies for the recycling and recovery of spent lithium-ion batteries. Renew Sust Energ Rev. 2016;60:195–205.10.1016/j.rser.2015.12.363Search in Google Scholar

[24] Yang YX, Meng XQ, Cao HB, Lin X, Liu C, Sun Y, et al. Selective recovery of lithium from spent lithium iron phosphate batteries: a sustainable process. Green Chem. 2018;20:3121–33.10.1039/C7GC03376ASearch in Google Scholar

[25] Li H, Xing SZ, Liu Y, Li FJ, Guo H. Recovery of lithium, iron, and phosphorus from spent LiFePO4 batteries using stoichiometric sulfuric acid leaching system. ACS Sustain Chem Eng. 2017;5:8017–24.10.1021/acssuschemeng.7b01594Search in Google Scholar

[26] Yang Y, Zheng X, Zhao C, Lin X, Cao H, Ning P, et al. A closed-loop process for selective metal recovery from spent lithium iron phosphate batteries through mechanochemical activation. ACS Sustain Chem Eng. 2017;5:9972–80.10.1021/acssuschemeng.7b01914Search in Google Scholar

[27] Li Z, Liu DF, Xiong JC, He LH, Zhao ZW, Wang D. Selective recovery of lithium and iron phosphate/carbon from spent lithium iron phosphate cathode material by anionic membrane slurry electrolysis. Waste Manag. 2020;107:1–8.10.1016/j.wasman.2020.03.017Search in Google Scholar PubMed

[28] Gannavarapu KP, Thakkar M, Veerapaga S, Wei L, Dandamudi RB, Mitra S. Novel diatom-FeOx composite as highly active catalyst in photodegradation of Rhodamine-6G. Nanotechnol Rev. 2018;7:247–55.10.1515/ntrev-2017-0218Search in Google Scholar

[29] Chen S, He T, Lu Y, Su Y, Tian J, Li N, et al. Renovation of LiCoO2 with outstanding cycling stability by thermal treatment with Li2CO3 from spent Li–ion batteries. J Energy Storage. 2016;8:262–73.10.1016/j.est.2016.10.008Search in Google Scholar

[30] Sonoc A, Jeswiet J, Soo VK. Opportunities to improve recycling of automotive lithium ion batteries. Procedia CIRP. 2015;29:752–7.10.1016/j.procir.2015.02.039Search in Google Scholar

[31] Li L, Bian Y, Zhang X, Xue Q, Fan E, Wu F, et al. Economical recycling process for spent lithium-ion batteries and macro- and micro-scale mechanistic study. J Power Sources. 2018;377:70–9.10.1016/j.jpowsour.2017.12.006Search in Google Scholar

[32] Sun Z, Xiao Y, Sietsma J, Agterhuis H, Yang Y. A cleaner process for selective recovery of valuable metals from electronic waste of complex mixtures of end-of-life electronic products. Env Sci Technol. 2015;49:7981–8.10.1021/acs.est.5b01023Search in Google Scholar PubMed

[33] Fan E, Li L, Zhang X, Bian Y, Xue Q, Wu J, et al. Selective recovery of Li and Fe from spent lithium-ion batteries by an environmentally friendly mechanochemical approach. ACS Sustain Chem Eng. 2018;6:11029–35.10.1021/acssuschemeng.8b02503Search in Google Scholar

[34] Yang YX, Meng XQ, Cao HB, Lin X, Liu CM, Sun Y, et al. Selective recovery of lithium from spent lithium iron phosphate batteries: a sustainable process. Green Chem. 2018;20:3121.10.1039/C7GC03376ASearch in Google Scholar

[35] Li L, Bian Y, Zhang X, Yao Y, Xue Q, Fan E, et al. A green and effective room-temperature recycling process of LiFePO4 cathode materials for lithium-ion batteries. Waste Manag. 2019;85:437–44.10.1016/j.wasman.2019.01.012Search in Google Scholar PubMed

[36] Nayl AA, Elkhashab RA, Badawy SM, El-Khateeb MA. Acid leaching of mixed spent Li–ion batteries. Arab J Chem. 2017;10:S3632–9.10.1016/j.arabjc.2014.04.001Search in Google Scholar

[37] Ali Z, Li T, Khan M, Ali N, Zhang Q. Immobilization of lipase on iron oxide organic/inorganic hybrid particles: a review article. Rev Adv Mater Sci. 2018;53:106–17.10.1515/rams-2018-0008Search in Google Scholar

[38] Yang Y, Xu S, He Y. Lithium recycling and cathode material regeneration from acid leach liquor of spent lithium-ion battery via facile co-extraction and coprecipitation processes. Waste Manag. 2017;64:219–27.10.1016/j.wasman.2017.03.018Search in Google Scholar PubMed

[39] Harper G, Sommerville R, Kendrick E, Driscoll L, Slater P, Stolkin R, et al. Recycling lithium-ion batteries from electric vehicles. Nature. 2019;575:75–86.10.1038/s41586-019-1682-5Search in Google Scholar PubMed

[40] Ku H, Jung YJ, Jo M, Park S, Kim S, Yang D, et al. Recycling of spent lithium-ion battery cathode materials by ammoniacal leaching. J Hazard Mater. 2016;313:138–46.10.1016/j.jhazmat.2016.03.062Search in Google Scholar PubMed

[41] Liu K, Tan Q, Liu LL, Li JH. Acid-free and selective extraction of lithium from spent lithium iron phosphate batteries via a mechanochemically induced isomorphic substitution. Env Sci Technol. 2019;53:9781–8.10.1021/acs.est.9b01919Search in Google Scholar PubMed

[42] Setiawan H, Petrus HTBM, Perdana I. Reaction kinetics modeling for lithium and cobalt recovery from spent lithium-ion batteries using acetic acid. Int J Miner Metall Mater. 2019;26:98–107.10.1007/s12613-019-1713-0Search in Google Scholar

[43] Zheng Y, Song W, Mo W, Zhou L, Liu JW. Lithium fluoride recovery from cathode material of spent lithium-ion battery. RSC Adv. 2018;8:8990–8.10.1039/C8RA00061ASearch in Google Scholar PubMed PubMed Central

Received: 2020-11-23
Revised: 2020-12-22
Accepted: 2020-12-24
Published Online: 2020-12-31

© 2020 Tingting Yan et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 19.4.2024 from https://www.degruyter.com/document/doi/10.1515/ntrev-2020-0119/html
Scroll to top button