Next Article in Journal
The Study of Structural Dynamic Response of Wind Turbine Blades under Different Inflow Conditions for the Novel Variable-Pitch Wind Turbine
Next Article in Special Issue
Impact of Data Corruption and Operating Temperature on Performance of Model-Based SoC Estimation
Previous Article in Journal
A Rule-Based Energy Management Technique Considering Altitude Energy for a Mini UAV with a Hybrid Power System Consisting of Battery and Solar Cell
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Direct Sonochemical Leaching of Li, Co, Ni, and Mn from Mixed Li-Ion Batteries with Organic Acids

1
Faculty of Materials Engineering, Silesian University of Technology, 2A Akademicka Street, 44-100 Gliwice, Poland
2
Faculty of Production Engineering and Materials Technology, Czestochowa University of Technology, 69 Dabrowskiego Street, 42-201 Czestochowa, Poland
*
Author to whom correspondence should be addressed.
Energies 2024, 17(16), 4055; https://doi.org/10.3390/en17164055
Submission received: 10 July 2024 / Revised: 25 July 2024 / Accepted: 27 July 2024 / Published: 15 August 2024

Abstract

:
Metals such as nickel, cobalt, lithium, and manganese are widely used in lithium-ion batteries (LIBs) in electronic devices and electric vehicles. It is forecast that there will be a strong increase in the number of electronic devices and electric vehicles in the coming years. (1) Background: In this paper, the application of ultrasound waves on improving Li, Co, Mn, and Ni leaching efficiency from mixed active cathode materials from different types of LIBs is presented. (2) Methods: Environmentally friendly, low-concentrated (0.75 M) organic acids (oxalic acid, citric acid) and, additionally, sulfuric acid, were used in sonochemical and chemical leaching (stirring process) at a temperature of 60 °C. (3) Results: The results showed significantly higher leaching efficiency of metals with ultrasound-assisted treatment, especially when using organic acids. An average of 50% better leaching results were obtained for Li in oxalic acid (99.6%) and for Co (93.1%) in citric acid during sonochemical leaching. (4) Conclusions: Based on the theory of hydrogen peroxide formation during ultrasound wave transition in solutions, the role of H2O2 as one of the most effective reductants used to enhance cobalt, manganese, and nickel leaching from LIBs is indicated.

1. Introduction

Lithium and cobalt belong to the group of Critical Raw Materials (CRMs) as essential elements for the development of strategic sectors such as renewable energy, electric mobility, defense, aerospace, and digital technologies [1]. These two metals, together with Ni and Mn, are an essential component of lithium-ion batteries (LIBs). Depending on the type of lithium batteries (LiMn2O4—LMO, LiCoO2—LCO, LiFePO4—LFP, LiNi0.33Mn0.33Co0.33O2—NMC, LiNi0.8Co0.15Al0.05O2—NCA), the metal content in active cathode material varies: Li: 3.8–7.8%; Co: 6.0–60.2%; Ni: 20.1–48.8%; and Mn: 5.6–60.7% [2,3,4]. Forecasts for the coming years undeniably confirm the growth tendency of the lithium cell market, electro-plasticity, and the demand for these metals. According to analyses [5] global Electric Vehicle (EV) production is set to increase from 3.4 million vehicles in 2020 to 12.7 million by 2024, and lithium demand will grow from 47.3 kilotons to 117.4 kt. It is estimated the EU will require up to 160 times more lithium and 15 times more cobalt in 2050 compared to levels of demand in 2018 to develop rapid low-carbon strategies [6]. The main thing that attracts attention is lithium, which the European Union recently added to its list of strategic minerals and for which it wants to establish more robust supply chains [7]. The usage, manufacturing rates, end-of-life handling, and recycling of lithium-ion batteries are expected to increase in the coming years as they play a crucial role in the EU’s transition to a climate-neutral economy. Today, lithium recovery in the EU is more limited and the volume of recovered and more economically valuable metals used in battery manufacturing is low. Only 12% of Al, 22% of Co, 8% of Mn, and 16% of Ni used within the EU is recycled [8]. Chemical, physical, and biological processes have been widely discussed [9,10,11,12] for the recovery of valuable metals from LIBs. The most popular approaches are hydrometallurgical methods, with strong inorganic acids such as HCl [13], HNO3 [14], Na2S2O8, and H2SO4 [3,15,16,17,18] commonly being used as leaching reagents. Traditional leaching treatments of LIBs rely on expensive reagents and can lead to the release of toxic gases, and the waste acid solution after leaching is a threat to the environment. Therefore, the most sustainable techniques with not-so-aggressive conditions, a closed loop of leaching agents, and improved environmental impact are desired. At the forefront are biological methods (microbial leaching) as important assets in future industrial implementation that are economically sound and have reduced environmental problems (reduced footprint from CO2 emissions, as well as reduced impact on groundwater and soil quality) [19,20]. An equally important and more environmentally friendly approach is the use of organic acids.
The advantages of this approach relate to biodegradability and non-toxic gas emissions [21,22], the involvement of protonation and chelation reactions that accelerate the dissolution of metals [23], and the reduction of the footprint of recycling processes [24]. Many organic acids have been used to improve the extraction rate of valuable metals from LIBs. Although the acidity (correlated with acidic strength) of organic acids is considerably lower than that of strong inorganic acids, their leaching efficiency is high. This is because organic acids can be used as moderate reducing agents for multivalent transition metals and to stabilize the dissolution of metal ions by forming chelating complexes [25,26]. The most popular is citric acid, usually applied together with a reducing agent in the form of H2O2 [27,28,29,30], or ascorbic acids [29]. Other acids, such as oxalic [31], malic [28], formic [32], succinic [33], or glutaric [34] acids, have been used as well. In most studies, the high metal leaching efficiency with organic acid (>95%), comparable to the use of conventional acids (H2SO4, HCl), has been demonstrated [28,30,31,32]. However, most experiments have been carried out over a temperature range of 80−100 °C [9,10], while only a few works report lower temperatures bringing equally good metal leaching results [27,32]. In addition to the commonly used factors intensifying or supporting LIB metal dissolution—such as temperature, the addition of a reducing agent, or speed of mixing, a highly enhancing factor—ultrasounds are still among the least used and tested methods in metal leaching. Due to its advantages, ultrasound technology has been widely used in many areas, such as medical diagnostic and therapy [35], micro- and nanomotors [36,37], and hydrometallurgy, where ultrasound is applied to enhance the extraction of metals from precious or rare ores ( e-waste e.g., PCB and LCD [38,39,40,41,42]) and releasing metals in a shorter time using acids at low concentration [43] during leaching and bioleaching [41,44,45,46,47]. Cavitation and bubbles created by ultrasound waves in the liquid cause high-speed micro-jets (300–1000 m/s) [48], and high local temperature creates localized material erosion, destroys the surface of the grain, enhances mass transfer, accelerates reaction rates, and increases metal recovery during leaching [38,39,42]. Ultrasounds can be introduced into the system by either directly immersing the ultrasound source in a washing medium (direct sonication) or by immersing a vessel containing a mixture of test material and leaching agent in a tank containing an ultrasonically treated liquid, which is usually water (indirect sonication) [43,44]. The available literature and laboratory-scale sonoleaching tests for Li and Co recovery under different sonic conditions are presented in Table 1. The results show that, in direct sonoleaching conditions, the leaching time is significantly reduced compared to that of the method with an ultrasonic cleaning bath/vessels using inorganic acids. The leaching of Co and Li was found to be more efficient with citric acid than with the two inorganic acids [27], remarkable results were achieved for citric acid, whereby nearly 100% of the Co, Li, and Ni were recovered in 35 minute process [45]. It should be noted that these studies used to a cathode material (LiCoO2) selected from one type of battery, limiting the analyzed composition to Li and Co. The current market and technological trend are related to the lowering use of expensive cobalt in the composition of the batteries in favor of a greater concentration of Ni and mixed metal cathode production [10,49]. This forces extending the battery recycling process by focusing not only on cobalt but, for more practical operation, also on other metals derived from cobalt-deficient cathodes and from a mixture of different cathode materials. In our study, we focused on combining the potential of high-efficiency metal recovery using organic acids (0.75 M oxalic and 0.75 M citric acids) with the application of ultrasound waves to shorten process time and improve metal’s leaching rate. A mixture of a variety of cathode active materials (containing Li, Co, Ni, Mn) was used and the extraction efficiency of these metals at the temperature of 60 °C was determined.

2. Materials and Methods

2.1. Materials

Li-ion cells from used cell phones (Samsung prismatic cells) and portable computer equipment (laptop) (cylindrical batteries, type 8858) were used in our research. Common cell compositions based on information in the Material Safety Data Sheet (MSDS) provided by the manufacturer are presented in Table 2.
The batteries were first manually dismantled. Next, cathode active material with lithium transition metal oxides (Co, Ni, Mn) on Al foil was separated from the rest of the components (steel cases, conductive carbon particles, polymer binder). Then Al collector foil with the active cathode material was cut into small pieces (1 × 1 cm) and was ground by a knife mill (ChemLand FW135, Stargard, Poland). Ground mixed materials were washed with H2O demi water, and left to dry for 24 h at 50 °C. Sieve analysis was used to divide the granular material into smaller size by applying an electromagnetic sieve shaker (Multiserw-Morek LPzE-2e, Marcyporeba, Poland) equipped with sieves of standard sizes: >2.0, 1.0, 0.5, 0.2, 0.1 and <0.1 mm. Separating the material into individual fractions allowed for a significant separation of the cathode active material from the shredded aluminum foil (Figure 1). Material with the finest fragmentation (<0.1 mm), characterized by the largest contact surface, was selected for further leaching tests. A particle size analysis of the initial samples and residues was performed by using a Malvern Mastersizer 3000 laser (Malvern, UK) particle size analyzer.

2.2. Methods—Leaching Procedure

Data in the literature reports that a concentration of organic acids in the range of 1–2 M is most commonly used in LIBs cell leaching experiments [9,27,50,51,52,53,54,55]. In the present study, lower acid concentrations were applied. i.e., 0.75 M oxalic acid (C2H2O4) and 0.75 M citric acid (C6H8O7), these concentrations were nonetheless sufficient to ensures significant acid reactivity, especially in the case of oxalic acid, which is one of the strongest carboxylic acids [56]. For comparison, 0.75 M H2SO4 was used as an example of inorganic acid, that has been widely used as a high-potential leaching agent. In our experiment, no additional reducing agent was used. Leching was carried out with a sample density of 50 g/L, at a temperature of 60 °C or 4 h. One leaching test with 0.75 M oxalic acid (C2H2O4) was performed additionally at room temperature (22 °C) and 40 °C for 4 h.
Experiments were conducted under two different conditions, i.e., —the mixed cathode material was subjected to acid leaching:
(i)
mechanical stirring leaching—in closed cylindrical reaction vessels with a thermostatic jacket and mechanical stirrer (300 rpm) (Lenz with Heidolph SN stirrer and thermostat Grant LT Ecocool 150 SN, Royston, UK),
(ii)
sonochemical leaching—in closed vessels with immersed titanium probe conductive ultrasound produced by an ultrasonic generator at 20 kHz frequency, and amplitude range 65% (SONOPhULS HD 4200, Bandelin, Berlin, Germany). Under the assumed operating parameters of the generator, the sonotrode generated a temperature of 60 °C in the solution, which was maintained as a constant value in the sonochemical leaching experiments.
Table 3 shows the conditions for the mixing and ultrasonic leaching experiments. To test the pH, a pH-Eh-meter (Hydromet CP-505, Gliwice, Poland) was used. The progress of leaching was monitored by taking 5 mL samples at different time intervals and analyzing the metal contents after filtration. For the analysis of the Ni, Co, Mn and separate Li concentrations, a Microwave Plasma Atomic Emission Spectrometer (Agilent MP-AES 4200, Santa Clara, CA, USA) and Atomic Absorption Spectrometer (ContrAA300, Analytik Jena, Germany) were used, respectively. Analysis of the morphology and structure of the ground samples and leaching residues were made with a Hitachi S-4200 scanning electron microscope (Naka, Japan) using secondary electron detectors (SE). Chemical composition tests were performed using an X-ray energy dispersion spectrometer (EDS) from Thermo Noran (System Seven) at a voltage accelerating the electron beam of 15 keV.
The metals leaching efficiency was calculated by the following equation:
E = C v · V m 0 · C 0 / 100 · 100 %
where E is leaching efficiency of metal (%); Cv is metal concentration in leachate (mg/L); V is volume of leach solution (L); m0 is the initial mass of sample (g); and C0 refers to metal content in leaching material (%).

3. Results and Discussion

3.1. Cathode Active Material Characterization

The scanning electron microscopy images (SEM) with microanalysis (EDS) samples of the active cathode material from mobile phones and laptops and mixed materials are shown in Figure 2.
The presence of cobalt was particularly noticeable, accounting for 56% of the ground material in the laptop battery, 98% of the material from the mobile phone, and 57% of the mixed material of batteries from both mobile phones and laptops. In the sample of material from a laptop, nickel and manganese were present, with mass fractions of 14% and 24%, respectively, these metals were not detected in the sample from the phone battery. These data verified the model composition of the cathode material of lithium cells used in electronic and electrical devices. In the case of mobile phones, the cathode mass was composed of lithium cobalt oxide (LiCoO2) (Figure 2a). For batteries used in laptops, the presence of additional Ni and Mn indicated that it was an NMC battery type, containing lithium-nickel-manganese-cobalt oxide (LiNi0.33Mn0.33Co0.33O2)—Figure 2b. The aluminum present in the SEM-EDS results is a remnant of aluminum foil from cathode material after the process of grinding. Due to the detection limit of the SEM-EDS, Li was not identified, but its content, together with those of other metals (Co, Ni, Mn), was determined by AAS and MP-AES methods, respectively. The content of metals in the mixed cathode active material of laptop and mobile phone batteries intended for leaching experiments is shown in Table 4.

3.2. Effect of Time on Leaching Efficiency

In our study, sulfuric acid and 2 organic acids were used: Citric acid recognized as excellent chelating agent, binding metals by making them soluble and the cheapest organic acid [29] and oxalic acid which forms water-soluble oxalate compounds and can be used simultaneously as reducing agent to reduce Co(III) to Co(II) and as effective precipitating agent [9,10,43,57]. Figure 3, Figure 4 and Figure 5 shows changes in the metals leaching efficiency (Li, Co, Ni and Mn) in inorganic and organic acids over a period of 4 h. Depending on the method (sonochemical or stirring leaching) and type of acid, different mobilization efficiencies and selectivities of the metals were observed.

3.2.1. Leaching with Oxalic Acid

Figure 3 confirms the almost complete mobilization of Li after 120 min, where 95.2% of Li was dissolved in oxalic acid (with ultrasound). Cobalt was released in parallel at 31.2%, while Ni and Mn did not dissolve significantly, remaining in the solid phase at 98%. At the same time, 50.3% of Li (with a final value 59.2%) was leached in the system with mechanical stirring (Figure 3b). Ni, Co and Mn were co-mobilized in percentages, not exceeding 1% for Mn and 4.5% for Ni and Co. A similar observation may be found in [28,31,58,59], i.e., when oxalic acid was used to leach transition metals from a cathodes material (Ni, Mn, Co) oxalates were formed that precipitated out of solution.
The oxalate chemistry explains the different behavior of the considered metals. In the presence of oxalic acid, Mn, Ni and Co form water-insoluble oxalate compounds such as MnC2O4·2H2O, NiC2O4·2H2O, and CoC2O4·2H2O [60], while Li goes into solution easily. Oxalic acid works as leaching and reducing agent simultaneously, releasing H+ and producing CO2 from predominant H2C2O4 species at a pH below 1.23 [9,61], as follows:
2LiCoO2(s) + 4H2C2O4(aq) → Li2C2O4(aq) + 2CoC2O4(s) + H2O + 2CO2(aq)
In many works [5,28,31,59] oxalic acid was applied as a precipitation agent for metals from LIBs with good results. When oxalic acid was used for LiCoO2 (LCO) cathode active materials, lithium was extracted easily in the leaching solution, while cobalt was precipitated as cobalt oxalate. Zeng et al. [31] reported that 98% of Li was leached into the aqueous phase, while 97% of Co was precipitated out as cobalt(II) oxalate (study conditions: 1 M oxalic acid, 2.5 h, 95 °C, 400 rpm, 15 g/L). Also, positive Li separation results were obtained for the mixed cathode material. Zhang et al. [58] achieved 84% Li recovery in the aqueous phase with a LiNixMnyCozO2 (NMC) cathodes material, while 98% of the Co, Ni, and Mn were precipitated (study conditions: 0.6 M oxalic acid, 2 h, 75 °C, 20 g/L). Our results also confirm the possibility of the selective recovery of Li from mixed cathode material, which is particularly important for practical separation techniques. During the first 60 min of sonochemical leaching, 94.1% Li recovery was achieved while Co stayed in the solid phase (Figure 3a). However, this time, the transition of cobalt onto the solution (31.2%) could be observed. The increased solubility of Co was related to the reduced high valence of Co to lower Co3+ to Co2+, which was related to the presence of ultrasound in the chemical system. Therefore reaction (2) describes (i) the behavior of Li and other metals in a mechanical stirring system (Figure 3b), where Co and (implicitly), Ni, and Mn also are in the solid phase and (ii) the first 60 min of sonochemical leaching, where we selectively separated Li from a leached mixed active cathode material.

3.2.2. Leaching with Sulfuric Acid

The solubility of metals increases with time in most cases. The leaching rate was very fast with sulfuric acid and ultrasounds for Li and Ni, allowing for the transition of these metals into the solution to occur in amounts of >95% during the first 60 min (Figure 4a). Co and Mn were also released from the cathode material, although their extraction was at a lower degree, i.e., 83.7% and 18.5% (after 60 min), respectively.
The leaching of nickel during mechanical mixing was limited in the first hour of the process, where 73.8% Ni was obtained. Only over time did the degree of Ni transition to solution increase. Despite the use of a low concentration of 0.75 M H2SO4, the efficiency of Ni was 89.7% after 120 min, and the final value of the obtained efficiency of Ni was 97.4%. Our results differ from those of Urbańska [55] and Wang et al. [62], who indicated lower efficiencies of Ni leaching. Those authors showed that at temperature of 55–60 °C and 90 °C using 1.5 M or 2.0 M H2SO4, Ni leaching did not exceed 45% within 120 min. On the other hand, Meshram et al. [63], using 1 M H2SO4, 90 °C, 4 h and no reductant agent, achieved a high degree of recovery of Ni 96.3%, Li 94.3%, Mn 50.2%, and Co 66.2% from Commercial LIBs of various compositions. The findings of that study align with our results, showing high extraction efficiency for Ni and Li and lower efficiency for Mn and Co. It appears that a longer processing time and a lower concentration of the applied acid enhance metal extraction. Additionally, the use of H2O2 increases the dynamics of Ni extraction as well as the extraction of the less soluble Co and Mn from spent batteries. For instance, as reported in [62], the extraction rate of Ni increases from 48.9% to 83.7% when the concentration of H2O2 is raised from 0 to 25 g/L. This suggests that the presence of an additional factor, such as a catalyst for the dissolution reaction, may be responsible for the high efficiency of leaching within the first 60 min, achieving 98.5% for Ni, 95.1% for Li, and 83.7% for Co during sonochemical leaching (Figure 4a).

3.2.3. Leaching with Citric Acid

The dynamics of the changes in concentration over time of all metals with citric acid, both in sonochemical leaching and with mixing, proceed much slower, in a characteristic linear manner, compared to oxalic and sulfuric acid were used (Figure 5).
Visible is, however, that, Co, and not Li reached the higher efficiency of transition to the solution (93.1%) after 240 minutes—while Mn did not exceed 1.8% leaching degree in systems with ultrasounds. Also, under in a vessel with stirring conditions, low leaching efficiency of Mn < 22% and other metals (not exceeding 40%) were observed.
The leaching efficiency of organic acids is directly correlated with their acidic strength (Table 5). From this, it follows that citric acid with triprotic nature, comprising three carboxylic acid groups, compared to oxalic acid, is a highly potent leaching agent. Oxalic acid, despite of diprotic nature, exhibited poor leaching efficiency for Ni, Mn, and Co from LIBs due to the formation of a lowly soluble metal oxalate precipitate [64], this resulted in low recovery values for these metals in vessels with stirring, i.e., not exceeding 1.9% (Figure 3b).
Research by Chen et al. [65] and other authors [30] clearly indicates the dependence of metal leaching efficiency on the volume concentration of the introduced H2O2 additive and the concentration of the citric acid. The lack of a reducing agent limited the leaching of Ni, Co, Mn, and Li to 51%, 48%, 43%, and 59%, respectively (time: 60 min, 80 °C) [62], and using a low concentration of citric acid (0.5 M) the efficiency of metal transition to solution did not exceed 60%. Increasing the citric acid concentration to 1.0 M allowed for the recovery of 89%, 86%, 86%, and 80% of Ni, Co, Mn, and Li, respectively [65]. Therefore, it may be concluded that the low concentration of 0.75 M citric acid used in our study and the absence of reducing agent in the system with mechanical mixing were the reasons for the low efficiency of metal recovery.

3.3. Effect of Leaching Temperature on Li Efficiency with Oxalic Acid

From hydrometallurgical practice it is known that generally, temperature intensifies the leaching of metals from solid phase. Therefore, the effect of lower temperature on the leaching efficiency of the valuable metals in LIBs was studied only for one case, i.e., using oxalic acid. We compared the mobilization of Li at room temperature (around 22 °C), 40 °C and at 60 °C when mechanical stirring and ultrasound (only at 60 °C) were applied (Figure 6). Due to the low leaching efficiency of Ni, Co, and Mn (<1%) with mechanical mixing, these metals were not included in the chart. As expected, at lower temperatures, the efficiency of Li leaching was at low level of 23.0% and 37.1% respectively, at 22 °C and 40 °C. However, Li extraction increased significantly with leaching temperature in the ultrasound system. Ultrasonic cavitation produces high local temperature, increasing with time from 30 °C after 5 min to a constant of 60 °C after 60 min. This was accompanied by a clear increase in the efficiency of Li leaching from 4.4% after 5 min at 30 °C to 94.1% when the system reached a temperature of 60 °C. This concurs with research by Yoon et al. [66], who showed, that increasing the temperature from 25 °C to 60 °C led to an increase in Li recovery from 7% to 100%. Similar conclusions were found in many works dedicated influence temperature on recover valuable metals from spent lithium-ion batteries [9,10,52,59,62,65,67]. Temperature is a factor, which significantly affects the increase of effectiveness of leaching Li and other metals, i.e., Co, Ni, and Mn, especially when an additional factor, reductant agent, is present [62]. For example, with increasing reaction temperature from 30°C to 50 °C, the leaching efficiency of Li, Ni, Co and Mn improved from 90.7%, 87.2%, 87.5% and 87.2% to 98.9%, 94.8%, 95.6% and 94.3%, respectively, [52] Also Okonkwo [53] improved that, compared to conventional mechanical stirring, the application of ultrasonic waves improves the leaching of metals by ~22–74%, with the degree of enhancement dependent on the temperature and the metal. The findings illustrated in Figure 6 confirm the expected results and align with data published in the literature. Specifically, the data demonstrate that lithium (Li) leaching efficiency was significantly higher at 60 °C, as anticipated. The best leaching results were observed at this temperature, particularly when ultrasonic stirring (US) was employed. This corroborates previous studies indicating that increased temperature and the presence of ultrasonic waves enhance Li extraction efficiency. Therefore, maintaining a constant temperature of 60 °C from the very beginning of the process should be considered in further research on the recovery of metals from LIBs.

3.4. Leaching under Sonic Conditions vs. a Mechanical Stirring System

Our experiments were conducted without the addition of reductants such as H2O2, glucose, NaHSO3 or ascorbic acid, applying ultrasound as the only intensifying factor. In the systems with ultrasounds present, the effectiveness of transitioning metals to a solution was higher in relation to systems applying the stirring leaching method, as displayed in Figure 7.
Depending on the leaching medium used, in addition to Li, Co, Ni and Mn also pass into the solution. During the ultrasound process, vacuum holes are produced in the solution, and the formed bubbles are enlarged gradually when the sound pressure increased via the transmission of ultrasonic waves [51,68,69]. Growing bubbles attack the surface of the ground material and finally burst with high pressure accelerating the leaching of metals. Ultrasonic waves promote not only cavitation but also convective motion, which increases the contact area of solid and liquid, providing a large amount of energy, which facilitates the dissolution of material [27]. Ultrasound significantly improves the mass transfer rates by causing cavitation, which leads to the formation of micro-jets and shock waves. This phenomenon enhances the penetration of the leaching agents into the solid matrix. Yan et al. [51] reported up to 10% increase in the leaching efficiency of Li, Co, Ni, and Mn using ultrasound compared to traditional methods. Also other authors [45,50,51,54] using different organic media demonstrated higher ultrasound leaching efficiencies for cobalt and lithium. Importantly, in our research, we also considered the possibility of the presence of an additional effect of ultrasound waves phenomenon—the formation of H2O2. In the presence of ultrasound, in aqueous systems and organic acids, in addition to the local increase in temperature, pressure formation, and the growth and implosive collapse of bubbles, transient cavitation initiates the formation of hydrogen atoms and free hydroxyl radicals which may recombine to form hydrogen and hydrogen peroxide, depicted by reactions [70,71,72]:
H2O → H + HO
H + O 2 H O O
2 H O H 2 O 2
2 H O O H 2 O 2
The creation of free radicals and the presence of H2O2 in systems with ultrasound supports the extraction and the transition of Co, Ni and Mn into solution. Hydrogen peroxide is one of the most effective reductants in terms of enhancing cobalt, manganese and nickel leaching from LIBs in organic [10,28,30,59] and inorganic (H2SO4, HCl, HNO3) acids [10,62,73,74], as follows:
2LiCoO2(s) + 3H2C2O4(aq) + H2O2(aq) → Li2C2O4(aq) + 2CoC2O4(s) + 4H2O + O2(g)
2LiCoO2(s) + 3H2SO4(aq) + H2O2(aq) → Li2SO4(aq) + 2CoSO4(s) + 4H2O + O2(g)
LiNi1/3Mn1/3Co1/3O2(s) + 9H2SO4(aq) + H2O2(aq) → 2NiSO4(aq)+ 2MnSO4(aq) + 2CoSO4(aq) + 3Li2SO4(aq) + 2O2(g) + 10H2O
Hydrogen peroxide has been used successfully as a reducing agent to increase the leaching rate Li and Co in leaching systems without ultrasounds. The authors of [75] reported increase in Li, Ni, Co, and Mn leaching efficiencies to over 98% within 30 min by adding H2O2 at a temperature, 50 °C. Bertuol et al. [76] explored the use of H2O2 and showed that at atmospheric pressure, cobalt leaching was favoured by increasing the amount of H2O2 (from 0 to 8% v/v) reaching more than 95 wt.% of cobalt during 5 min with H2SO4 and CO2. Also, Chen et al. [77] added peroxide to phosphoric acid, increasing Co recovery up to 99%. Comparing the behavior of Co, Ni, and Mn in mixing and ultrasound systems in oxalic acid, it was observed, that only Co transitioned into the solution throughout sonoleaching. It can therefore be concluded, that not only it was this caused directly by the temperature, but that the presence of ultrasound and the effect of the creation of H2O2 may have been responsible for catalyzing the dissolution reaction of Co. Cathode transition metals have low solubility as they exist in +3/+4 valence states in discharged cathodes and are difficult to leach due to the strong metal–oxide bonds [10,50]. It seems, that the rapid realize of Li first from the solid phase of the cathode causes the rupture of the bond structure of the LiCoO2 material, followed by the release of Co and the reduction of Co3+ to Co2+, which than goes into solution. Hydrogen peroxide production during cavitation depends on the intensities of the ultrasonic irradiation and time [72]. It has been indicated that the H2O2 concentration is higher where the ultrasonic intensity is greater and the time is longer [78,79]. This explains the release of Co after 60 min, i.e., the time needed for H2O2 to form; this peroxide is then used by the released Co faster than Mn and Ni, which remain bound in the solid phase. Additionally, the reduction of Co3+ to Co2+ by H2O2 might be further accelerated by increasing the temperature; this is in agreement with previous findings [27,30]. It was possible to recover 99.6% of the Li and 31.2% of the Co during sonochemical leaching with oxalic acid, while in the stirring system, it was only 59.2% of the Li and 4.4% of the Co. Ultrasound also supports the dissolution of metals from batteries in citric acid and H2SO4 solutions. In these cases, however, co-dissolution of all metals with varying efficiency was observed. In addition to the partial dissolution of Co and Mn in the H2SO4 solution, a slightly higher degree of Ni recovery was obtained, i.e., 99.9% compared to 94.7% in the mixing system. Also for Co, a higher leaching efficiency was achieved during sonoleaching: 62.8% compared to 42.8% during mechanical mixing. Citric acid was shown to be one of the best chelating agents, as it should accelerate the dissolution process [27]. However, compared to H2SO4 and oxalic acid, chelating only with Co2+ was s greatly accelerated citric acid by the cavitation action of ultrasound [27] and the possible presence of H2O2; 93.1% of Co was leached with ultrasound, while in the stirring system, it was 42.8%.

3.5. Cathode Active Material after Ultrasound-Assisted Leaching

Figure 8 and Figure 9 present, respectively, SEM pictures and the particle-size distribution of the initial cathode active material following the milling process and mechanical sieving (fraction < 0.1 mm) and the residue after ultrasound-assisted leaching with sulfuric acid. According to Figure 8, the initial material had a dense structure with porous agglomerates, while the residue was looser with visible individual grains of material and smaller particle sizes.
Leached material after the ultrasonic-assisted treatment shows a significant reduction of particle size relative to initial material, what is described in many works [27,80,81,82,83]. The results of particle size distribution (Figure 8) also confirm SEM analysis. The resulting curves differ significantly from one another. D90 value is below 35 µm, D50 below 15 µm for the initial cathode material before leaching and the proportion of the particle diameter below 10 μm was small (Figure 9). Compared with the residue after ultrasound leaching, the particle size below 10 μm of the cathode material was obviously predominant and D90 value was below 11 µm and D50 below 3 µm. Particle size reduction is directly related to the mechanical ultrasound effect and accompanied cavitation phenomenon damaged the solid surface [45,67,79,83]. Ultrasound increases the specific surface area of materials and provides contact of the solid phase with the solvents which accelerates the dissolution reaction and improves the leaching rate.

4. Conclusions

Ultrasound-assisted leaching is a suitable technique for metal recovery, as it significantly improved the efficiency of the Li, Co, Mn and Ni extraction process, particularly with organic acids (citric and oxalic acids), compared to the mechanical stirring leaching process. An average of 50% better leaching results were obtained for Li recovery in oxalic acid and of the Co in citric acid during sonochemical leaching, allowing us to obtain 99.6% of the Li and 93.1% after 4 hours. Oxalic acid, as a leaching and precipitating agent, has a good selectivity in the separation of Li from Co, Ni, and Mn, e.g., in the mixed cathode active material. The optimum Li leach conditions for separation were 0.75 M C2H2O4 at 60 °C for 60 min using ultrasound assistant leaching where 94.1% of Li was recovered. Based on the theory of free hydroxyl radicals and hydrogen peroxide formation during ultrasound waves action in acids solutions, it may be concluded that hydrogen peroxide is a support factor that reduces the high valence of metals to a lower state, enhancing cobalt, manganese and nickel leaching from LIBs.
Future research in ultrasound-assisted leaching should focus on optimizing ultrasound parameters such as frequency, power, and duration to enhance metal recovery efficiency. Further exploration of other organic acids, beyond citric and oxalic, could lead to improved selectivity and recovery rates for various metals. Mechanistic studies are essential to understand the influence of ultrasound on the chemical and physical properties of the leachate and its effect on the dissolution and precipitation of metals. Additionally, assessing the environmental impact and sustainability of the use of organic acids in this process is crucial, including evaluating the potential for recycling these acids to reduce waste and costs.
Scaling up the laboratory processes for industrial applications is another vital area of study, addressing challenges related to equipment design, process control, and energy consumption. The technology’s applicability to other battery types and electronic waste (e-waste) containing valuable metals should also be investigated, potentially broadening its scope. Moreover, integrating ultrasound-assisted leaching with other recovery techniques, such as bioleaching or electrochemical methods, could enhance overall efficiency and recovery rates. This research aims to refine and expand the use of ultrasound-assisted leaching, potentially revolutionizing metal recovery from various sources and contributing to more sustainable recycling practices.

Author Contributions

Conceptualization, J.W. and A.F.; methodology, J.W.; software, B.G.; validation, A.F. and J.W.; formal analysis, J.W.; investigation, A.S., D.W., A.M., B.P. and M.K.; resources, B.P. and M.K.; data curation, T.F. and A.M; writing—original draft preparation, J.W. and A.F.; writing—review and editing, A.F.; visualization, D.W. and A.S.; supervision, A.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the European Union from the European Social Fund in the framework of the project “Silesian University of Technology as a Center of Modern Education based on Research and Innovation” POWR.03.05.00-00-Z098/17.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Blengini, G.A.; El Latunussa, C.; Eynard, U.; Torres De Matos, C.; Wittmer, D.; Georgitzikis, K.; Pavel, C.; Carrara, S.; Mancini, L.; Unguru, M.; et al. Study on the EU’s List of Critical Raw Materials; Final Report 31.01.2020; European Commission: Brussels, Belgium, 2020. [Google Scholar]
  2. Gaines, L.; Richa, K.; Spangenberger, J. Key issues for Li-ion battery recycling. MRS Energy Sustain. 2018, 5, 12. [Google Scholar] [CrossRef]
  3. Leszczyńska-Sejda, K.; Chmielarz, A.; Kopyto, D.; Ochmański, M.; Benke, G.; Palmowski, A.; Sobianowska-Turek, A.; Łoś, P.; Fornalczyk, A.; Zygmunt, M.; et al. An Innovative Method of Leaching of Battery Masses Produced in the Processing of Li-Ion Battery Scrap. Appl. Sci. 2024, 14, 397. [Google Scholar] [CrossRef]
  4. Yu, W.; Guo, Y.; Xu, S.; Yang, Y.; Zhao, Y.; Zhang, J. Comprehensive recycling of lithium-ion batteries: Fundamentals, pretreatment, and perspectives. Energy Storage Mater. 2023, 5, 172–220. [Google Scholar] [CrossRef]
  5. Available online: https://www.nsenergybusiness.com/news/industry-news/global-lithium-demand-2024/ (accessed on 10 June 2024).
  6. Available online: https://rmis.jrc.ec.europa.eu/?page=crms-in-strategic-sectors-and-technologies-e8c632 (accessed on 10 June 2024).
  7. Fourth List of CRMs—In 2020-Communication from the Commission to the European Parliament, the Council, the European Economic and Social Committee and the Committee of the Regions. Critical Raw Materials Resilience: Charting a Path towards greater Security and Sustainability; COM(2020) 474 final; European Commission: Brussels, Belgium, 2020.
  8. Halleux, V. New EU Regulatory Framework for Batteries Setting Sustainability Requirements; BRIEFING EU Legislation in Progress PE 689.337; European Parliamentary Research Service: Brussels, Belgium, 2021. [Google Scholar]
  9. Pinegar, H.; Smith, Y.R. Recycling of End-of-Life Lithium-Ion Batteries, Part II: Laboratory-Scale Research Developments in Mechanical, Thermal, and Leaching Treatments. J. Sustain. Metall. 2020, 6, 142–160. [Google Scholar] [CrossRef]
  10. Or, T.; Gourley, S.W.D.; Kaliyappan, K.; Yu, A.; Chen, Z. Recycling of mixed cathode lithium-ion batteries for electric vehicles, Current status and future outlook. Carbon Energy 2020, 2, 6–43. [Google Scholar] [CrossRef]
  11. Velázquez-Martínez, O.; Valio, J.; Santasalo-Aarnio, A.; Reuter, M.; Serna-Guerrero, R. A Critical Review of Lithium-Ion Battery Recycling Processes from a Circular Economy Perspective. Batteries 2019, 5, 68. [Google Scholar] [CrossRef]
  12. Sobianowska-Turek, A.; Urbanska, W.; Janicka, A.; Zawiślak, M.; Matla, J. The Necessity of Recycling of Waste Li-Ion Batteries Used in Electric Vehicles as Objects Posing a Threat to Human Health and the Environment. Recycling 2021, 6, 35. [Google Scholar] [CrossRef]
  13. Guimarães, L.F.; Botelho Junior, A.B.; Espinosa, D.C.R. Sulfuric acid leaching of metals from waste Li-ion batteries without using reducing agent. Miner. Eng. 2022, 183, 107597. [Google Scholar] [CrossRef]
  14. Shin, S.M.; Kim, N.H.; Sohn, J.S.; Yang, D.H.; Kim, Y.H. Development of a metal recovery process from Li-ion battery wastes. Hydrometallurgy 2005, 79, 172–181. [Google Scholar] [CrossRef]
  15. Balázs Illés, I.; Kékesi, T. Extraction of pure Co, Ni, Mn, and Fe compounds from spent Li-ion batteries by reductive leaching and combined oxidative precipitation in chloride media. Miner. Eng. 2023, 201, 108169. [Google Scholar] [CrossRef]
  16. Sadia, I.; Rajiv Ranjan, S.; Vinay, K.S.; Ruan, C.; Hyunjung, K. Recovery of critical metals from spent Li-ion batteries: Sequential leaching, precipitation, and cobalt–nickel separation using Cyphos IL104. Waste Manag. 2022, 154, 175–186. [Google Scholar] [CrossRef]
  17. Qian, H.; Hong, Z.; Zhanfang, C. High-efficiency leaching of Li and Ni from spent lithium-ion batteries based on sodium persulfate. Sep. Purif. Technol. 2023, 325, 124653. [Google Scholar] [CrossRef]
  18. Song, D.; Wang, T.; Liu, Z.; Zhao, S.; Quan, J.; Li, G.; Zhu, H.; Huang, J.; He, W. Characteristic comparison of leaching valuable metals from spent power Li-ion batteries for vehicles using the inorganic and organic acid system. J. Environ. Chem. Eng. 2022, 10, 107102. [Google Scholar] [CrossRef]
  19. Sedlakova-Kadukova, A.; Marcincakova, R.; Mrazikova, A.; Willner, J.; Fornalczyk, A. Closing the loop: Key role of iron in metal-bearing waste recycling. Arch. Metall. Mater. 2017, 62, 1459–1466. [Google Scholar] [CrossRef]
  20. Arya, S.; Kumar, S. Bioleaching: Urban mining option to curb the menace of E-waste challenge. Bioengineered 2020, 11, 640–660. [Google Scholar] [CrossRef] [PubMed]
  21. Golmohammadzadeh, R.; Rashchi, F.; Vahidi, E. Recovery of lithium and cobalt from spent lithium-ion batteries using organic acids: Process optimization and kinetic aspects. Waste Manag. 2017, 64, 244–254. [Google Scholar] [CrossRef]
  22. Meshram, P.; Mishra, A.; Sahu, R. Environmental impact of spent lithium ion batteries and green recycling perspectives by organic acids—A review. Chemosphere 2020, 242, 125291. [Google Scholar] [CrossRef]
  23. Li, D.; Yin, Y.; Jiao, B.; Song, L.; Wang, Y.; Qiao, J.; Xu, Z. Chemical dissolving of citric acid on bioleaching of copper mine tailings by Acidithiobacillus ferrooxidans ATCC 23270. Res. J. Biotechnol. 2013, 8, 12. [Google Scholar]
  24. Musariri, B.; Akdogan, G.; Dorfling, C.; Bradshaw, S. Evaluating organic acids as alternative leaching reagents for metal recovery from lithium ion batteries. Miner. Eng. 2019, 137, 108–117. [Google Scholar] [CrossRef]
  25. Golmohammadzadeh, R.; Faraji, F.; Rashchi, F. Recovery of lithium and cobalt from spent lithium ion batteries (LIBs) using organic acids as leaching reagents: A review. Resour Conserv Recycl. 2018, 136, 418–435. [Google Scholar] [CrossRef]
  26. Li, L.; Lu, J.; Ren, Y.; Zhang, X.X.; Chen, R.J.; Wu, F.; Amine, K. Ascorbic-acid-assisted recovery of cobalt and lithium from spent Li-ion batteries. J. Power Sources 2012, 218, 21–27. [Google Scholar] [CrossRef]
  27. Li, L.; Zhai, L.; Zhang, X.; Lu, J.; Chen, R.; Wu, F.; Amine, K. Recovery of valuable metals from spent lithium-ion batteries by ultrasonic-assisted leaching process. J. Power Sources 2014, 262, 380–385. [Google Scholar] [CrossRef]
  28. Park, Y.; Lim, H.; Moon, J.-H.; Lee, H.-N.; Son, S.H.; Kim, H.; Kim, H.-J. High-yield one-pot recovery and characterization of nanostructured cobalt oxalate from spent lithium-ion batteries and successive re-synthesis of LiCoO2. Metals 2017, 7, 303. [Google Scholar] [CrossRef]
  29. Menga, F.; Liua, Q.; Kim, R.; Wang, J.; Liu, G.; Ghahreman, A. Selective recovery of valuable metals from industrial waste lithium-ion batteries using citric acid under reductive conditions: Leaching optimization and kinetic analysis. Hydrometallurgy 2020, 191, 105160. [Google Scholar] [CrossRef]
  30. Li, L.; Ge, J.; Wu, F.; Chen, R.; Chen, S.; Wu, B. Recovery of cobalt and lithium from spent lithium ion batteries using organic citric acid as leachant. J. Hazard. Mater. 2010, 176, 288–293. [Google Scholar] [CrossRef]
  31. Zeng, X.; Li, J.; Shen, B. Novel approach to recover cobalt and lithium from spent lithium-ion battery using oxalic acid. J. Hazard Mater. 2015, 295, 112–118. [Google Scholar] [CrossRef]
  32. Gao, W.; Zhang, X.; Zheng, X.; Lin, X.; Cao, H.; Zhang, Y.; Sun, Z. Lithium carbonate recovery from cathode scrap of spent lithium-ion battery: A closed loop process. Environ. Sci. Technol. 2017, 51, 1662–1669. [Google Scholar] [CrossRef]
  33. Li, L.; Qu, W.; Zhang, X.; Lu, J.; Chen, R.; Wu, F.; Amine, K. Succinic acid-based leaching system: A sustainable process for recovery of valuable metals from spent Li-ion batteries. J. Power Sources 2015, 282, 544–551. [Google Scholar] [CrossRef]
  34. Urbanska, W.; Osial, M. Investigation of the Physico-Chemical Properties of the Products Obtained after Mixed Organic-Inorganic Leaching of Spent Li-Ion Batteries. Energies 2020, 13, 6732. [Google Scholar] [CrossRef]
  35. Rix, A.; Lederle, W.; Theek, B.; Lammers, T.; Moonen, C.; Schmitz, G.; Kiessling, F. Advanced Ultrasound Technologies for Diagnosis and Therapy. J. Nucl. Med. 2018, 59, 740–746. [Google Scholar] [CrossRef]
  36. Xu, T.; Xu, L.P.; Zhang, X. Ultrasound propulsion of micro-/nanomotors. Appl. Mater. Today 2017, 9, 493–503. [Google Scholar] [CrossRef]
  37. Li, J.; Li, T.; Xu, T.; Kiristi, M.; Liu, W.; Wu, Z.; Wang, J. Magneto–Acoustic Hybrid Nanomotor. Nano Lett. 2015, 15, 4814–4821. [Google Scholar] [CrossRef] [PubMed]
  38. Huang, Z.Y.; Xie, F.C.; Ma, Y. Ultrasonic recovery of copper and iron through the simultaneous utilization of Printed Circuit Boards (PCB) spent acid etching solution and PCB waste sludge. J. Hazard Mater. 2011, 185, 155–161. [Google Scholar] [CrossRef] [PubMed]
  39. Toache-Pérez, A.D.; Bolarín-Miró, A.M.; Sánchez-De Jesús, F.; Lavine, G.T. Facile method for the selective recover of Gd and Pr from LCD screen wastes using ultrasound-assisted leaching. Sustain. Environ. Res. 2020, 30, 20. [Google Scholar] [CrossRef]
  40. Oza, R.; Shah, N.; Patel, S. Recovery of nickel from spent catalysts using ultrasonication-assisted leaching. J. Chem. Technol. Biotechnol. 2011, 86, 1276–1281. [Google Scholar] [CrossRef]
  41. Chang, J.; Zhang, E.; Yang, C.; Zhou, J.; Peng, J.; Zhang, L.; Srinivasakannan, C. Kinetics of ultrasound-assisted silver leaching from sintering dust using thiourea. Green Process. Synth. 2016, 5, 31–40. [Google Scholar] [CrossRef]
  42. Yin, S.; Pei, J.; Jiang, F.; Li, S.; Peng, J.; Zhang, L.; Ju, S.; Srinivasakannan, C. Ultrasound-assisted leaching of rare earths from the weathered crust elution deposited ore using magnesium sulfate without ammonia-nitrogen pollution. Ultrason. Sonochem. 2020, 41, 156–162. [Google Scholar] [CrossRef]
  43. Sobianowska-Turek, A.; Grudniewska, K.; Maciejewski, P.; Gawlik-Kobylińska, M. Removal of Zn(II) and Mn(II) by Ion Flotation from Aqueous Solutions Derived from Zn-C and Zn-Mn(II) Batteries Leaching. Energies 2021, 14, 1335. [Google Scholar] [CrossRef]
  44. Li, H.; Li, S.; Peng, J.; Srinivasakannan, C.; Zhang, L.; Yin, S. Ultrasound augmented leaching of nickel sulfate in sulfuric acid and hydrogen peroxide media. Ultrason. Sonochem. 2018, 40, 1021–1030. [Google Scholar] [CrossRef]
  45. Esmaeili, M.; Rastegar, S.O.; Beigzadeh, R.; Gu, T. Ultrasound-assisted leaching of spent lithium ion batteries by natural organic acids and H2O2. Chemosphere 2020, 254, 126670. [Google Scholar] [CrossRef]
  46. Vyas, S.; Ting, Y.P. A Review of the Application of Ultrasound in Bioleaching and Insights from Sonication in (Bio)Chemical Processes. Resources 2018, 7, 3. [Google Scholar] [CrossRef]
  47. Swamy, K.M.; Sarveswara, K.; Narayana, K.L.; Murty, J.S.; Ray, H.S. Application of Ultrasound in Leaching. Miner. Process. Extr. Metall. Rev. 1995, 14, 179–192. [Google Scholar] [CrossRef]
  48. Tzanakis, I.; Lebon, G.S.B.; Eskin, D.G.; Pericleous, K.A. Characterizing the cavitation development and acoustic spectrum in various liquids. Ultrason. Sonochem. 2017, 34, 651–662. [Google Scholar] [CrossRef]
  49. Available online: https://www.reuters.com/article/us-catl-batteries-idUSKCN25B0BA (accessed on 18 January 2024).
  50. Jiang, F.; Chen, Y.; Ju, S.; Zhu, Q.; Zhang, L.; Peng, J.; Wang, X.; Miller, J.D. Ultrasound-assisted leaching of cobalt and lithium from spent lithium-ion batteries. Ultrason. Sonochem. 2018, 48, 88–95. [Google Scholar] [CrossRef] [PubMed]
  51. Nshizirungu, T.; Rana, M.; Jo, Y.T.; Uwiragiye, E.; Kim, J.; Park, J.-H. Ultrasound-assisted sustainable recycling of valuable metals from spent Li-ion batteries via optimisation using response surface methodology. J. Environ. Chem. Eng. 2024, 12, 2371. [Google Scholar] [CrossRef]
  52. Yan, S.; Sun, C.; Zhou, T.; Gao, R.; Xie, H. Ultrasonic-assisted leaching of valuable metals from spent lithium-ion batteries using organic additives. Sep. Purif. Technol. 2021, 257, 117930. [Google Scholar] [CrossRef]
  53. Okonkwo, E.G.; Wheatley, G.; Liu, Y.; He, Y. A cavitation enabled green leaching of metals from spent lithium-ion batteries. Chem. Eng. Process.-Process. Intensif. 2024, 202, 109850. [Google Scholar] [CrossRef]
  54. Ning, P.; Meng, Q.; Dong, P.; Duan, J.; Xu, M.; Lin, Y.; Zhang, Y. Recycling of cathode material from spent lithium ion batteries using an ultrasound-assisted DL-malic acid leaching system. Waste Manag. 2020, 103, 52–60. [Google Scholar] [CrossRef]
  55. Urbańska, W. Recovery of Co, Li, and Ni from Spent Li-Ion Batteries by the Inorganic and/or Organic Reducer Assisted Leaching Method. Minerals 2020, 10, 555. [Google Scholar] [CrossRef]
  56. Morrison, R.T.; Boyd, R.N. Chap. 11. In Organic Chemistry, 6th ed.; Prentice Hall International, Inc.: London, UK, 1992. [Google Scholar]
  57. Krishnamurty, K.V.; Harris, G.M. The Chemistry of the Metal Oxalato Complexes. Chem. Rev. 1961, 61, 213–246. [Google Scholar] [CrossRef]
  58. Zhang, X.; Bian, Y.; Xu, S.; Fan, E.; Xue, Q.; Guan, Y.; Wu, F.; Li, L.; Chen, R. Innovative Application of Acid Leaching to Regenerate Li(Ni1/3Co1/3Mn1/3)O2 Cathodes from Spent Lithium-Ion Batteries. ACS Sustain. Chem. Eng. 2018, 6, 5959–5968. [Google Scholar] [CrossRef]
  59. Sun, L.; Qiu, K. Organic oxalate as leachant and precipitant for the recovery of valuable metals from spent lithium-ion batteries. Waste Manag. 2021, 32, 1575–1582. [Google Scholar] [CrossRef] [PubMed]
  60. Rumble, J.R.; Lide, D.R.; Bruno, T.J. CRC Handbook of Chemistry and Physics: A Ready-Reference Book of Chemical and Physical Data; CRC Press: Boca Raton, FL, USA, 2018. [Google Scholar]
  61. Verma, A.; Kore, R.; Corbin, D.R.; Shiflett, M.B. Metal Recovery Using Oxalate Chemistry: A Technical Review. Ind. Eng. Chem. Res. 2019, 58, 15381–15393. [Google Scholar] [CrossRef]
  62. Wang, H.; Friedrich, B. Development of a highly efficient hydrometallurgical recycling process for automotive Li–ion batteries. J. Sustain. Met. 2015, 1, 168–178. [Google Scholar] [CrossRef]
  63. Meshram, P.; Pandey, B.D.; Mankhand, T.R. Recovery of valuable metals from cathodic active material of spent lithium ionbatteries: Leaching and kinetic aspects. Waste Manag. 2015, 45, 306–313. [Google Scholar] [CrossRef] [PubMed]
  64. Yao, X.; Xu, Z.; Yao, Z.; Cheng, W.; Gao, H.; Zhao, Q.; Li, J.; Zhou, A. Oxalate co-precipitation synthesis of LiNi0.6Co0.2Mn0.2O2 for low-cost and high-energy lithium-ion batteries. Mater. Today Commun. 2019, 19, 262–270. [Google Scholar] [CrossRef]
  65. Chen, X.; Zhou, T. Hydrometallurgical process for the recovery of metal values from spent lithium-ion batteries in citric acid media. Waste Manag. Res. 2014, 32, 1083–1093. [Google Scholar] [CrossRef]
  66. Yoon, W.S.; Nam, K.-W.; Jang, D.; Chung, K.Y.; Hanson, J.; Chen, J.-M.; Yang, X.-Q. Structural study of the coating effect on the thermal stability of charged MgO-coated LiNi0.8Co0.2O2 cathodes investigated by in situ XRD. J. Power Sources 2012, 217, 128–134. [Google Scholar] [CrossRef]
  67. Sohn, J.-S.; Shin, S.-M.; Yang, D.-H.; Kim, S.-K.; Lee, C.-K. Comparison of Two Acidic Leaching Processes for Selecting the Effective Recycle Process of Spent Lithium ion Battery. Geosyst. Eng. 2006, 9, 1−6. [Google Scholar] [CrossRef]
  68. Eskin, G.I. Cavitation mechanism of ultrasonic melt degassing. Ultrason. Sonochem. 1995, 2, S137–S141. [Google Scholar] [CrossRef]
  69. Zhang, K.; Li, B.; Wu, Y.; Wang, W.; Li, R.; Zhang, Y.N.; Zuo, T. Recycling of indium from waste LCD: A promising non-crushing leaching with the aid of ultrasonic wave. Waste Manag. 2017, 64, 236–243. [Google Scholar] [CrossRef] [PubMed]
  70. Riesz, P.; Berdahl, D.; Christman, C.L. Free radical generation by ultrasound in aqueous and nonaqueous solutions. Environ. Health Perspect. 1985, 64, 233–252. [Google Scholar] [CrossRef]
  71. Thompson, L.H.; Doraiswamy, L.K. Sonochemistry: Science and engineering. Ind. Eng. Chem. Res. 1999, 38, 1215–1249. [Google Scholar] [CrossRef]
  72. Ziembowicz, S.; Kida, M.; Koszelnik, P. Sonochemical Formation of Hydrogen Peroxide. Proceedings 2018, 2, 188. [Google Scholar]
  73. Jha, M.K.; Kumari, A.; Jha, A.K.; Kumar, V.; Hait, J.; Pandey, B.D. Recovery of lithium and cobalt from waste lithium ion batteries of mobile phone. Waste Manag. 2013, 33, 1890–1897. [Google Scholar] [CrossRef] [PubMed]
  74. Lee, C.K.; Rhee, K.I. Preparation of LiCoO2 from spent lithium-ion batteries. J. Power Sources 2002, 109, 17–21. [Google Scholar] [CrossRef]
  75. Sattar, R.; Ilyas, S.; Bhatti, H.N.; Ghaffar, A. Resource recovery of critically-rare metals by hydrometallurgical recycling of spent lithium ion batteries. Sep. Purif. Technol. 2019, 209, 725–733. [Google Scholar] [CrossRef]
  76. Bertuol, D.A.; Machado, C.M.; Silva, M.L.; Calgaro, C.O.; Dotto, G.L.; Tanabe, E.H. Recovery of cobalt from spent lithium-ion batteries using supercritical carbon dioxide extraction. Waste Manag. 2016, 51, 245–251. [Google Scholar] [CrossRef] [PubMed]
  77. Chen, X.; Ma, H.; Luo, C.; Zhou, T. Recovery of valuable metals from waste cathode materials of spent lithium-ion batteries using mild phosphoric acid. J. Hazard. Mater. 2017, 326, 77–86. [Google Scholar] [CrossRef] [PubMed]
  78. Lim, M.; Son, Y.; Khim, J. The effects of hydrogen peroxide on the sonochemical degradation of phenol and bisphenol A. Ultrason. Sonochem. 2014, 21, 1976–1981. [Google Scholar] [CrossRef]
  79. Pétrier, C. 31: The use of power ultrasound for water treatment. In Power Ultrasonics; Woodhead Publishing: Oxford, UK, 2015; pp. 939–972. [Google Scholar]
  80. Franco, F.; Pérez-Maqueda, L.A.; Pérez-Rodríguez, J.L. The effect of ultrasound on the particle size and structural disorder of a well-ordered kaolinite. J. Colloid Interface Sci. 2004, 274, 107–117. [Google Scholar] [CrossRef]
  81. Pilli, S.; Bhunia, P.; Yan, S.; LeBlanc, R.; Tyagi, R.; Surampalli, R. Ultrasonic pretreatment of sludge: A review. Ultrason. Sonochem. 2011, 18, 1–18. [Google Scholar] [CrossRef]
  82. Mao, Y.; Xie, G.; Liang, L.; Xia, W.; Peng, Y. Effects of ultrasonic treatment on the particle size, shape and ash content of fine coal. Physicochem. Probl. Miner. Process. 2019, 55, 679–688. [Google Scholar] [CrossRef]
  83. Chang, J.; Zhang, E.-D.; Zhang, L.-B.; Peng, J.-H.; Zhou, J.-W.; Srinivasakannan, C.; Yang, C.-J. A comparison of ultrasound-augmented and conventional leaching of silver from sintering dust using acidic thiourea. Ultrason. Sonochem. 2017, 34, 222–231. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Active cathode material separated from Al foil.
Figure 1. Active cathode material separated from Al foil.
Energies 17 04055 g001
Figure 2. SEM-EDS analysis for cathode active material (a) coming from a computer (b) from mobile phones, and (c) a mixed material.
Figure 2. SEM-EDS analysis for cathode active material (a) coming from a computer (b) from mobile phones, and (c) a mixed material.
Energies 17 04055 g002
Figure 3. Leaching efficiency of metals with oxalic acid during (a) sonochemical leaching (b) mechanical stirring leaching.
Figure 3. Leaching efficiency of metals with oxalic acid during (a) sonochemical leaching (b) mechanical stirring leaching.
Energies 17 04055 g003
Figure 4. Leaching efficiency of metals during (a) sonoleaching (b) mechanical stirring with sulfuric acid.
Figure 4. Leaching efficiency of metals during (a) sonoleaching (b) mechanical stirring with sulfuric acid.
Energies 17 04055 g004
Figure 5. Leaching efficiency of metals during (a) sonoleaching (b) mechanical stirring with citric acid.
Figure 5. Leaching efficiency of metals during (a) sonoleaching (b) mechanical stirring with citric acid.
Energies 17 04055 g005
Figure 6. Effect of temperature on Li recovery in sonochemical (US) and stirring leaching (ST).
Figure 6. Effect of temperature on Li recovery in sonochemical (US) and stirring leaching (ST).
Energies 17 04055 g006
Figure 7. Comparison of metals leaching efficiency with different methods (0.75 M acids, 60 °C, 240 min): ultrasound leaching (US) and stirring leaching (ST).
Figure 7. Comparison of metals leaching efficiency with different methods (0.75 M acids, 60 °C, 240 min): ultrasound leaching (US) and stirring leaching (ST).
Energies 17 04055 g007
Figure 8. SEM images of cathode active material. (a) before leaching, (b) residue after ultrasound leaching.
Figure 8. SEM images of cathode active material. (a) before leaching, (b) residue after ultrasound leaching.
Energies 17 04055 g008
Figure 9. Particle size distribution of active cathode material (a) before leaching and (b) after ultrasound-assisted leaching.
Figure 9. Particle size distribution of active cathode material (a) before leaching and (b) after ultrasound-assisted leaching.
Energies 17 04055 g009
Table 1. Experimental conditions of the ultrasound-assisted leaching of metals from LIBs described in the literature.
Table 1. Experimental conditions of the ultrasound-assisted leaching of metals from LIBs described in the literature.
Type of MaterialLeaching AgentType of Sonic Source/PowerLeaching ConditionsLeaching EffectivenessRef.
LiCoO2 2 M H2SO4 + 5% H2O2direct/360 W0.5 h, 30 °C, 100 g/LCo: 91.6%
Li: 92.7%
[50]
LiCoO2citric acid + H2O2,indirect0.5 h, 40 °CLi: 100%
Co: 100%
Ni: 100%
[45]
LiCoO20.5 M C6H8O7 + 0.55 M H2O2,indirect/90 W5 h, 60 °C,
25 g/L
Co: 96.1%
Li: 98.4%
[27]
LiCoO22 M H2SO4 + 0.55 M H2O2indirect/90 W5 h, 40 °C,
25 g/L
Co: 45.7%
Li: 97.0%
LiCoO22 M HClindirect/90 W5 h, 60 °C,
25 g/L
Co: 76.4%
Li: 98.2%
LiCoO20.3 M C6H12O7indirect/300 W1 h, 180 °C,
3 g/200 mL
Co: 96.0%
Li: 98.0%
[51]
LiCoO20.5 M CH3COOH + 0.2 M C6H8O6 + bagasse pithdirect/450 W40 min, 50 °C,
20 g/L
Co: 98.6%
Li: 99.6%
[52]
LiCoO2CH4O3S +sugarcane molassesdirect/400 W1 h, 90 °C,
20 g/L
Co: 97.0%
Li: 97.0%
[53]
LiCoO21 M DL C4H6O5 + H2O2indirect/90 W0.5 h, 80 °C,
5 g/L
Co: 97.3%
Li: 98.0%
[54]
Table 2. Composition of lithium-ion battery LG Chem, Ltd. (Wrocław, Poland) and Samsung (Yongin, Korea) SDI lithium-ion cell/battery [47,48].
Table 2. Composition of lithium-ion battery LG Chem, Ltd. (Wrocław, Poland) and Samsung (Yongin, Korea) SDI lithium-ion cell/battery [47,48].
Ingredient Content, %
Aluminum Foil2–10
Metal Oxide (proprietary)20–50
Polyvinylidene Fluoride (PVDF)<5
Copper Foil2–10
Carbon (proprietary)10–30
Electrolyte (proprietary10–20
Stainless steel, Nickel and inert materials residue
Table 3. Conditions of the experiments.
Table 3. Conditions of the experiments.
Type of AcidLeaching SystemTemperature,
°C
Initial pH of Leach Liquor
C2H2O4Mechanical stirrer220.68
40
60
Ultrasound60
C6H8O7Mechanical stirrer600.94
Ultrasound60
H2SO4Mechanical stirrer60−0.25
Ultrasound60
Table 4. Chemical composition of homogenized cathode LIBs material.
Table 4. Chemical composition of homogenized cathode LIBs material.
ElementCoNiMnLiAl
wt%19.595.028.903.120.14
Table 5. Acidic strength of applied acids [8,60].
Table 5. Acidic strength of applied acids [8,60].
Type of AcidThe Dissociation Constants
H2SO4pKa = 1.26 × 10−2
C6H8O7pKa1 = 3.08
pKa2 = 4.74
pKa3 = 5.40
C2H2O4pKa1 = 1.23,
pKa2 = 4.19
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Willner, J.; Fornalczyk, A.; Gajda, B.; Figlus, T.; Swieboda, A.; Wegrzyński, D.; Mlonka, A.; Perenc, B.; Kander, M. Direct Sonochemical Leaching of Li, Co, Ni, and Mn from Mixed Li-Ion Batteries with Organic Acids. Energies 2024, 17, 4055. https://doi.org/10.3390/en17164055

AMA Style

Willner J, Fornalczyk A, Gajda B, Figlus T, Swieboda A, Wegrzyński D, Mlonka A, Perenc B, Kander M. Direct Sonochemical Leaching of Li, Co, Ni, and Mn from Mixed Li-Ion Batteries with Organic Acids. Energies. 2024; 17(16):4055. https://doi.org/10.3390/en17164055

Chicago/Turabian Style

Willner, Joanna, Agnieszka Fornalczyk, Bernadeta Gajda, Tomasz Figlus, Adam Swieboda, Dawid Wegrzyński, Aleksander Mlonka, Bartosz Perenc, and Michał Kander. 2024. "Direct Sonochemical Leaching of Li, Co, Ni, and Mn from Mixed Li-Ion Batteries with Organic Acids" Energies 17, no. 16: 4055. https://doi.org/10.3390/en17164055

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop