[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (294)

Search Parameters:
Keywords = lithium-ion battery recycling

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
18 pages, 4990 KiB  
Article
Disassembly and Its Obstacles: Challenges Facing Remanufacturers of Lithium-Ion Traction Batteries
by Gregor Ohnemüller, Marie Beller, Bernd Rosemann and Frank Döpper
Processes 2025, 13(1), 123; https://doi.org/10.3390/pr13010123 - 5 Jan 2025
Viewed by 504
Abstract
Lithium-ion batteries are major drivers to decarbonize road traffic and electric power systems. With the rising number of electric vehicles comes an increasing number of lithium-ion batteries reaching their end of use. After their usage, several strategies, e.g., reuse, repurposing, remanufacturing, or material [...] Read more.
Lithium-ion batteries are major drivers to decarbonize road traffic and electric power systems. With the rising number of electric vehicles comes an increasing number of lithium-ion batteries reaching their end of use. After their usage, several strategies, e.g., reuse, repurposing, remanufacturing, or material recycling can be applied. In this context, remanufacturing is the favored end-of-use strategy to enable a new use cycle of lithium-ion batteries and their components. The process of remanufacturing itself is the restoration of a used product to at least its original performance by disassembling, cleaning, sorting, reconditioning, and reassembling. Thereby, disassembly as the first step is a decisive process step, as it creates the foundation for all further steps in the process chain and significantly determines the economic feasibility of the remanufacturing process. The aim of the disassembly depth is the replacement of individual cells to replace the smallest possible deficient unit and not, as is currently the case, the entire battery module or even the entire battery system. Consequently, disassembly sequences are derived from a priority matrix, a disassembly graph is generated, and the obstacles to non-destructive cell replacement are analyzed for two lithium-ion traction battery systems, to analyze the distinctions between battery electric vehicle (BEV) and plug-in hybrid electric vehicle (PHEV) battery systems and identify the necessary tools and fundamental procedures required for the effective management of battery systems within the circular economy. Full article
(This article belongs to the Special Issue Green Manufacturing and Energy-Efficient Production)
Show Figures

Figure 1

Figure 1
<p>Main components of the BMW F48 X1 xDrive25e numbered and named in <a href="#processes-13-00123-t004" class="html-table">Table 4</a>.</p>
Full article ">Figure 2
<p>Disassembly priority graph for the BMW F48 X1 xDrive25e.</p>
Full article ">Figure 3
<p>Main components of the SMART EQ battery system.</p>
Full article ">Figure 4
<p>Disassembly priority graph.</p>
Full article ">Figure 5
<p>BMW F48 xDrive25e battery module. (<b>a</b>) The red rectangles mark the opened weld seams. (<b>b</b>) Spot-welded contacts and the partially removed wiring harness (voltage/current and temperature sensors).</p>
Full article ">Figure 6
<p>SMART EQ battery module. (<b>a</b>) Module housing with yellow rings marking the rivets. (<b>b</b>) Opened module housing and unfanned cells.</p>
Full article ">
22 pages, 2361 KiB  
Review
Advances in Recycling Technologies of Critical Metals and Resources from Cathodes and Anodes in Spent Lithium-Ion Batteries
by Shuwen Wang, Yanrong Lai, Jingran Yang, Jiaxue Zhao, Yushan Zhang, Miaoling Chen, Jinfeng Tang, Junhua Xu and Minhua Su
Separations 2025, 12(1), 4; https://doi.org/10.3390/separations12010004 - 30 Dec 2024
Viewed by 316
Abstract
With the rapid economic development and the continuous growth in the demand for new energy vehicles and energy storage systems, a significant number of waste lithium-ion batteries are expected to enter the market in the future. Effectively managing the processing and recycling of [...] Read more.
With the rapid economic development and the continuous growth in the demand for new energy vehicles and energy storage systems, a significant number of waste lithium-ion batteries are expected to enter the market in the future. Effectively managing the processing and recycling of these batteries to minimize environmental pollution is a major challenge currently facing the lithium-ion battery industry. This paper analyzes and compares the recycling strategies for different components of lithium-ion batteries, providing a summary of the main types of batteries, existing technologies at various pre-treatment stages, and recycling techniques for valuable resources such as heavy metals and graphite. Currently, pyrometallurgy and hydrometallurgy processes have matured; however, their high energy consumption and pollution levels conflict with the principles of the current green economy. As a result, innovative technologies have emerged, aiming to reduce energy consumption while achieving high recovery rates and minimizing the environmental impact. Nevertheless, most of these technologies are currently limited to the laboratory scale and are not yet suitable for large-scale application. Full article
(This article belongs to the Section Purification Technology)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Applications of lithium-ion batteries; (<b>b</b>) The shape and components of some Li-ion battery configurations; (<b>c</b>) Flow-chart showing the typical recycling process; (<b>d</b>) Schematic diagram of the LIB working principle.</p>
Full article ">Figure 2
<p>Schematic diagrams of pyrometallurgy, hydrometallurgy, and direct recovery processes [<a href="#B48-separations-12-00004" class="html-bibr">48</a>].</p>
Full article ">Figure 3
<p>Characteristics of different pyrometallurgical technologies used to treat spent LIBs for the recovery of strategic metals.</p>
Full article ">
19 pages, 3282 KiB  
Article
The Effect of Plasma Pretreatment on the Flotation of Lithium Aluminate and Gehlenite Using Light-Switchable Collectors
by Ali Zgheib, Maximilian Hans Fischer, Stéphanie Mireille Tsanang, Iliass El Hraoui, Shukang Zhang, Annett Wollmann, Alfred P. Weber, Ursula E. A. Fittschen, Thomas Schirmer and Andreas Schmidt
Separations 2024, 11(12), 362; https://doi.org/10.3390/separations11120362 - 23 Dec 2024
Viewed by 364
Abstract
The pyridinium phenolate punicine is a switchable molecule from Punica granatum. Depending on the pH, punicine exists as a cation, neutral molecule, anion, or dianion. In addition, punicine reacts to light, under the influence of which it forms radical species. We report [...] Read more.
The pyridinium phenolate punicine is a switchable molecule from Punica granatum. Depending on the pH, punicine exists as a cation, neutral molecule, anion, or dianion. In addition, punicine reacts to light, under the influence of which it forms radical species. We report on three punicine derivatives that possess an adamantyl, 2-methylnonyl, or heptadecyl substituent and on their performance in the flotation of lithium aluminate, an engineered artificial mineral (EnAM) for the recycling of lithium, e.g., from lithium-ion batteries. By optimizing the parameters: pH and light conditions (daylight, darkness), recovery rates of 92% of LiAlO2 are achieved. In all cases, the flotation of the gangue material gehlenite (Ca2Al[AlSiO7]) is suppressed. IR, the contact angle, zeta potential measurements, TG-MS, and PXRD confirm that the punicines interact with the surface of LiAlO2, which is covered by LiAl2(OH)7 after contact to water, resulting in a hydrophobization of the particle. The plasma pretreatment of the lithium aluminate has a significant influence on the flotation results and increases the recovery rates of lithium aluminate in blank tests by 58%. The oxidative plasma leads to a partial dehydratisation of the LiAl2(OH)7 and thus to a hydrophobization of the particles, while a reductive plasma causes a more hydrophilic particle surface. Full article
(This article belongs to the Special Issue Green Separation and Purification Technology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>TGA measurements of pure lithium aluminate and 5 min oxidative plasma-pretreated lithium aluminate samples.</p>
Full article ">Figure 2
<p>TG-MS of pure LiAlO<sub>2</sub> (<b>left</b>) and 5 min plasma pretreated LiAlO<sub>2</sub> (<b>right</b>).</p>
Full article ">Figure 3
<p>Comparison of the recovery rates of lithium aluminate at different light scenarios in the presence of the collectors <b>1</b>–<b>3</b> under identical conditions, i.e., 2.00 g of LiAlO<sub>2</sub>, 1 min of mixing (500 rpm) with 25 mL of distilled water, 60 µL (1 × 10<sup>−5</sup> M) of the corresponding collector, 1 min of mixing, 30 µL of frother, and another 1 min of mixing. The flotation was done for 3 min at an air flow rate of 32 mL/min using 250 mL distilled water in daylight at a pH of 10.9 ± 0.3.</p>
Full article ">Figure 4
<p>Comparison of the recovery rates of lithium aluminate at different pH values in the presence of the collectors <b>1</b>–<b>3</b> under identical conditions, i.e., 2.00 g of LiAlO<sub>2</sub>, 1 min of mixing (500 rpm) with 25 mL of distilled water, 60 µL (1 × 10<sup>−5</sup> M) of the corresponding collector, 1 min of mixing, 30 µL of frother, and another 1 min of mixing. The flotation was done for 3 min at an air flow rate of 32 mL/min using 250 mL distilled water in daylight.</p>
Full article ">Figure 5
<p>Single mineral flotation of lithium aluminate and gehlenite using punicine <b>3</b> as collector at pH 10.9 ± 0.3 for lithium aluminate and pH 8.3 ± 0.3 for gehlenite at different light scenarios. The conditioning was as follows: 120 µL collector, 1 min mixing, 30 µL frother, 1 min mixing, stirring speed 500 rpm, air flow rate of 32 cm<sup>3</sup>/min using 250 mL distilled water, and flotation time of 3 min.</p>
Full article ">Figure 6
<p>Single mineral flotation of plasma-pretreated lithium aluminate and the gehlenite-rich material without collector at pH 10.9 ± 0.3 for lithium aluminate and pH 8.3 ± 0.3 for gehlenite as a function of plasma treatment time. Conditioning was as follows: 30 µL frother, 1 min mixing, stirring speed 500 rpm, air flow rate of 32 cm<sup>3</sup>/min using 250 mL distilled water, and flotation time of 3 min.</p>
Full article ">Figure 7
<p>Single mineral flotation of the oxidative plasma pretreated lithium aluminate and gehlenite using punicine <b>3</b> as a collector at pH 10.9 ± 0.3 for lithium aluminate, and pH 8.3 for gehlenite as a function of plasma treatment time in min. The conditioning was as follows: 60 µL collector, 1 min mixing 30 µL frother, 1 min mixing, stirring speed of 500 rpm, air flow rate of 32 cm<sup>3</sup>/min using 250 mL distilled water, and applying a flotation time of 3 min.</p>
Full article ">Figure 8
<p>Single mineral flotation of the reductive plasma pretreated lithium aluminate in blank and using punicine <b>3</b> as a collector at pH 10.9 ± 0.3 as a function of plasma treatment time in min. The conditioning was as follows: 60 µL collector, 1 min mixing 30 µL frother, 1 min mixing, stirring speed of 500 rpm, air flow rate of 32 cm<sup>3</sup>/min using 250 mL distilled water, and flotation time of 3 min.</p>
Full article ">Figure 9
<p>Contact angle on LiAlO<sub>2</sub> before and after flotation with punicine <b>3</b> under identical conditions, i.e., 2.00 g of the mineral, 1 min mixing (500 rpm) with 25 mL distilled water, 120 µL (2 × 10<sup>−5</sup> M) of the corresponding collector, 1 min mixing, 30 µL foaming agent, and 1 min mixing again. Flotation was carried out for 3 min at an air flow rate of 32 mL/min with 250 mL of distilled water in daylight and a pH of 10.9 ± 0.3.</p>
Full article ">Figure 10
<p>Zeta potential of pure lithium aluminate, lithium aluminate with 7.2 µL of 42.5 mmol/L punicine <b>1</b>, lithium aluminate with 7.2 µL of 42.5 mmol/L punicine <b>2,</b> and lithium aluminate with 7.2 µL of 42.5 mmol/L punicine <b>3</b> in distilled water and under different pH values.</p>
Full article ">Figure 11
<p>FTIR spectra of LiAlO<sub>2</sub> in the presence of 60 µL 42.5 µmol/L punicine <b>3</b> with and without pretreatment with plasma.</p>
Full article ">Figure 12
<p>Zeta potential of pure and oxidative plasma-pretreated lithium aluminate, with and without 10 µL of 42.5 mmol/L punicine <b>3</b> at pH 10.9 ± 0.3.</p>
Full article ">Scheme 1
<p>Schematic representation of the switchability of punicine with respect to charges and radical status.</p>
Full article ">Scheme 2
<p>Intermolecular interactions and complexes of punicine.</p>
Full article ">Scheme 3
<p>Synthesis of the punicine derivatives.</p>
Full article ">Scheme 4
<p>Reactions of lithium aluminate surfaces with water.</p>
Full article ">Scheme 5
<p>Interactions of punicines with lithium aluminum surfaces.</p>
Full article ">
17 pages, 12063 KiB  
Article
The CaO Enhanced Defluorination and Air-Jet Separation of Cathode-Active Material Coating for Direct Recycling Li-Ion Battery Electrodes
by Piotr Siwak, Volf Leshchynsky, Emil Strumban, Mircea Pantea, Dariusz Garbiec and Roman Maev
Metals 2024, 14(12), 1466; https://doi.org/10.3390/met14121466 - 23 Dec 2024
Viewed by 374
Abstract
With the rapid growth of the lithium-ion battery (LIBs) market, recycling and re-using end-of-life LIBs to reclaim the critical Li, Co, Ni, and Mn has become an urgent task. Presently, high temperature, strong acid, and alkali conditions are required to extract blended critical [...] Read more.
With the rapid growth of the lithium-ion battery (LIBs) market, recycling and re-using end-of-life LIBs to reclaim the critical Li, Co, Ni, and Mn has become an urgent task. Presently, high temperature, strong acid, and alkali conditions are required to extract blended critical metals (CM) from the typical battery cathode. Hence, there is a need for more effective recycling processes for recycling blended Li, Co, Ni, and their direct regeneration for re-use in LIBs. The goal of the offered paper is the development of recycling technology for degraded battery cathode-active materials based on the thermal decomposition of polyvinylidene fluoride (PVDF) using calcination and air-jet stripping of active materials. The proposed air-jet erosion method of calcined cathode material stripping from Al foil allows for the flexible industry-applicable separation process, which is damage-free for both particles and substrate. The CaO calcination air-jet separation process and equipment can significantly improve the PVDF decomposition and the separation efficiency of the cathode materials. It is demonstrated that low-temperature CaO calcination at 350–450 °C associated with air-jet separation of active material is characterized by low environmental impact, high purity of the recycled material, and low cost as compared to pyro- and hydrometallurgical methods. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Cathode structure schematics; (<b>b</b>) Example of sample cut from cathode tape before thermal treatment (dimensions: 25.4 mm × 150 mm); (<b>c</b>) Aluminum foil after cathode heat treatment at 600 °C and passing between serrated jaws; (<b>d</b>) Aluminum foil after cathode heat treatment at 550 °C and vibration at 60 Hz between stainless-steel mesh pieces.</p>
Full article ">Figure 2
<p>Flowchart of technology examination.</p>
Full article ">Figure 3
<p>Air erosion test for determination of air erosion resistance. (<b>a</b>) schematics of the test rig; (<b>b</b>) Interaction of air stream with the active material layer.</p>
Full article ">Figure 4
<p>NMC weight loss during the cathode calcination and thermogravimetric curves at the temperatures 50–600 °C.</p>
Full article ">Figure 5
<p>TGA curves of low-temperature NMC powder calcination (<b>a</b>), NMC + CaO calcination (<b>b</b>), and heating of preliminary calcined NMC (<b>c</b>).</p>
Full article ">Figure 6
<p>The conversion rate-temperature dependences for the first (<b>a</b>) and second (<b>b</b>) NMC active material decomposition stages.</p>
Full article ">Figure 7
<p>SEM and EDS of spent LIB cathode NMC active material. (<b>a</b>,<b>b</b>) SEM images; (<b>c</b>) Elemental analysis by EDX.</p>
Full article ">Figure 8
<p>SEM and EDX of spent LIB cathode LiCoO<sub>2</sub> active material. (<b>a</b>,<b>b</b>) SEM micrograph; (<b>c</b>) Elemental analysis by EDX.</p>
Full article ">Figure 9
<p>SEM and EDS of the LIB cathode NMC active material calcined at 400 °C. (<b>a</b>,<b>b</b>) SEM micrograph; (<b>c</b>) calcined with CaO; (<b>d</b>) elemental analysis by EDX.</p>
Full article ">Figure 10
<p>SEM and EDS of the LIB cathode LiCoO<sub>2</sub> active material calcined at 400 °C. (<b>a</b>,<b>b</b>) SEM micrographs; (<b>c</b>) Elemental analysis by EDX.</p>
Full article ">Figure 11
<p>Erosion spot topography of active materials calcined at 400 °C during 30 min: (<b>a</b>) NMC, (<b>b</b>) LiCoO<sub>2</sub>.</p>
Full article ">Figure 12
<p>Air-jet erosion spot geometry dependence on calcination temperature for NMC active material.</p>
Full article ">Figure 13
<p>SEM micrographs of LiCoO<sub>2</sub> cathode cross-section before (<b>a</b>) and after (<b>b</b>) air-jet erosion test.</p>
Full article ">Figure 14
<p>Schematics of calcination-separation machine for cathode-active material recycling.</p>
Full article ">
18 pages, 3016 KiB  
Article
Concepts for the Sustainable Hydrometallurgical Processing of End-of-Life Lithium Iron Phosphate (LFP) Batteries
by Marius Müller, Hüseyin Eren Obuz, Sebastian Keber, Firat Tekmanli, Luka Nils Mettke and Bengi Yagmurlu
Sustainability 2024, 16(24), 11267; https://doi.org/10.3390/su162411267 - 23 Dec 2024
Viewed by 604
Abstract
Lithium-ion batteries with an LFP cell chemistry are experiencing strong growth in the global battery market. Consequently, a process concept has been developed to recycle and recover critical raw materials, particularly graphite and lithium. The developed process concept consists of a thermal pretreatment [...] Read more.
Lithium-ion batteries with an LFP cell chemistry are experiencing strong growth in the global battery market. Consequently, a process concept has been developed to recycle and recover critical raw materials, particularly graphite and lithium. The developed process concept consists of a thermal pretreatment to remove organic solvents and binders, flotation for anode–cathode separation, and hydrometallurgical processes for product recovery. It has been shown that a pretreatment step is necessary for efficient flotation. By increasing the thermal treatment temperatures up to 450 °C, recovery rates of up to 73% are achieved. Similar positive effects are observed with leaching, where leaching efficiencies increase with higher treatment temperatures up to 400 °C. The results indicate that the thermal treatment of the black mass significantly influences both flotation and hydrometallurgical processes. Full article
Show Figures

Figure 1

Figure 1
<p>Generalized LFP processing methods.</p>
Full article ">Figure 2
<p>XRD diffractogram of the initial LFP sample.</p>
Full article ">Figure 3
<p>TGA data of the LFP black mass and possible reactions occurring during the thermal treatment [<a href="#B21-sustainability-16-11267" class="html-bibr">21</a>,<a href="#B22-sustainability-16-11267" class="html-bibr">22</a>,<a href="#B23-sustainability-16-11267" class="html-bibr">23</a>].</p>
Full article ">Figure 4
<p>Flotation behavior in the dependency of the thermal treatment. (<b>a</b>) Improper foam formation (thermal treatment at 350 °C; flotation parameters: cell volume, 0.125 L; flotation time, 10 min; conditioning time—collector, 3 min; conditioning time—foamer, 2 min; collector concentration, 350 g/t; foamer concentration, 150 g/t). (<b>b</b>) Foam after incomplete flotation. (<b>c</b>) Proper foam formation (thermal treatment at 450 °C; flotation parameters: cell volume, 0.125 L; flotation time, 10 min; conditioning time—collector, 3 min; conditioning time—foamer, 2 min; collector concentration, 350 g/t; foamer concentration, 150 g/t). (<b>d</b>) Cathode concentration after the flotation.</p>
Full article ">Figure 5
<p>Influence of the roasting temperature during the thermal pretreatment for 60 min on the graphite flotation efficiency (flotation parameters: cell volume, 0.125 L; flotation time, 10 min; conditioning time—collector, 3 min; conditioning time—foamer, 2 min; collector concentration, 350 g/t; foamer concentration, 150 g/t).</p>
Full article ">Figure 6
<p>Leaching yields with different stoichiometric ratios (STC) of acid concentrations (leaching conditions: solid/liquid ratio, 1/10; leaching time, 60 min; leaching temperature, 30 °C).</p>
Full article ">Figure 7
<p>Effect of oxidative media (H<sub>2</sub>O<sub>2</sub>) on the leaching yields (leaching conditions: stoichiometric amount of H<sub>2</sub>SO<sub>4</sub>; solid/liquid ratio, 1/10; leaching time, 60 min; leaching temperature, 30 °C).</p>
Full article ">Figure 8
<p>Effect of different roasting temperatures (roasting in a muffle furnace under atmospheric conditions with a heating rate of 10 K/min) on the leaching (leaching conditions: stoichiometric amount of H<sub>2</sub>SO<sub>4</sub>; solid/liquid ratio, 1/10; leaching time, 60 min; leaching temperature, 30 °C).</p>
Full article ">Figure 9
<p>The pH vs. precipitation yields of the constituent cathode metals.</p>
Full article ">Figure 10
<p>Flow sheet for the proposed process for leaching LFP black mass using phosphoric acid.</p>
Full article ">Figure 11
<p>Diffractogram of the precipitation product.</p>
Full article ">Figure 12
<p>Possible process routes for the recycling of LFP batteries.</p>
Full article ">
25 pages, 1241 KiB  
Review
Recycling Lithium-Ion Batteries—Technologies, Environmental, Human Health, and Economic Issues—Mini-Systematic Literature Review
by Geani Teodor Man, Andreea Maria Iordache, Ramona Zgavarogea and Constantin Nechita
Membranes 2024, 14(12), 277; https://doi.org/10.3390/membranes14120277 - 21 Dec 2024
Viewed by 869
Abstract
Global concerns about pollution reduction, associated with the continuous technological development of electronic equipment raises challenge for the future regarding lithium-ion batteries exploitation, use, and recovery through recycling of critical metals. Several human and environmental issues are reported, including related diseases caused by [...] Read more.
Global concerns about pollution reduction, associated with the continuous technological development of electronic equipment raises challenge for the future regarding lithium-ion batteries exploitation, use, and recovery through recycling of critical metals. Several human and environmental issues are reported, including related diseases caused by lithium waste. Lithium in Li-ion batteries can be recovered through various methods to prevent environmental contamination, and Li can be reused as a recyclable resource. Classical technologies for recovering lithium from batteries are associated with various environmental issues, so lithium recovery remains challenging. However, the emergence of membrane processes has opened new research directions in lithium recovery, offering hope for more efficient and environmentally friendly solutions. These processes can be integrated into current industrial recycling flows, having a high recovery potential and paving the way for a more sustainable future. A second method, biolexivation, is eco-friendly, but this point illustrates significant drawbacks when used on an industrial scale. We discussed toxicity induced by metals associated with Li to iron-oxidizing bacteria, which needs further study since it causes low recycling efficiency. One major environmental problem is the low efficiency of the recovery of Li from the water cycle, which affects global-scale safety. Still, electromembranes can offer promising solutions in the future, but there is needed to update regulations to actual needs for both producing and recycling LIB. Full article
(This article belongs to the Section Membrane Applications for Energy)
Show Figures

Figure 1

Figure 1
<p>The number of articles published yearly is based on the imposed interrogation criteria in the WOS online database and the cumulative number of citations per year.</p>
Full article ">Figure 2
<p>Schematic interaction between different Li sources of lithium and the effects of contamination in the environment and humans.</p>
Full article ">
33 pages, 5779 KiB  
Review
Electric Vehicle Battery Technologies and Capacity Prediction: A Comprehensive Literature Review of Trends and Influencing Factors
by Vo Tri Duc Sang, Quang Huy Duong, Li Zhou and Carlos F. A. Arranz
Batteries 2024, 10(12), 451; https://doi.org/10.3390/batteries10120451 - 19 Dec 2024
Viewed by 1456
Abstract
Electric vehicle (EV) battery technology is at the forefront of the shift towards sustainable transportation. However, maximising the environmental and economic benefits of electric vehicles depends on advances in battery life cycle management. This comprehensive review analyses trends, techniques, and challenges across EV [...] Read more.
Electric vehicle (EV) battery technology is at the forefront of the shift towards sustainable transportation. However, maximising the environmental and economic benefits of electric vehicles depends on advances in battery life cycle management. This comprehensive review analyses trends, techniques, and challenges across EV battery development, capacity prediction, and recycling, drawing on a dataset of over 22,000 articles from four major databases. Using Dynamic Topic Modelling (DTM), this study identifies key innovations and evolving research themes in battery-related technologies, capacity degradation factors, and recycling methods. The literature is structured into two primary themes: (1) “Electric Vehicle Battery Technologies, Development & Trends” and (2) “Capacity Prediction and Influencing Factors”. DTM revealed pivotal findings: advancements in lithium-ion and solid-state batteries for higher energy density, improvements in recycling technologies to reduce environmental impact, and the efficacy of machine learning-based models for real-time capacity prediction. Gaps persist in scaling sustainable recycling methods, developing cost-effective manufacturing processes, and creating standards for life cycle impact assessment. Future directions emphasise multidisciplinary research on new battery chemistries, efficient end-of-life management, and policy frameworks that support circular economy practices. This review serves as a resource for stakeholders to address the critical technological and regulatory challenges that will shape the sustainable future of electric vehicles. Full article
Show Figures

Figure 1

Figure 1
<p>Methodology framework.</p>
Full article ">Figure 2
<p>Distribution of articles by publication year.</p>
Full article ">Figure 3
<p>Top 20 journals by article count.</p>
Full article ">Figure 4
<p>Publication trends by journal (top 20 journals).</p>
Full article ">Figure 5
<p>Coherence score vs. number of topics (k) for Dynamic Topic Modelling of Theme 1: “Electric Vehicle Battery Technologies, Development &amp; Trend”.</p>
Full article ">Figure 6
<p>Coherence score vs. number of topics (k) for Dynamic Topic Modelling of Theme 2: “Electric Vehicle Battery Capacity Prediction: Influencing Factors”.</p>
Full article ">Figure 7
<p>Sematic keyword visualisation in Theme 1 in 1976 [<a href="#B6-batteries-10-00451" class="html-bibr">6</a>,<a href="#B7-batteries-10-00451" class="html-bibr">7</a>].</p>
Full article ">Figure 8
<p>Sematic keywords visualisation in Theme 1 in 2024 [<a href="#B6-batteries-10-00451" class="html-bibr">6</a>,<a href="#B7-batteries-10-00451" class="html-bibr">7</a>].</p>
Full article ">Figure 9
<p>Sematic keywords visualisation in Theme 2 in 1976 [<a href="#B6-batteries-10-00451" class="html-bibr">6</a>,<a href="#B7-batteries-10-00451" class="html-bibr">7</a>].</p>
Full article ">Figure 10
<p>Sematic keywords visualisation in Theme 2 in 2024 [<a href="#B6-batteries-10-00451" class="html-bibr">6</a>,<a href="#B7-batteries-10-00451" class="html-bibr">7</a>].</p>
Full article ">
18 pages, 3357 KiB  
Article
Deep Eutectic Solvent (TOPO/D2EHPA/Menthol) for Extracting Metals from Synthetic Hydrochloric Acid Leachates of NMC-LTO Batteries
by Arina V. Kozhevnikova, Nikita A. Milevskii, Dmitriy V. Lobovich, Yulia A. Zakhodyaeva and Andrey A. Voshkin
Metals 2024, 14(12), 1441; https://doi.org/10.3390/met14121441 - 16 Dec 2024
Viewed by 446
Abstract
The recycling of lithium-ion batteries is increasingly important for both resource recovery and environmental protection. However, the complex composition of cathode and anode materials in these batteries makes the efficient separation of metal mixtures challenging. Hydrometallurgical methods, particularly liquid extraction, provide an effective [...] Read more.
The recycling of lithium-ion batteries is increasingly important for both resource recovery and environmental protection. However, the complex composition of cathode and anode materials in these batteries makes the efficient separation of metal mixtures challenging. Hydrometallurgical methods, particularly liquid extraction, provide an effective means of separating metal ions, though they require periodic updates to their extraction systems. This study introduces a hydrophobic deep eutectic solvent composed of trioctylphosphine oxide, di(2-ethylhexyl)phosphoric acid, and menthol, which is effective for separating Ti(IV), Co(II), Mn(II), Ni(II), and Li+ ions from hydrochloric acid leachates of NMC (LiNixMnyCo1−x−yO2) batteries with LTO (Li4Ti5O12) anodes. By optimising the molar composition of the trioctylphosphine oxide/di(2-ethylhexyl)phosphoric acid/menthol mixture to a 4:1:5 ratio, high extraction efficiency was achieved. The solvent demonstrated stability over 10 cycles, and conditions for its regeneration were successfully established. At room temperature, the DES exhibited a density of 0.89 g/mL and a viscosity of 56 mPa·s, which are suitable for laboratory-scale extraction processes. Experimental results from a laboratory setup with mixer-settlers confirmed the efficiency of separating Ti(IV) and Co(II) ions in the context of their extraction kinetics. Full article
(This article belongs to the Section Extractive Metallurgy)
Show Figures

Figure 1

Figure 1
<p>A photo and scheme of a laboratory installation in operation: <span class="html-italic">a</span>—raffinate solution, <span class="html-italic">b</span>—extract solution, <span class="html-italic">c</span>—initial solution with appropriate metal concentration and HCl, <span class="html-italic">d</span>—HDES, P—pump, E—extractor.</p>
Full article ">Figure 2
<p>Dependence of density (<b>a</b>) and dynamic viscosity (<b>b</b>) of the HDES composed of TOPO/D2EHPA/menthol in a 4:1:5 ratio on temperature.</p>
Full article ">Figure 3
<p>Dependence of the degree of extraction of metal ions on the menthol content in the HDES. HDES phase: χ<sub>TOPO</sub>/χ<sub>D2EHPA</sub> = 1:1. Aqueous phase: [HCl] = 8 mol/L; V<sub>aq</sub>/V<sub>HDES</sub> = 1/1; mixing time was 30 min.</p>
Full article ">Figure 4
<p>The dependence of the degree of extraction of metal ions on the content of TOPO and D2EHPA in the HDES. HDES phase: χ<sub>menthol</sub> = 4. Aqueous phase: [HCl] = 8 mol/L, 0.069 g/L Li, 0.589 g/L Co, 0.587 g/L Ni, 0.549 g/L Mn, and 0.479 g/L Ti; V<sub>aq</sub>/V<sub>HDES</sub> = 1/1; mixing time was 30 min.</p>
Full article ">Figure 5
<p>Dependence of the degree of metal ion extraction on the molar fraction of D2EHPA. Aqueous phase: [HCl] = 8 mol/L, 0.069 g/L Li, 0.589 g/L Co, 0.587 g/L Ni, 0.549 g/L Mn, and 0.479 g/L Ti. HDES phase: χ<sub>TOPO</sub> = 4; V<sub>aq</sub>/V<sub>HDES</sub> = 1/1; mixing time was 30 min.</p>
Full article ">Figure 6
<p>Dependence of the degree of extraction of metal ions on the phase contact time. HDES phase: TOPO/D2EHPA/menthol = 4:1:5. Aqueous phase: [HCl] = 6 mol/L, 0.069 g/L Li, 0.589 g/L Co, 0.587 g/L Ni, 0.549 g/L Mn, and 0.479 g/L Ti; V<sub>aq</sub>/V<sub>HDES</sub> = 1/1.</p>
Full article ">Figure 7
<p>The dependence of the degree of extraction of metal ions on the concentration of HCl. HDES phase: TOPO/D2EHPA/menthol = 4:1:5. Aqueous phase: 0.069 g/L Li, 0.589 g/L Co, 0.587 g/L Ni, 0.549 g/L Mn, and 0.479 g/L Ti; V<sub>aq</sub>/V<sub>HDES</sub> = 1/1; mixing time was 30 min.</p>
Full article ">Figure 8
<p>Dependence of the degree of extraction of metal ions on the volume ratio of the aqueous and organic phases. HDES phase: TOPO/D2EHPA/menthol = 4:1:5. Aqueous phase: 0.069 g/L Li, 0.589 g/L Co, 0.587 g/L Ni, 0.549 g/L Mn, and 0.479 g/L Ti; mixing time was 30 min.</p>
Full article ">Figure 9
<p>Isotherms of the Me extraction with TOPO/D2EHPA/menthol = 4:1:5. Aqueous phase: [HCl] = 1, 6, 9 mol/L for Ti(IV), Co(II), and Mn(II) ions, respectively; V<sub>aq</sub>/V<sub>HDES</sub> = 1/1; mixing time was 30 min.</p>
Full article ">Figure 10
<p>Dependence of the degree of stripping of Ti(IV), Co(II), and Mn(II) ions on the phase contact time. Extraction conditions for Ti(IV), Co(II), and Mn(II): [HCl] = 1, 6, and 9 mol/L, respectively. Stripping conditions for Ti(IV): [H<sub>3</sub>PO<sub>4</sub>] = 3 mol/L, 3 vol.% H<sub>2</sub>O<sub>2</sub>; for Co(II) and Mn(II): [HCl] = 1 mol/L.</p>
Full article ">Figure 11
<p>The possibility of the reuse of TOPO/D2EHPA/menthol 4:1:5. Conditions for the stripping of Ti(IV): [H<sub>3</sub>PO<sub>4</sub>] = 3 mol/L, 3 vol.% H<sub>2</sub>O<sub>2</sub>; Co(II) and Mn(II) ions: [HCl] = 1 mol/L.</p>
Full article ">Figure 12
<p>Dependence of the degree of extraction of metals from HCl 1 mol/L solution (<b>a</b>) and Co(II) and Mn(II) from HCl 6 mol/L solution (<b>b</b>) on the volume flow rate of the phases.</p>
Full article ">
3 pages, 491 KiB  
Correction
Correction: Wiechers et al. Development of a Process for Direct Recycling of Negative Electrode Scrap from Lithium-Ion Battery Production on a Technical Scale and Its Influence on the Material Quality. Batteries 2024, 10, 218
by Patrick Wiechers, Anna Hermann, Sofia Koob, Fabian Glaum and Marco Gleiß
Batteries 2024, 10(12), 441; https://doi.org/10.3390/batteries10120441 - 12 Dec 2024
Viewed by 309
Abstract
In the original publication [...] Full article
Show Figures

Figure 9

Figure 9
<p>Adhesive force of the non-calendered negative electrodes with recycled coating, as well as the references with pristine coatings.</p>
Full article ">
23 pages, 2200 KiB  
Review
Recent Advancements in Artificial Intelligence in Battery Recycling
by Subin Antony Jose, Connor Andrew Dennis Cook, Joseph Palacios, Hyundeok Seo, Christian Eduardo Torres Ramirez, Jinhong Wu and Pradeep L. Menezes
Batteries 2024, 10(12), 440; https://doi.org/10.3390/batteries10120440 - 11 Dec 2024
Viewed by 642
Abstract
Battery recycling has become increasingly crucial in mitigating environmental pollution and conserving valuable resources. As demand for battery-powered devices rises across industries like automotive, electronics, and renewable energy, efficient recycling is essential. Traditional recycling methods, often reliant on manual labor, suffer from inefficiencies [...] Read more.
Battery recycling has become increasingly crucial in mitigating environmental pollution and conserving valuable resources. As demand for battery-powered devices rises across industries like automotive, electronics, and renewable energy, efficient recycling is essential. Traditional recycling methods, often reliant on manual labor, suffer from inefficiencies and environmental harm. However, recent artificial intelligence (AI) advancements offer promising solutions to these challenges. This paper reviews the latest developments in AI applications for battery recycling, focusing on methodologies, challenges, and future directions. AI technologies, particularly machine learning and deep learning models, are revolutionizing battery sorting, classification, and disassembly processes. AI-powered systems enhance efficiency by automating tasks such as battery identification, material characterization, and robotic disassembly, reducing human error and occupational hazards. Additionally, integrating AI with advanced sensing technologies like computer vision, spectroscopy, and X-ray imaging allows for precise material characterization and real-time monitoring, optimizing recycling strategies and material recovery rates. Despite these advancements, data quality, scalability, and regulatory compliance must be addressed to realize AI’s full potential in battery recycling. Collaborative efforts across interdisciplinary domains are essential to develop robust, scalable AI-driven recycling solutions, paving the way for a sustainable, circular economy in battery materials. Full article
(This article belongs to the Special Issue Towards a Smarter Battery Management System: 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>LIB recycling market size 2023 to 2033 (USD billion). Adapted from [<a href="#B8-batteries-10-00440" class="html-bibr">8</a>].</p>
Full article ">Figure 2
<p>Broad applications of computer vision across multiple sectors. Reproduced with permission from [<a href="#B23-batteries-10-00440" class="html-bibr">23</a>].</p>
Full article ">Figure 3
<p>Workflow of a digital twin system for comprehensive battery lifecycle management. Reproduced with permission from [<a href="#B53-batteries-10-00440" class="html-bibr">53</a>].</p>
Full article ">
14 pages, 20689 KiB  
Article
Enhancing Lithium Recovery from Slag Through Dry Forced Triboelectric Separation: A Sustainable Recycling Approach
by Mehran Javadi, Cindytami Rachmawati, Annett Wollmann, Joao Weiss, Hugo Lucas, Robert Möckel, Bernd Friedrich, Urs Peuker and Alfred P. Weber
Minerals 2024, 14(12), 1254; https://doi.org/10.3390/min14121254 - 10 Dec 2024
Viewed by 506
Abstract
The increasing use of lithium-containing materials highlights the urgent need for their recycling to preserve resources and protect the environment. Lithium-containing slags, produced during the pyrometallurgical process in lithium-ion battery recycling, represent an essential resource for lithium recovery efforts. While multiple methods for [...] Read more.
The increasing use of lithium-containing materials highlights the urgent need for their recycling to preserve resources and protect the environment. Lithium-containing slags, produced during the pyrometallurgical process in lithium-ion battery recycling, represent an essential resource for lithium recovery efforts. While multiple methods for lithium recycling exist, it is crucial to emphasize environmentally sustainable approaches. This study employs dry forced triboelectrification (FTC) to recover valuable components from slag powder, commonly known as engineered artificial minerals (EnAMs). The FTC method is used to change the charge of the target material and achieve a neutral state while other materials remain charged. The downstream electrostatic separator enables the charged particles to be separated from the target material, which in this study is lithium aluminate. The results show that the method is effective, and lithium aluminate can be successfully enriched. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic of (<b>a</b>) liberation by comminution (crushing and grinding), (<b>b</b>) slag components liberated to various degrees of free surface area (FSA) (partially and fully surface-liberated target phases).</p>
Full article ">Figure 2
<p>Experimental setup: (<b>a</b>) combination of dry forced triboelectric charging (FTC) of powders and the Faraday cup electrometer (FCE) test to determine the PZNC of the used materials and (<b>b</b>) the electrostatic separator, including five collection bins at the bottom.</p>
Full article ">Figure 3
<p>Slag’s transformation from (<b>a</b>) solid block to (<b>b</b>) powder form and (<b>c</b>) mineral characterization using an electron probe micro-analyzer (EPMA).</p>
Full article ">Figure 4
<p>(<b>a</b>) Particle size distribution of Li-Slag obtained by MLA and (<b>b</b>) X-ray diffraction (XRD) pattern of Li-Slag and single-phase components [<a href="#B8-minerals-14-01254" class="html-bibr">8</a>].</p>
Full article ">Figure 5
<p>Characterization of the liberation of the lithium aluminate component in the Li-Slag: (<b>a</b>) examples of particles extracted from EPMA analysis (from <a href="#minerals-14-01254-f003" class="html-fig">Figure 3</a>c, lithium aluminate in pink color) exhibiting different degrees of FSA, (<b>b</b>) lithium aluminate percentage in each liberation class by free surface area, and (<b>c</b>) particle size classes in the 90%–100% liberation class by free surface area.</p>
Full article ">Figure 6
<p>The specific charge of powders as a function of the applied voltage in the range of −12 to +12 kV for pure lithium aluminate, Li-Slag, and their mixture (10%/90%). From the fits of the two powders, the charging behavior of the mixture was calculated as a weighted average (black dashed line), which agrees very well with the measurements (green points).</p>
Full article ">Figure 7
<p>Free fall test of (<b>a</b>) Li-Slag and (<b>b</b>) lithium aluminate powders, demonstrating particle settling behavior through the separator without applied voltages (V<sub>chute</sub> = 0 kV, V<sub>sep</sub> = 0 kV).</p>
Full article ">Figure 8
<p>Electrostatic separation of original Li-Slag (<b>a</b>) at −3 kV (PZNC of Li-Slag) and (<b>b</b>) at −10.3 kV (PZNC of lithium aluminate). The numbers in the figure give the mass fraction of lithium slag (red) and of lithium aluminate (blue) (the sum of a mass fraction over all bins accounts for 100%).</p>
Full article ">Figure 9
<p>Electrostatic separation of the mixture of Li-Slag and lithium aluminate (<b>a</b>) at −3 kV (PZNC of Li-Slag) and (<b>b</b>) at −10.3 kV (PZNC of lithium aluminate). The numbers in the figure show the mass fraction of lithium slag (red) and lithium aluminate (blue).</p>
Full article ">Figure A1
<p>Conductivity of the mixture of Li-Slag and lithium aluminate for adding different amounts of lithium aluminate.</p>
Full article ">
15 pages, 4366 KiB  
Article
Separation of Magnesium and Lithium Ions Utilizing Layer-by-Layer Polyelectrolyte Modification of Polyacrylonitrile Hollow Fiber Porous Membranes
by Danai Koukoufilippou, Ioannis L. Liakos, George I. Pilatos, Niki Plakantonaki, Alexandros Banis and Nikolaos K. Kanellopoulos
Materials 2024, 17(23), 5878; https://doi.org/10.3390/ma17235878 - 30 Nov 2024
Viewed by 520
Abstract
This study explores the layer-by-layer (LBL) modification of polyacrylonitrile (PAN) hollow fibers for effective Mg2+/Li+ separation. It employs an LBL method of surface modification using polyelectrolytes, specifically aiming to enhance ion selectivity and improve the efficiency of lithium extraction from [...] Read more.
This study explores the layer-by-layer (LBL) modification of polyacrylonitrile (PAN) hollow fibers for effective Mg2+/Li+ separation. It employs an LBL method of surface modification using polyelectrolytes, specifically aiming to enhance ion selectivity and improve the efficiency of lithium extraction from brines or lithium battery wastes, which is critical for battery recycling and other industrial applications. The modification process involves coating the hydrolyzed PAN fibers with alternating layers of positively charged polyelectrolytes, such as poly(allylamine hydrochloride) (PAH), polyethyleneimine (PEI), or poly(diallyldimethylammonium chloride) (PDADMAC) and negatively charged polyelectrolytes, such as poly(styrene sulfonate) (PSS), to form polyelectrolyte multilayers (PEMs). This study evaluates the modified membranes in Mg2+ and Li+ salt solutions, demonstrating significant improvements in selectivity for Mg2+/Li+ separation. PAH was identified as the optimal positively charged polyelectrolyte. PAN hollow fibers modified with ten bilayers of PAH/PSS achieved rejection rates of 95.4% for Mg2+ ions and 34.8% for Li+ ions, and a permeance of 0.39 LMH/bar. This highlights the potential of LBL techniques for effectively addressing the challenges of ion separation across a variety of applications. Full article
(This article belongs to the Section Porous Materials)
Show Figures

Figure 1

Figure 1
<p>Illustration of LBL deposition on PAN hollow fibers and the chemical structures of polyelectrolytes and GA (light red and purple line layers represents positively and negatively charged polyelectrolytes respectively; double arrow green lines represent the cross-linker).</p>
Full article ">Figure 2
<p>Experimental set-up for testing the metal ion concentration through the membrane (permeance).</p>
Full article ">Figure 3
<p>SEM images of PAN hollow fiber substrates: (<b>a</b>) unmodified substrate, exhibiting an average pore size of 200 nm (±50 nm); (<b>b</b>) modified substrates after a 2 h pretreatment with 2M NaOH; (<b>c</b>,<b>d</b>) modified with 2 h pretreatment with NaOH 2M and subsequent deposition of five bilayers of (PAH/PSS); (<b>e</b>,<b>f</b>) modified substrates after a 2 h pretreatment with NaOH 2M and subsequent deposition of 10 bilayers (PAH/PSS).</p>
Full article ">Figure 4
<p>SEM images of PAN hollow fiber substrates: (<b>a</b>,<b>b</b>) unmodified substrate, exhibiting an average pore size of 200 nm (±50 nm); (<b>c</b>) modified with 30 min pretreatment with NaOH 2M and subsequent deposition of five bilayers of (PAH/PSS); (<b>d</b>) modified substrates after a 30 min pretreatment with NaOH 2M and subsequent deposition of 10 bilayers (PAH/PSS).</p>
Full article ">Figure 5
<p>FTIR spectra of PAN hollow fiber with different pretreatment and (PAH/PSS) modification (<b>a</b>) PAN hollow fiber without pretreatment or modification, with 30 min and 2 h 2M NaOH pretreatment; (<b>b</b>–<b>d</b>) PAN hollow fiber without pretreatment or modification, with 2 h 2M NaOH pretreatment and 10 bilayer (PAH/PSS) deposition.</p>
Full article ">Figure 6
<p>(<b>a</b>) Mg<sup>2+</sup> rejection and Li<sup>+</sup> rejection of all membranes with (PAH/PSS) deposition (<b>b</b>) Mg<sup>2+</sup> rejection, Li<sup>+</sup> rejection and permeance of PAN hollow fibers with 1.5–2.5 h time of hydrolysis and 5, 10 bilayers of (PAH/PSS) (<b>c</b>) Mg<sup>2+</sup> rejection, Li<sup>+</sup> rejection and permeance of (PAH/PSS)<sub>10</sub>-2 membrane (<b>d</b>) Mg<sup>2+</sup> rejection, Li<sup>+</sup> rejection and permeance of (PAH/PSS)<sub>10</sub>-2.5 membrane.</p>
Full article ">Figure 7
<p>(<b>a</b>) Comparison of the performance of PAH/PSS, PDADMAC/PSS and PEI/PSS LBL membranes. (<b>b</b>) comparison of the selectivity coefficient and permeance of PAH/PSS, PDADMAC/PSS and PEI/PSS LBL membranes.</p>
Full article ">
15 pages, 3474 KiB  
Article
Selective Leaching for the Recycling of Lithium, Iron, and Phosphorous from Lithium-Ion Battery Cathodes’ Production Scraps
by Martina Bruno, Carlotta Francia and Silvia Fiore
Batteries 2024, 10(12), 415; https://doi.org/10.3390/batteries10120415 - 27 Nov 2024
Viewed by 629
Abstract
The market for lithium iron phosphate (LFP) batteries is projected to grow in the near future. However, recycling methods targeting LFP batteries, especially production scraps, are still underdeveloped. This study investigated the extraction of iron phosphate and lithium from LFP production scraps using [...] Read more.
The market for lithium iron phosphate (LFP) batteries is projected to grow in the near future. However, recycling methods targeting LFP batteries, especially production scraps, are still underdeveloped. This study investigated the extraction of iron phosphate and lithium from LFP production scraps using selective leaching, considering technical and economic aspects. Two leaching agents, sulfuric acid (0.25–0.5 M, 25 °C, 1 h, 50 g/L) and citric acid (0.25–0.5 M, 25 °C, 1 h, 70 g/L) were compared; hydrogen peroxide (3–6%vv.) was added to prevent iron and phosphorous solubilization. Sulfuric acid leached up to 98% of Li and recovered up to 98% of Fe and P in the solid residues. Citric acid leached 18–26% of Li and recovered 98% of Fe and P. Totally, 28% of Li was precipitated for sulfuric acid process, while recovery with citric acid did not produce enough precipitate for a characterization. Sulfur is the main impurity present in the precipitates. The total operative costs associated with reagents and energy consumption of the sulfuric acid route were below 3.00 €/kg. In conclusion, selective leaching provided a viable and economic method to recycle LFP production scraps, and it is worth further research to optimize Lithium recovery. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Outline of the experimental procedure to recycle Fe, Li, and P from LFP production scraps.</p>
Full article ">Figure 2
<p>XRD spectrum of LFP cathodes’ production scraps.</p>
Full article ">Figure 3
<p>Leaching efficiency (%) of lithium, iron, and phosphorous achieved in the selective leaching tests with sulfuric and citric acids.</p>
Full article ">Figure 4
<p>Comparison between XRD spectra of (<b>A</b>) initial sample of LFP cathodes’ production scraps, (<b>B</b>) solid residues after leaching with sulfuric acid and hydrogen peroxide, and (<b>C</b>) citric acid and hydrogen peroxide.</p>
Full article ">Figure 5
<p>FESEM images of (<b>A</b>) initial LFP powder, (<b>B</b>) residues after sulfuric acid leaching, and (<b>C</b>) after citric acid leaching.</p>
Full article ">Figure 6
<p>XRD spectrum of precipitate obtained at pH 12 from selective leaching with sulfuric acid and hydrogen peroxide.</p>
Full article ">Figure 7
<p>Operative costs of the considered routes applied to 1 kg of LFP cathodic material.</p>
Full article ">
22 pages, 2134 KiB  
Article
Unlocking Economic and Environmental Gains Through Lithium-Ion Battery Recycling for Electric Vehicles
by Bianca Ifeoma Chigbu and Ikechukwu Umejesi
Resources 2024, 13(12), 163; https://doi.org/10.3390/resources13120163 - 21 Nov 2024
Viewed by 829
Abstract
Amid South Africa’s shift towards electric vehicles (EVs), building a lithium-ion battery (LIB) recycling sector is essential for promoting sustainable development and generating employment opportunities. This study employs qualitative methodologies to collect insights from 12 critical stakeholders in the automotive, mining, and recycling [...] Read more.
Amid South Africa’s shift towards electric vehicles (EVs), building a lithium-ion battery (LIB) recycling sector is essential for promoting sustainable development and generating employment opportunities. This study employs qualitative methodologies to collect insights from 12 critical stakeholders in the automotive, mining, and recycling sectors and academia to examine the feasibility and advantages of establishing such an industry. We implemented purposeful and snowball sampling to guarantee an exhaustive array of viewpoints. Thematic analysis of the interview data reveals that LIB recycling has substantial social, environmental, and economic implications. The results emphasize the pressing necessity of recycling infrastructure to mitigate environmental impacts and attract investment. The economic feasibility and employment potential of LIB recycling is promising despite the early stage of the EV industry in South Africa. These potentials are influenced by EV adoption rates, technological advancements, regulatory frameworks, and industry growth. In this sector, employment opportunities are available in various phases: battery collection, transportation, disassembly, testing, mechanical crushing, hydrometallurgical processes, valuable metal recovery, manufacturing, reuse, research and development, and administrative roles. Each of these roles necessitates a unique set of skills. This interdisciplinary research investigates vital elements of economic growth, employment creation, environmental sustainability, policymaking, technological innovation, and global collaboration. The study offers valuable guidance to policymakers and industry stakeholders trying to establish a sustainable and robust LIB recycling industry in South Africa by utilizing Transition Management Theory to develop a framework for improving the sustainability and circularity of the EV LIB recycling sector. Full article
Show Figures

Figure 1

Figure 1
<p>Flowchart of the EV LIB Recycling Process.</p>
Full article ">Figure 2
<p>Theoretical Foundations for LIB Recycling: A Sustainable Approach.</p>
Full article ">Figure 3
<p>LIB recycling landscape.</p>
Full article ">Figure 4
<p>Employment creation in LIB recycling sector.</p>
Full article ">Figure 5
<p>Practical insights and beneficiaries.</p>
Full article ">
64 pages, 12203 KiB  
Review
Beyond Lithium: Future Battery Technologies for Sustainable Energy Storage
by Alan K. X. Tan and Shiladitya Paul
Energies 2024, 17(22), 5768; https://doi.org/10.3390/en17225768 - 18 Nov 2024
Viewed by 1318
Abstract
Known for their high energy density, lithium-ion batteries have become ubiquitous in today’s technology landscape. However, they face critical challenges in terms of safety, availability, and sustainability. With the increasing global demand for energy, there is a growing need for alternative, efficient, and [...] Read more.
Known for their high energy density, lithium-ion batteries have become ubiquitous in today’s technology landscape. However, they face critical challenges in terms of safety, availability, and sustainability. With the increasing global demand for energy, there is a growing need for alternative, efficient, and sustainable energy storage solutions. This is driving research into non-lithium battery systems. This paper presents a comprehensive literature review on recent advancements in non-lithium battery technologies, specifically sodium-ion, potassium-ion, magnesium-ion, aluminium-ion, zinc-ion, and calcium-ion batteries. By consulting recent peer-reviewed articles and reviews, we examine the key electrochemical properties and underlying chemistry of each battery system. Additionally, we evaluate their safety considerations, environmental sustainability, and recyclability. The reviewed literature highlights the promising potential of non-lithium batteries to address the limitations of lithium-ion batteries, likely to facilitate sustainable and scalable energy storage solutions across diverse applications. Full article
(This article belongs to the Section D: Energy Storage and Application)
Show Figures

Figure 1

Figure 1
<p>Chart showing the number of papers highlighting non-lithium-ion batteries across the years. Search results were obtained from Web of Science, searching titles with the keywords “X-ion AND batteries”, where X = Na, K, Mg, Al, Zn, Ca.</p>
Full article ">Figure 2
<p>Flowchart outlining the methodology employed in this literature review. Search results are accurate as of September 2024.</p>
Full article ">Figure 3
<p>Schematics showcasing the working principle behind the different types of batteries: (<b>a</b>) the metal-ion battery; (<b>b</b>) the metal–air battery (MAB) in an aqueous, alkaline electrolyte.</p>
Full article ">Figure 4
<p>The periodic table with elements of interest as the active species shown. Lithium, the current state-of-the-art, is highlighted in green. Values of the standard electrode potential for the reaction <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>M</mi> </mrow> <mrow> <mi>n</mi> <mo>+</mo> </mrow> </msup> <mo>+</mo> <mi>n</mi> <msup> <mrow> <mi>e</mi> </mrow> <mrow> <mo>−</mo> </mrow> </msup> <mo>→</mo> <mo> </mo> <mi>M</mi> </mrow> </semantics></math>, with reference to the standard hydrogen electrode (SHE), are shown below each element and obtained from [<a href="#B20-energies-17-05768" class="html-bibr">20</a>].</p>
Full article ">Figure 5
<p>Structural model for the O3- and P2-type layered transition metal oxides. The blue and green spheres represent, respectively, the transition metals and sodium ions. The different oxygen stacking sequences (each distinct layer is denoted by a letter “A”, “B”, or “C”) generate either octahedral or trigonal prismatic Na coordination, accordingly. Reproduced from Ref. [<a href="#B64-energies-17-05768" class="html-bibr">64</a>] with permission from the Wiley Online Library.</p>
Full article ">Figure 6
<p>Schematic showing the three-dimensional (3D) sodium-ion transport channels in Na<sub>3</sub>V<sub>2</sub>(PO<sub>4</sub>)<sub>3</sub>, along the (<b>a</b>) x, (<b>b</b>) y, and (<b>c</b>) curved z directions. The purple tetrahedron represents a PO<sub>4</sub> unit while the green octahedron represents a VO<sub>6</sub> interlink. In the middle of each octahedron is one Na<sup>+</sup> ion. Reproduced from Ref. [<a href="#B79-energies-17-05768" class="html-bibr">79</a>] with permission from the Royal Society of Chemistry.</p>
Full article ">Figure 7
<p>Schematic showing the Jahn–Teller distortion for high-spin d<sup>4</sup> species in octahedral sites. The green spheres represent high-spin Mn<sup>3+</sup> or Cr<sup>2+</sup> (with the d<sup>4</sup> electronic configuration) in a six-fold octahedral coordination with O<sup>2−</sup> ions in blue.</p>
Full article ">Figure 8
<p>Two innovative electrolytes to improve the electrochemical performance of PIBs: (<b>a</b>) schematic of the surface-modification strategy for mitigating structural instability due to Mn dissolution in the KMnF electrode. Reproduced from Ref. [<a href="#B99-energies-17-05768" class="html-bibr">99</a>] with permission from Springer Nature. (<b>b</b>) Schematic illustrations of the solvation structure and SEI formation on graphite in 1:8 (KFSI:TMP) electrolyte (i) and 3:8 (KFSI:TMP) electrolyte (ii), respectively. (iii) the schematic illustration shows a pre-cycled graphite electrode with an F-rich interface cycling in 1:8 (KFSI:TMP) electrolyte. (The orange and red symbols represent the TMP solvent molecule and the FSI− anion, respectively). Reproduced from Ref. [<a href="#B103-energies-17-05768" class="html-bibr">103</a>] with permission from the Wiley Online Library.</p>
Full article ">Figure 9
<p>Schematic for the proposed electrochemical reaction mechanism of PTCDA in KIBs during the discharge/charge process. Reprinted from Ref. [<a href="#B110-energies-17-05768" class="html-bibr">110</a>], copyright (2024) Elsevier, with permission from Elsevier.</p>
Full article ">Figure 10
<p>Heteroatom doping of graphene layers to promote long-term cycling stability of graphitic materials: (<b>a</b>) the structure of N-doped carbon nanosheets with pyridinic N (N-5), pyrrolic N (N-6), and graphitic N (N-Q) defects. Reproduced from Ref. [<a href="#B114-energies-17-05768" class="html-bibr">114</a>] with permission from the Wiley Online Library. (<b>b</b>) The structure of N/S dual-doped carbon. Reproduced from Ref. [<a href="#B118-energies-17-05768" class="html-bibr">118</a>] with permission from the Wiley Online Library.</p>
Full article ">Figure 11
<p>Pourbaix diagram of a Zn/H<sub>2</sub>O system at 25 °C. Reprinted from Ref. [<a href="#B170-energies-17-05768" class="html-bibr">170</a>], copyright (2024) Elsevier, with permission from Elsevier.</p>
Full article ">Figure 12
<p>Schematics highlighting the intricate balance needed in the degree of hydrophilicity of an electrode material: (<b>a</b>) some degree of hydrophobicity aids in the de-solvation of Zn<sup>2+</sup> ions, facilitating the deposition of zinc metal without side reactions, while (<b>b</b>) too much hydrophobicity would lead to sluggish electrochemical kinetics.</p>
Full article ">Figure 13
<p>Schematic showing the benefits of functionalised separators in AZIBs: (<b>a</b>) control of the crystallographic orientation of the Zn deposits to prevent dendrite formation, and (<b>b</b>) the de-solvation of Zn<sup>2+</sup> ions when passing through the separator.</p>
Full article ">Figure 14
<p>AFM images of (<b>a</b>) the pristine Zn electrode and the Zn electrodes after electrochemical cycling (<b>b</b>) with the GF separator and (<b>c</b>) with the CF separator. Top-view SEM images of the cycled Zn electrodes after electrochemical cycling (<b>d</b>,<b>e</b>) with the GF separator and (<b>f</b>,<b>g</b>) with the CF separator. Adapted from Ref. [<a href="#B167-energies-17-05768" class="html-bibr">167</a>], copyright (2024) Elsevier, with permission from Elsevier.</p>
Full article ">Figure 15
<p>Overview of the charging and discharging mechanisms of aluminium-ion batteries using the AlCl<sub>3</sub>/[EmIm]Cl electrolyte system. Instead of the usual rocking-chair mechanism where Al<sup>3+</sup> is shuttled between the cathode and anode, the active species Al<sub>2</sub>Cl<sub>7</sub><sup>−</sup> is reduced to Al (deposited at the anode) and AlCl<sub>4</sub><sup>−</sup> (deposited at the cathode) during charging.</p>
Full article ">Figure 16
<p>The electrochemical storage of the divalent AlCl<sup>2+</sup> ion with the synthesised tetradiketone macrocycle. Adapted and reproduced from Ref. [<a href="#B183-energies-17-05768" class="html-bibr">183</a>] under the CC-BY-4.0 license (<a href="https://creativecommons.org/licenses/by/4.0/" target="_blank">https://creativecommons.org/licenses/by/4.0/</a>).</p>
Full article ">Figure 17
<p>Schematic of the proposed lattice change during cycling. Three forces were identified by the authors during the intercalation process: electrostatic repulsion between Al<sup>3+</sup> and the PBA (F1), bond cooperation between metal ions and -CN- groups (F2), and homo-ionic repulsion between M<sup>2+</sup> and M<sup>2+</sup> ions (F3). Reproduced from Ref. [<a href="#B203-energies-17-05768" class="html-bibr">203</a>] with permission from the Wiley Online Library.</p>
Full article ">Figure 18
<p>(<b>a</b>) Crystal structure of the Chevrel phase with cavities for 3-dimensional ion transport. Reprinted with permission from Ref. [<a href="#B235-energies-17-05768" class="html-bibr">235</a>]. Copyright 2019 American Chemical Society. (<b>b</b>) Crystal structure of the spinel phase of general formula AB<sub>2</sub>X<sub>4</sub> (red: A ions; centre of blue octahedra: B ions; corners of blue octahedra: X ions). Adapted from Ref. [<a href="#B236-energies-17-05768" class="html-bibr">236</a>], copyright (2024) Elsevier, with permission from Elsevier.</p>
Full article ">Figure 19
<p>Summary of MIB anodes displaying different Mg-ion storage mechanisms. (<b>a</b>) Schematics of the electrochemical behaviour of bare Mg and Mg–M@Mg (M = tin and bismuth) anodes. Reproduced from Ref. [<a href="#B225-energies-17-05768" class="html-bibr">225</a>] with permission from the Royal Society of Chemistry. (<b>b</b>) Schematic illustration of the Mg-ion full-cell using an intercalation-type anode. Reproduced from Ref. [<a href="#B220-energies-17-05768" class="html-bibr">220</a>] with permission from Springer Nature. (<b>c</b>) Schematic illustration of the working mechanism of the magnesium dual-ion battery (Mg-DIB) based on the PTCDI anode and EG cathode. Reprinted from Ref. [<a href="#B215-energies-17-05768" class="html-bibr">215</a>], copyright (2024) Elsevier, with permission from Elsevier.</p>
Full article ">Figure 20
<p>(<b>a</b>) Schematic illustration of the urea-dominant solvation shell in the absence of acetamide (left) and the competitive solvation shell with acetamide (right). Adapted with permission from Ref. [<a href="#B226-energies-17-05768" class="html-bibr">226</a>]. Copyright 2024 American Chemical Society. (<b>b</b>) Schematic illustrations of the reactions occurring on the interface of the Mn-NVO electrode when assembled with the 1M MgCl<sub>2</sub>/H<sub>2</sub>O and optimised MgCl<sub>2</sub>·6H<sub>2</sub>O: acetamide: urea electrolyte (1:1:7). Adapted with permission from Ref. [<a href="#B226-energies-17-05768" class="html-bibr">226</a>]. Copyright 2024 American Chemical Society.</p>
Full article ">Figure 21
<p>Shielding effect of Mg<sup>2+</sup> in the hydrated V<sub>2</sub>O<sub>5</sub>·<span class="html-italic">n</span>H<sub>2</sub>O. The strong polarisation of the divalent Mg<sup>2+</sup> could be significantly reduced by solvating with water of crystallisation. Reprinted from Ref. [<a href="#B265-energies-17-05768" class="html-bibr">265</a>], copyright (2024) Elsevier, with permission from Elsevier.</p>
Full article ">Figure 22
<p>A conceptual illustration of the evolution of the DS-V<sub>2</sub>O<sub>5</sub> structure upon cycling. Reproduced from Ref. [<a href="#B262-energies-17-05768" class="html-bibr">262</a>] with permission from the Wiley Online Library.</p>
Full article ">Figure 23
<p>(<b>a</b>) The mechanism diagram of one-layer 2COF-18Ca_1. Reprinted with permission from Ref. [<a href="#B254-energies-17-05768" class="html-bibr">254</a>]. Copyright 2023 American Chemical Society. (<b>b</b>) Structural evolution of PTHAT-COF repeat unit during the discharge process. Reproduced from Ref. [<a href="#B256-energies-17-05768" class="html-bibr">256</a>] with permission from the Wiley Online Library.</p>
Full article ">Figure 24
<p>Computed electrochemical conversion path of Se electrode for CIBs. Reproduced from Ref. [<a href="#B253-energies-17-05768" class="html-bibr">253</a>] with permission from the Wiley Online Library.</p>
Full article ">Figure 25
<p>Box-and-whisker plots of initial capacity for full cells in this review at a current density of (<b>a</b>) 100 mA g<sup>−1</sup>; (<b>b</b>) 1000 mA g<sup>−1</sup>. Box-and-whisker plots of cycle life for full cells in this review at a current density of (<b>c</b>) 100 mA g<sup>−1</sup>; (<b>d</b>) 1000 mA g<sup>−1</sup>. All plots were generated using an inclusive median.</p>
Full article ">Figure 26
<p>Load peak shaving by battery energy storage system. Reproduced from Ref. [<a href="#B274-energies-17-05768" class="html-bibr">274</a>], copyright (2014) Elsevier, with permission from Elsevier.</p>
Full article ">Figure 27
<p>CO<sub>2</sub> emissions of LIB production, manufacturing, operation, and recycling. Emissions for the operation (or charging) were estimated using the Greenhouse Gas Emissions from Electric and Plug-In Hybrid Vehicles—Results from the US Department of Energy and converted assuming a 40 kWh battery has a range of 226 miles. Reprinted from Ref. [<a href="#B278-energies-17-05768" class="html-bibr">278</a>], copyright (2024) Elsevier, with permission from Elsevier.</p>
Full article ">Figure 28
<p>A schematic highlighting the potential solutions (in green) for a circular battery economy. Free-for-use icons made by Freepik, Konkapp, Smashicons from <a href="http://www.flaticon.com" target="_blank">www.flaticon.com</a>, accessed on 12 September 2024.</p>
Full article ">Figure 29
<p>Comparison of the characteristic temperatures of the three types of batteries. T<sub>onset</sub> represents the temperature where self-heating occurs; T<sub>open</sub> represents the temperature at which the battery’s safety valve is opened to release pressure; T<sub>sc</sub> represents the temperature at which the separator collapses, resulting in internal short-circuiting; T<sub>max</sub> represents the maximum temperature during thermal runaway. Reprinted from Ref. [<a href="#B294-energies-17-05768" class="html-bibr">294</a>], copyright (2024) Elsevier, with permission from Elsevier.</p>
Full article ">Figure 30
<p>The flash points of commonly used liquid electrolytes in this review, sorted in ascending order; the electrolytes are thus sorted in terms of increasing safety. Note that water and solid-state electrolytes do not have flash points. The data were obtained from Refs. [<a href="#B296-energies-17-05768" class="html-bibr">296</a>,<a href="#B297-energies-17-05768" class="html-bibr">297</a>,<a href="#B298-energies-17-05768" class="html-bibr">298</a>].</p>
Full article ">
Back to TopTop