You are on page 1of 13

Hydrometallurgy 190 (2019) 105169

Contents lists available at ScienceDirect

Hydrometallurgy
journal homepage: www.elsevier.com/locate/hydromet

Recovery of metals from spent lithium-ion batteries using organic acids T



Jessica de Oliveira Demarco , Jéssica Stefanello Cadore, Franciele da Silveira de Oliveira,
Eduardo Hiromitsu Tanabe, Daniel Assumpção Bertuol
Environmental Processes Laboratory (LAPAM), Chemical Engineering Department, Federal University of Santa Maria (UFSM), Av. Roraima 1000, Santa Maria, RS
97105-900, Brazil

A R T I C LE I N FO A B S T R A C T

Keywords: An effective and environmentally friendly process was developed for recovering metals present in lithium-ion
LIB batteries (LIBs), using a mechanical process followed by heat treatment and leaching using malic, citric, and
Recycling formic acids at concentrations of 2 M. The techniques employed to characterize the process were TGA, DSC,
Leaching XRD, SEM, FT-IR, and EDXRF. The leaching was carried out using different experimental conditions of tem-
Valuable metal
perature, volume of hydrogen peroxide (H2O2), S/L ratio, and extraction time. The leachate solutions were
Organic acid
analyzed by FAAS. The characterization results showed that heat treatment at 700 °C for 2 h was effective for
degradation of the graphite and PVDF present in the LIBs. Over 90% of the Co, Li, and Mn present could be
extracted using the following conditions: 2 M DL-malic acid, 6% (v/v) H2O2, S/L of 1:20 (m/v), 95 °C, and
extraction time of 60 min. The process for recovering metals from spent LIBs using DL-malic acid could be
considered economically and environmentally correct, avoiding negative impacts in the environment and re-
covering metals with high added value that could be used in the manufacturing of new products.

1. Introduction Therefore, the recovery of these metals has a strong economic influence
on the development of a battery recycling process (Georgi-Maschler
In the last decades, the requirements for batteries as mobile power et al., 2012). Furthermore, the irresponsible disposal of spent LIBs, in
sources have steadily increased, largely due to the implementation of addition to failing to meet the requirement for sustainable use of va-
new technologies in the electronics industry, attractive designs for the luable materials, can lead to environmental contamination (Chen et al.,
consumer, and the wide range of equipment used on a daily basis. 2015).
Portable devices, such as mobile phones and microcomputers, make a Currently, pyrometallurgical and hydrometallurgical processes are
major contribution to the increased demand. Currently, lithium-ion the two routes commonly used to recover valuable metals from spent
batteries (LIBs) are the type most widely employed, with their use ex- LIBs (Al-Thyabat et al., 2013; Chen et al., 2015; Georgi-Maschler et al.,
pected to increase further as a result of their application in the auto- 2012). However, pyrometallurgical processes have disadvantages, such
motive sector (Freitas et al., 2010; Meshram et al., 2015). as the release of hazardous gases and high energy consumption. Fur-
The widespread and growing use of LIBs generates large quantities thermore, a hydrometallurgical process is usually needed in order to
of spent batteries that must be recycled by means of environmentally refine the waste and obtain purer forms, such as salts, hydroxides, and
friendly and economically viable processes. Ideally, closed-loop re- metals (Chen et al., 2015). Consequently, several studies have shown
cycling should provide materials for the production of new batteries. that use of a hydrometallurgical process can be advantageous for re-
LIBs contain high amounts of valuable metals, such as aluminum, covering valuable metals from spent LIBs (Lee and Rhee, 2003; Nayaka
copper, lithium, cobalt, nickel, and manganese, of which the most va- et al., 2016; Swain et al., 2007). However, it is necessary to use pro-
luable is cobalt (Co) (Georgi-Maschler et al., 2012). cesses that do not cause new environmental threats. One option is to use
Based on the assumption that the metal content can be completely organic acids, instead of the traditional leaching agents, since these
recovered in the metallic form, the monetary value of the Co present in acids are less corrosive and can be considered ecologically correct
1 ton of spent batteries is approximately US$ 7200 (Georgi-Maschler (Argenta et al., 2017).
et al., 2012). The value of Li is significantly lower, on average US$ 4530 The recovery of valuable metals from spent LIBs generally employs
per ton, but has increased since 2009 (U.S. Geological Survey, 2011). acid leaching in the presence of a reducing agent, which converts the


Corresponding author.
E-mail address: jessica-demarco@hotmail.com (J. de Oliveira Demarco).

https://doi.org/10.1016/j.hydromet.2019.105169
Received 20 December 2018; Received in revised form 9 May 2019; Accepted 14 June 2019
Available online 15 October 2019
0304-386X/ © 2019 Elsevier B.V. All rights reserved.
J. de Oliveira Demarco, et al. Hydrometallurgy 190 (2019) 105169

Table 1
Hydrometallurgical methods used for recovering cobalt and lithium from spent LIBs.
Reference Leaching agent H2O2 (v/v) T (°C) Time (min) S/L ratio (m/v) Leaching efficiency

Li et al., 2010a Citric acid (1.25 M) 1% 90 30 1:20 91% - Co


99% - Li
Golmohammadzadeh et al., 2017 Citric acid (2 M) 1.25% 60 300 1:30 96.46% - Co
99.8% - Li
Li et al., 2012 Ascorbic acid (1.25 M) – 70 20 1:25 94.8% - Co
98.5% - Li
Sun and Qiu, 2011 Sulfuric acid (2 M) 5% 80 60 1:50 99% - Co
99% - Li
Chen and Zhou, 2014 Citric acid (2 M) 2% 80 90 1:30 95% - Co
99% - Li
Zhang et al., 2015 Oxalic acid (1 M) – 95 150 1:15 97% - Co
98% - Li
Li et al., 2010b DL-malic acid (1.5 M) 2% 90 40 1:20 93% - Co
94% - Li
Pinna et al., 2017 Phosphoric acid (2%) 2% 90 60 1:8 99% - Co
88% - Li
Gao et al., 2018b Acetic acid (3.5 M) 4% 60 60 1:40 93.62% - Co
99.93% - Li

metals into a more soluble oxidation state (Bertuol et al., 2016). When a by dilution using deionized water.
strongly acidic solution is used for the leaching of Co and Li, > 99% (by
mass) of these metals can be recovered. However, Cl2, SO3, and NOx are
2.2. Dismantling and characterization of the LIBS
released during this process, while the acid residue obtained after
leaching also presents a risk to the environment (Li et al., 2010b). Be-
Mass characterization was performed by manually disassembling
cause of the growing interest in the sustainable management of natural
three LIBs of the same brand and model, with separation of the different
resources and in reducing environmental pollution, the recovery of
components including the external covering, aluminum case, cathode,
metals from spent LIBs is becoming increasingly important, because it
anode, polymeric material, aluminum foil (support for the LiCoO2
can help to alleviate potential environmental pressures and resolve the
cathode), and copper foil (support for the LiCγ anode). The mass of each
crisis of scarcity of Co and Li (Wang et al., 2016).
component was determined. The LiCoO2 powder was separated by
Due to the environmental concerns, it is necessary to separate and
scraping, followed by weighing.
recycle all constituents of LIBs, in order to recover the valuable metals
The LiCoO2 powder used in the leaching stages was obtained from
(Natarajan et al., 2018). The present work focuses on the recovery of Co
30 batteries of the same brand and model. The charge level of each
and Li. The main methods found in the literature for recovering these
battery was checked by measuring the voltage, in order to avoid the
metals are presented in Table 1.
possibility of short circuit and self-ignition. The batteries that had any
As shown in Table 1, Co and Li can be readily leached from LIB
charge were discharged in 10% Na2SO4 solution for 2 h, followed by
residues using dilute acid solutions. However, reducing agents such as
washing with deionized water and drying in an oven at 80 °C for 12 h.
hydrogen peroxide (H2O2) may be required in order to achieve dis-
The batteries were then comminuted in a hammer mill (Model A4,
solutions comparable to those obtained using higher concentration acid
Tiger) with a grid opening size of 5 mm, for release of the metals from
solutions (Ferreira et al., 2009). Furthermore, the studies shown in
the electrodes, with the aim of reducing the cost of the separation steps
Table 1 generally used processes different to those employed in the
(Silveira et al., 2017).
present work, such as manual separation of battery components and
The particles that passed through the grid were dried for 24 h in an
ultrasonic shaking, which may be impracticable on a large scale.
oven at 80 °C, for evaporation and elimination of the organic solvents
The aim of this study was to develop an effective and en-
present in the electrolyte of the batteries. The material was then sub-
vironmentally friendly technique for recovering metals present in LIBs,
jected to granulometric separation using vibrating sieves (Tyler 65,
using a mechanical process followed by heat treatment and leaching
with opening of 212 μm) to separate the fine fraction (LiCoO2 and
using malic, citric, and formic acids. Different characterization techni-
graphite) from the metal and polymer fractions. The particle size se-
ques were used to confirm the presence of the target metals in the LIBs,
paration increased the selectivity towards the active materials, hence
as well as to evaluate the efficiency of the heat treatment for removal of
decreasing the cost of subsequent separation steps (Lee and Rhee,
graphite and PVDF. In order to obtain the best process parameters,
2002).
evaluation was made of the effects of the solid/liquid ratio, volume of
Thermogravimetric analysis (TGA) was performed using a Shimadzu
hydrogen peroxide, temperature, and extraction time.
TGA-50 instrument, obtaining the change in mass as a function of the
thermal degradation of the sample. The sample was analyzed in an
2. Experimental atmosphere of synthetic air at a flow rate of 50 mL.min−1, with heating
from ambient temperature to 1000 °C, at a rate of 10 °C.min−1.
2.1. Materials and reagents Differential scanning calorimetry (DSC) was used to observe changes
during heating of the sample from room temperature to 600 °C, at a rate
A total of 33 batteries of the same brand and model were collected of 10 °C.min−1, under a flow of nitrogen gas supplied at 50 mL.min−1.
from a mobile phone maintenance company. The leaching procedures Samples of the fraction with granulometry < 212 μm were sub-
employed hydrochloric acid (HCl, Synth), nitric acid (HNO3, Synth), jected to pretreatment in a muffle furnace at 600, 650, and 700 °C,
anhydrous sodium sulfate (Na2SO4, Synth), anhydrous DL-malic acid during 2 h for each temperature, in order to eliminate graphite and
(C4H6O5, Synth), anhydrous citric acid (C6H8O7, Proquimios), and an- organic compounds.
hydrous formic acid (CH2O2, Êxodo Científica). All these reagents were The crystalline structures of the original material and the samples
analytical grade. Hydrogen peroxide (H2O2, 30%, Synth) was used as a obtained after the different thermal treatments were characterized by
reducing agent. The solutions were prepared at different concentrations X-ray diffraction (XRD), using a Rigaku Miniflex 300 instrument

2
J. de Oliveira Demarco, et al. Hydrometallurgy 190 (2019) 105169

operated in the 2θ range 5–98°, with Cu Kα radiation (γ = 1.5418 А), 2010b; Chen and Zhou, 2014; Golmohammadzadeh et al., 2017). The
voltage of 30 kV, and current of 10 mA. leachate solutions were separated from the solid material by filtration
Analysis by scanning electron microscopy (SEM) with energy dis- and were analyzed by flame atomic absorption spectrometry (FAAS),
persive spectroscopy (Vega-3G, Tescan) was performed in order to using an Agilent 240 FS instrument.
obtain the morphological characteristics and compositions of the sam- The first step was to determine the best S/L ratio for the process,
ples before and after the heat treatment. For this analysis, the samples using ratios of 1:10, 1:20, 1:30, and 1:50 (m/v), while the other vari-
were metalized by sputter-coating with gold (using a current of 20 mA ables remained fixed. The best ratio was then used in the second stage,
for 90 s) and were immobilized using Ni/Cu tape. evaluating the influence of H2O2 at concentrations of 0, 1, 2, 4, and 6%.
Analyses by Fourier transform infrared spectrometry (FTIR), em- In the third step, the effect of temperature on extraction of the desired
ploying a Shimadzu Prestige 21 instrument, were used to confirm the metals was evaluated using temperatures of 55, 65, 75, 85, and 95 °C.
effectiveness of the heat treatment for removal of PVDF. Semi-quanti- Finally, the effect of the leaching time was investigated using times of
tative characterization of the elements present in the samples was 30, 60, 120, 180, and 240 min, while keeping the other variables at
performed by energy dispersive X-ray fluorescence (EDXRF), using a their optimum values.
Bruker S2 Puma instrument.

3. Results and discussion


2.3. Recovery of metals
3.1. Characterization of the lithium-ion batteries
The system used for the leaching experiments consisted of a heating
plate with a magnetic stirrer operated at 300 rpm, which was sub- 3.1.1. Mass balance
merged in a water bath and connected to a recirculation cooling system Fig. 2 shows the different constituents present in a LIB (a) after
composed of a reflux condenser and ultrathermostatic bath at 5 °C, in manual separation: polymeric casing (b); metallic casing (c); connectors
order to avoid losses by evaporation. All the leaching stages were making the electrical contacts of the battery (d); anode composed of a
performed with the material that had been submitted to heat treatment. copper electrode covered with layers of graphite (e); polymer separator
The sequence of these steps is shown in Fig. 1. between the cathode and the anode (f); cathode composed of an alu-
Aqua regia (HCl:HNO3, 3:1 ratio) was used for complete digestion of minum electrode covered by LiCoO2 (g); LiCoO2 powder after scraping
the LIB powder prior to quantification of the metal contents, using (h).
1:50 S/L ratio, temperature of 90 °C, duration of 120 min, and agitation Table 2 shows the masses of the components of three batteries. The
at approximately 300 rpm (Li et al., 2012; Wang et al., 2016; Gao et al., total masses of the batteries before dismantling were from 20.12 to
2017; Golmohammadzadeh et al., 2017). 21.47 g. The cathode mass was between 8.24 and 9.6 g, containing from
The leaching processes using DL-malic acid, formic acid, and citric 3.05 to 5.94 g of active LiCoO2 material. As pointed out by Georgi-
acid were performed with the acids at concentrations of 2 M. The ex- Maschler et al. (2012), battery producers produce their own specific
periments were conducted using different conditions of temperature, types of LIBs, so it is difficult to provide accurate general values for the
volume of hydrogen peroxide (H2O2), S/L ratio, and extraction time, masses of the components, because the composition varies according to
which were selected based on studies reported in the literature (Li et al., the manufacturing process (Bertuol et al., 2016). As can be observed

Fig. 1. Processes carried out to recover the metals from the spent LIBs.

3
J. de Oliveira Demarco, et al. Hydrometallurgy 190 (2019) 105169

Fig. 2. Different components of a dismantled spent lithium-ion battery: a) lithium-ion battery; b) polymeric casing; c) metallic casing; d) connectors; e) anode; f)
polymeric separator; g) cathode; h) LiCoO2 powder.

here, even batteries of the same brand and model can differ in terms of the cathodic conductor and the active materials. During these reactions,
their mass composition. the metals in the active cathode materials were reduced from a high
The prior segregation of battery components would be very profit- charge to a low charge state. The reduction reactions of the transition
able for recycling, considering that the plastic case, the polymer foil, metals make their leaching easier and more efficient (Li et al., 2012;
the solvent, the steel case, the electrical contacts, and the aluminum Yang et al., 2016). A maximum mass loss peak was present at around
and copper foils are all directly recyclable after separation (Paulino 700 °C. According to Kurajica et al. (2011), the reduction of Co3+ to
et al., 2008). On average, these components constitute 65% (by weight) Co2+ is thermodynamically favored at temperatures higher than
of the spent battery. However, the manual separation of the different 700 °C. The heat treatment eventually causes breakdown of the metal
components would be impracticable on a large scale. Therefore, the oxide cathode material, releasing oxygen. The positively charged active
mechanical processes were carried out using whole batteries and the materials can disproportionate at elevated temperatures, as shown in
segregation of components was performed by particle size separation. Eqs. (1) and (2) (Wang et al., 2012; Diaz et al., 2019).
1 1
Li x CoO2 → xLiCoO2 + (1 − x ) Co3 O4 + (1 − x ) O2
3.1.2. Evaluation of calcination processes and chemical composition 3 3 (1)
The techniques described below were used to characterize both the
LIBs powder and the efficiency of the heat treatment for removal of Co3 O4 → 3CoO + 0.5O2 CoO → Co + 0.5O2 (2)
graphite and PVDF. A weight loss of 1.7% wt. was observed from room temperature to
275 °C, with a corresponding endothermic peak, due to the loss of water
3.1.2.1. Thermogravimetric analysis (TGA) and differential scanning bound in the material (Li et al., 2012). A small endothermic peak be-
calorimetry (DSC). The results of the TGA and DSC analyses of the tween 197 and 215 °C could be attributed to the melting of LiPF6 (Yang
LIB powder are presented in Fig. 3. The thermogravimetric analysis et al., 2006). The decomposition of LiPF6 is represented by Eq. 3
curves showed that there were significant mass losses when the (Ravdel et al., 2003), where the second product is a strong Lewis acid
temperature was ramped from 50 to 800 °C. that can react with water to form HF and POF3 (Genieser et al., 2018).
The first mass decrease, below 500 °C, could be attributed to the
LiPF6 ↔ LiF + PF5 (3)
formation of H2O and CO2 in the gas phase (Li et al., 2012). A sig-
nificant mass loss of 22.8% occurred from 580 to 845 °C, probably as- A mass loss of 10.8% from 275 to 580 °C was associated with several
sociated with degradation of the graphite and redox reactions between exothermic peaks. The mass loss above 500 °C could be explained by the

Table 2
Characterization of the masses of the LIB components.
Component Battery 1 Battery 2 Battery 3

Mass (g) Fraction (%) Mass (g) Fraction (%) Mass (g) Fraction (%)

Polymeric casing 0.39 1.94 0.36 1.68 0.22 1.03


Metallic casing 2.11 10.49 2.37 11.08 3.03 14.11
Connectors 1.49 7.41 1.54 7.20 1.16 5.40
Anode 4.08 20.28 4.12 19.25 4.63 21.56
Cathode 8.78 43.64 8.24 38.50 9.60 44.72
Polymeric separator 0.59 2.93 0.71 3.32 0.81 3.77
Aluminum foil 5.73 28.48 3.42 15.98 3.66 17.05
LiCoO2 3.05 15.16 4.82 22.52 5.94 27.67
Copper foil 1.72 8.55 1.47 6.87 2.06 9.59
Graphite 2.36 11.73 2.65 12.38 2.57 11.97
Electrolyte 2.68 13.31 4.06 18.97 2.02 9.41
Ʃ 20.12 100 21.40 100 21.47 100

4
J. de Oliveira Demarco, et al. Hydrometallurgy 190 (2019) 105169

Fig. 3. (a) Thermogravimetric analysis (TGA) and (b) differential scanning calorimetry (DSC) analysis of the LIB powder.

Table 3
Semi-quantitative analysis of the LIB powder by
EDXRF.
Element Weight (%)

Al 1.39 ( ± 0.18)
Mn 51.87 ( ± 0.18)
Fe 2.83 ( ± 0.03)
Co 34.15 ( ± 0.32)
Li 3.97 ( ± 0.29)
Ni 3.35 ( ± 0.10)
Cu 0.45 ( ± 0.04)
P 1.8 ( ± 0.17)
Others 0.19 ( ± 0.11)

Fig. 4. X-ray diffractograms of the LIB powder before and after heat treatment
at different temperatures. Fig. 5. FTIR spectra obtained before (a) and after (b) heat treatment of the LIB
powder at 700 °C.
thermal degradation of PVDF, which decomposes at around 500 °C
(Yang et al., 2016). the carbon electrode is lithiated. PVDF is dehydrofluorinated according
As reported by Wang et al. (2012), the reactions do not proceed in a to Eq. (4). A possible reaction between the binder and the LixC6 elec-
precise order, since they can influence each other, hence inducing er- trode is represented by Eq. (5) (Wang et al., 2012).
ratic behavior. For example, the PVDF–LixC6 reactions are strongly af-
fected by the degree of lithiation of the graphite, only occurring when − CH2 − CF2 → −CH = CF − +HF (4)

5
J. de Oliveira Demarco, et al. Hydrometallurgy 190 (2019) 105169

Fig. 6. SEM micrographs obtained before (a) and after (b) heat treatment at 700 °C (magnification of 2500×).

− CH2 − CF2 + Li → LiF + −CH = CF − +0.5H2 (5) and was added to the EDXRF results. The Mn was probably derived
The gas generated in the pretreatment of LIBs can be removed using from the LiMn2O4 complex, which is used in some cathodes in order to
activated carbon. Adsorption is considered one of the most viable op- reduce the production costs of batteries, since Co is expensive (Deboer
tions that can be applied in industrial flue gas treatment processes and Lammertsma, 2013; Swart et al., 2014; Huang et al., 2016; Xin
(Shafeeyan et al., 2011). Among several adsorbents, carbonaceous et al., 2016; Xiao et al., 2017). Therefore, considering the high Mn
materials show excellent potential because of their large surface area, content in the sample, the extraction efficiency was evaluated for this
plentiful pore structure, high stability, and relatively low cost. Fur- metal, as well as for Co and Li.
thermore, most of the material adsorbed onto the carbon materials can Increasing attention has been given to lithium manganese oxides,
be recycled by means of desorption processes performed under various due to their high energy density, low cost, and environmentally friendly
conditions (Zhang et al., 2017). characteristics. Fisher et al. (2013) developed film cathodes for lithium-
ion batteries using a Li–Mn–O system with Mn mass contribution of
53.2%. Here, the presence of phosphorus suggested that the electrolyte
3.1.2.2. Characterization by X-ray diffraction. Fig. 4 shows X-ray
was lithium hexafluorophosphate (LiPF6) (Silva et al., 2018). The re-
diffractograms of the cathodic and anodic active materials (powders)
sults shown are averages of three replicates.
of the batteries without heat treatment and after heat treatment. For the
Dutta et al. (2018) analyzed the composition of the cathodic ma-
sample without calcination, all the XRD peaks could be indexed to
terial present in different brands and models of LIBs and obtained mass
LiCoO2 and graphite (Silveira et al., 2017; Bertuol et al., 2016; Zhang
concentrations of 20% for Co, 2.4% for Li, 2.25% for Mn, 1.2% for Al,
et al., 2014; Ferreira et al., 2009). The graphite in the sample was from
0.3% for Fe, and 0.8% for Ni, after comminution of the batteries and
the anode and the LiCoO2 was from the cathode.
centrifugation to separate the polymer, metal, and LiCoO2 powder
After two hours of calcination, when the temperature reached about
fractions. However, the analysis employed a sample with 90% of the
700 °C, the absence of the graphite peaks indicated that the graphite
particles smaller than 1040 μm, while the present work employed
had combusted. Hence, 700 °C was considered the ideal temperature for
particles smaller than 212 μm.
the calcination prior to the leaching step. This was in agreement with
Natarajan et al. (2018) studied the metals content of the cathodic
the TGA data, corroborating the results of Zheng et al. (2018) and en-
material and noted that the composition of LIBs can differ according to
abling a shorter heat treatment time, compared to several earlier stu-
the manufacturer, obtaining values of 5.1% Li, 26.3% Co, and 28.2%
dies (Li et al., 2010b; Li et al., 2012; Golmohammdadzadeh et al.,
Mn. Differently, Ku et al. (2016) reported a composition of 15.3% Ni,
2017).
14.3% Mn, 6.0% Co, 2.3% Al, 0.3% Cu, and 61.8% other elements.
3.1.2.3. Energy dispersive X-ray fluorescence (EDXRF). The residues of
LIBs contain many valuable metals, such as Cu, Al, Fe, Li, Co, Mn, and 3.1.2.4. Fourier transform infrared spectrometry (FT-IR). Fig. 5 shows the
Ni (Huang et al., 2016; Xiao et al., 2017). Table 3 shows the chemical FT-IR spectra of the LIB powder before and after heat treatment at
composition of the LIB powder with granulometry < 212 μm, after the 700 °C. A broad absorption peak between 3600 and 3000 cm−1 could
heat treatment at 700 °C. The results showed that after comminution in be attributed to OeH stretching vibration, while a peak at 2933 cm−1
a hammer mill and particle size separation, Mn was the main element was associated with stretching vibrations of methyl and methylene
present in the LIB powder (51.87%), followed by Co (34.15%), Li groups. A peak at 1644 cm−1 corresponded to vibration of C]C in
(3.97%), Ni (3.35%), Fe (2.83%), and Al (1.39%). The Li concentration H2C=CF–R groups. Peaks at 1394 cm−1 and between 1194 and
was obtained by FAAS, due to the low molecular weight of the element, 866 cm−1 indicated the presence of the fluorocarbon group from the

6
J. de Oliveira Demarco, et al. Hydrometallurgy 190 (2019) 105169

Fig. 7. Effect of the S/L ratio on the leaching efficiencies for (a) Co, (b) Li, and (c) Mn, using 2 M acid, 1% (v/v) H2O2, and 65 °C, for 60 min.

PVDF binder (Sun and Qiu, 2011; Pant and Dolker, 2017). After the 3.2. Recoveries of the metals
heat treatment, these absorption peaks attributed to the PVDF binder
disappeared, confirming the effectiveness of the heat treatment at The recoveries of the metals were evaluated varying the S/L ratio,
700 °C for the removal of organic fluorocarbon compounds. H2O2 volume, temperature, and extraction time. The recovery achieved
The results of Fourier transform infrared spectroscopy analysis re- using extraction with aqua regia (3 HCl: 1 HNO3) was assumed to be
ported by Hanisch et al. (2015) indicated the production of unwanted 100% for all the metals.
compounds such as HCN, HF, CH4, HCHO, COF2, HNCO, higher hy-
drocarbons, nitrogen oxides, CO, and CO2, among others, when elec-
trodes were incinerated in air. These byproducts must not be released to 3.2.1. Effect of the solid/liquid ratio
the environment, since they may be hazardous to humans, the en- The effect of the solid/liquid ratio was studied in the range from
vironment, and equipment. Adsorptive and absorptive methods such as 1:10 to 1:50 (m/v), while keeping the other parameters constant at 2 M
the use of activated carbon and gas scrubbers can prevent the release of acid, 1% (v/v) H2O2, temperature of 65 °C, and time of 60 min. The
these compounds into the atmosphere. leaching efficiencies obtained for Co, Li, and Mn are shown in Fig. 7.
The results (Fig. 7) indicated that S/L of 1:20 was most suitable for
recovering the metals from the LiCoO2 powder, in agreement with the
3.1.2.5. Characterization by scanning electron microscopy with energy results reported by Sun and Qiu (2011), Chen et al. (2017), and Gao
dispersive spectroscopy (SEM/EDS). Morphological analysis using et al. (2017). It can be seen that the leaching efficiencies for Co, Li, and
scanning electron microscopy was used to observe the microstructural Mn did not increase significantly when lower S/L ratios were used. At a
changes caused by the heat treatment of the different samples (Fig. 6). low S/L ratio, the leaching solution included abundant water mole-
Prior to the heat treatment, the particles formed agglomerates (Fig. 6a), cules, so the organic acids became increasingly weak. At a high S/L
while after the heat treatment, these agglomerates were broken up, ratio, there was weaker ionization capacity and higher acid con-
resulting in dispersed particles (Fig. 6b). This provided further evidence centration. These characteristics provided an explanation for the fact
of the effectiveness of the calcination at 700 °C during 2 h, facilitating that the leaching efficiencies of Co and Li initially increased and then
contact with the leaching agent. decreased, as the S/L ratio was reduced (Zheng et al., 2018).
The use of S/L of 1:20 resulted in a higher concentration gradient of
H+ between the solid-liquid interface and the solution, which promoted
ion transfer in the solution, leading to higher leaching efficiency. With

7
J. de Oliveira Demarco, et al. Hydrometallurgy 190 (2019) 105169

Fig. 8. Effect of H2O2 volume on the leaching efficiencies for (a) Co, (b) Li, and (c) Mn, using 2 M acid, S/L ratio of 1:20, and 65 °C, during 60 min.

further addition of the acids, the effects of diffusion velocity and visc- While hydrogen peroxide assisted the dissolution of Co, the dissolution
osity became weaker, with the leaching efficiency decreasing for Li and of Li was also promoted, because the two metals were contained in the
not changing significantly for Co and Mn (Guo et al., 2016). same oxide compound (Shin et al., 2005).
Therefore, S/L of 1:20 was considered ideal from both economic and The use of 1% H2O2 with DL-malic acid resulted in Mn extraction of
extraction efficiency perspectives, providing Co extraction of 30.63% 98.90% (Fig. 8(c)). However, while the Co extraction increased pro-
for DL-malic acid, 20.78% for formic acid, and 43.39% for citric acid. gressively using H2O2 volumes of 2, 4, and 6%, the Mn extraction de-
For Li, the extraction efficiencies were 95.77, 63.25, and 82.18%, re- creased from 98.90 to 85.01%. It could be concluded from these results
spectively, while for Mn the efficiencies were 98.90, 18.81, and 19.09% that the use of 1% H2O2 resulted in easier dissolution of the LiMn2O4
for DL-malic acid, formic acid, and citric acid, respectively. As can be complex, compared to the LiCoO2 complex (Zhang et al., 2018).
seen from the results, Li was leached more easily than Co. This could be When formic acid was used together with 1% (v/v) H2O2, the Co
explained by the lamellar structure of LiCoO2 and the different dis- leaching efficiency increased from 1.19 to 20.78%, while the use of 1%
tribution patterns of Co and Li in the cathodic materials (Li et al., 2017; (v/v) H2O2 with citric acid increased the efficiency from 21.62 to
Chen et al., 2018a). 43.39%. When the H2O2 volume was increased above 1%, the efficiency
decreased. Therefore, the use of H2O2 at a concentration of 1% was
3.2.2. Effect of the hydrogen peroxide volume most suitable for both acids.
The effect of the H2O2 volume on the extraction of the metals was According to Zheng et al. (2018), Co3+ is reduced to Co2+ with
evaluated in the range 0–6% (Fig. 8). A volume of 8% was also used for H2O2. However, when the dosage of the reducing agent is increased
the extraction of Co and Li using DL-malic acid, because the leaching (Fig. 8), the molecules of H2O2 break down, due to the excess of H2O2,
efficiency increased as the H2O2 volume was increased in the range so the agent changes from being a reducer to being an oxidant (Eq. 6).
0–6%. The other parameters were kept constant at 2 M acid con-
H2 O2 (aq) = H2 O (aq) + O2 (g) (6)
centration, temperature of 65 °C, time of 60 min, and S/L ratio of 1:20.
As shown in Fig. 8(a), the leaching efficiency for Co using DL-malic The use of 1% (v/v) H2O2 with citric acid resulted in extraction
acid increased when the H2O2 volume was increased from 0 to 6%. The efficiencies of 82.18% for Li and 19.09% for Mn. The same volume of
chemical bond between Co and oxygen is strong, making acid leaching H2O2 in formic acid provided leaching efficiencies of 63.28% for Li and
of LiCoO2 difficult. When hydrogen peroxide was added, the oxygen 18.81% for Mn.
produced from the decomposition of H2O2 converted Co3+ to Co2+, The use of 6% (v/v) H2O2 with DL-malic acid resulted in leaching
which assisted the dissolution (Saeki et al., 2004; Li et al., 2010a). efficiencies of 67.21% for Co, 99.30% for Li, and 85.01% for Mn, so this

8
J. de Oliveira Demarco, et al. Hydrometallurgy 190 (2019) 105169

Fig. 9. Effect of temperature on the leaching efficiencies for (a) Co, (b) Li, and (c) Mn, using 2 M acid, S/L ratio of 1:20, and H2O2 volumes of 6% with DL-malic acid
and 1% with formic and citric acids, during 60 min.

volume of H2O2 was defined as being most suitable for use with DL- hydrogen bonds. Hence, the hydrogen bonds between the organic acids
malic acid. and the metals were weakened at temperatures above 65 °C (Yu et al.,
The leaching efficiencies of the metals observed in the absence of 2018).
H2O2 could be explained by the fact that Li+ and Co2+ ions can be For Li, the highest extraction efficiencies obtained using formic and
readily leached by DL-malic acid, while Co3+ and Mn4+ require the citric acids were 63.25 and 82.18%, respectively, at 65 °C. In the case of
presence of a reducing agent (Yao et al., 2015; Zhang et al., 2015; Sun DL-malic acid, the highest efficiency was 99.33% at 65 °C, while lower
et al., 2017). extraction of 93.22% was obtained at 95 °C. Increase of the temperature
from 55 to 65 °C (Fig. 9) resulted in the recovery of Li increasing, while
the leaching efficiency decreased, due to decomposition of the Li
3.2.3. Effect of temperature
complex at higher temperatures (Golmohammadzadeh et al., 2018).
The effect of temperature was studied while maintaining the other
In the case of Mn, the highest leaching efficiency was 99.53%, using
parameters constant at S/L of 1:20, H2O2 volumes of 6% with DL-malic
DL-malic acid at 95 °C, while values of 19.09 and 39.72% were obtained
acid and 1% with citric and formic acids, and extraction time of 60 min.
for citric acid at 65 °C and formic acid at 75 °C, respectively. However,
The leaching efficiencies obtained are shown in Fig. 9.
since Co is the most valuable metal present in LIBs, the recovery effi-
When the temperature was increased to 95 °C, 90.57% extraction of
ciency for this metal has a strong economic impact on the development
Co was achieved using DL-malic acid, indicating that under these
of a suitable battery recycling process (Georgi-Maschler et al., 2012).
conditions, the higher temperature accelerated the leaching, according
Therefore, the temperatures selected for recovering the metals from
to a mechanism of homogeneous corrosion of the particles (Joulié et al.,
LIBs were 95 °C for malic acid and 65 °C for formic and citric acids.
2014). The ability of cobalt to form chelates with organic acids is based
on the potential required for Co3+ to Co2+ conversion. The chelate
formation reaction of cobalt with this organic acid is favored by the 3.2.4. Effect of leaching time
dissociation of DL-malic acid at high temperatures (Li et al., 2014; Evaluation of the effect of the leaching process time was performed
Golmohammadzadeh et al., 2017). This does not occur for formic and in experiments in which the other parameters were kept constant at the
citric acids, so the Co leaching efficiency did not increase, with max- best values identified previously: S/L ratio of 1:20, H2O2 volumes of 6%
imum efficiencies of 20.78 and 43.39%, respectively, obtained at 65 °C. with DL-malic acid and 1% with formic and citric acids, and tempera-
The kinetic energies of molecules containing Li and Co generally in- tures of 95 °C with malic acid and 65 °C with formic and citric acids.
crease with increasing temperature, leading to the ability to break down The results are shown in Fig. 10.

9
J. de Oliveira Demarco, et al. Hydrometallurgy 190 (2019) 105169

Fig. 10. Effect of leaching time on the extraction efficiencies of (a) Co, (b) Li, and (c) Mn, using 2 M acid, S/L ratio of 1:20, 6% (v/v) H2O2 and 95 °C with DL-malic
acid, and 1% (v/v) H2O2 and 65 °C with formic and citric acids.

respectively, and for Mn were 99.53, 18.82, and 19.09%, respectively.


0 Extending the leaching time for longer than 60 min did not increase the
H+ extraction of Co and Mn, which could be attributed to the instability of
H2O2 and its decomposition to H2O and O2 (Chen and Zhou, 2014). It is
OH also important to mention the possibility that Mn could react with the
-2
O2 released by the decomposition of H2O2, leading to the precipitation
of Mn, which is feasible according to the diagram shown in Fig. 11.
Log Conc.

-4
There have been few studies reporting the behavior of Mn with these
Mn3+
acids. In most cases, the extraction times used were longer than 1 h,
Mn2O3(cr) which highlights the relevance of the present study. In the work of Chen
et al. (2018b), it was found that an extended leaching time could result
-6
in lower extraction of Li, Co, and Mn.
The Pourbaix diagrams for Co, Li, and Mn are provided in Fig. 12,
showing the possible equilibrium phases of the aqueous electrochemical
-8
systems.
The pH values fluctuated in the ranges 2.1 to 2.98 for DL-malic acid,
2 4 6 8 10 12 0.14 to 3.28 for formic acid, and 1.81 to 2.59 for citric acid. The po-
pH
tential values varied from 0.4579 to 0.4909, from 0.4299 to 0.6149,
Fig. 11. Speciation diagram for Mn. Source: HYDRA-MEDUSA speciation dia- and from 0.4689 to 0.5119 for DL-malic, formic, and citric acids, re-
grams (available from: https://www.kth.se/che/medusa/downloads-1. spectively. Hence, considering the Pourbaix diagrams presented in
386254). Fig. 12, all the leaching stages were in the corrosion area, enabling the
recovery of metals by dissolution.
As can be seen in Fig. 10, the Co, Li, and Mn leaching efficiencies However, the variations in the physicochemical properties of dif-
tended to increase and then decrease, as the duration of the process was ferent acid solutions could lead to difficulties in establishing the best
extended. Leaching for 60 min resulted in extraction efficiencies for Co parameters for the recycling of metals from LIBs. The characteristics of
of 90.57, 20.67, and 43.39%, using DL-malic, formic, and citric acids, the interactions between the materials and the acid solutions, the var-
respectively. The values for Li were 93.22, 63.25, and 82.18%, iation in the physicochemical properties of the solutions, and the

10
J. de Oliveira Demarco, et al. Hydrometallurgy 190 (2019) 105169

(a) (b)
1.0 CoO 2 (cr) 1.0

Co(OH)3(s)

0.5 0.5
Co 2+

/ V
/ V

0.0 Co(OH ) 2 (s) 0.0 L i+

SHE
SHE

E
E

-0.5 -0.5
Co(cr)

-1.0 -1.0
2 4 6 8 10 12 2 4 6 8 10 12
pH pH

(c)
1.0
M nO 4−
M nO 2 (cr)

0.5
M n 2O 3(cr)
/ V

M n 3O 4 (s)

0.0
SHE

M n 2+
E

-0.5 M n(OH)2

-1.0
2 4 6 8 10 12
pH
Fig. 12. Pourbaix diagrams for (a) Co, (b) Li, and (c) Mn. Source: HYDRA-MEDUSA speciation diagrams (available from: https://www.kth.se/che/medusa/
downloads-1.386254).

relationship between the leaching parameters are critical aspects of a in the presence of citric acid, compared to Co. In the leaching reaction,
comprehensive evaluation of the effectiveness of a leaching process the Co and Li were leached as Co(C6H7O7)2, Li(C6H7O7), Co2+, and Li+.
(Gao et al., 2018a). The possible reactions that may have occurred The residues were Co3O4 and C, which could not be leached because
during the leaching processes are described below. Co3O4 does not dissolve completely in citric acid (Li et al., 2010b).
Citric acid presented extraction efficiency below 50%, showing that Formic acid also showed low extraction efficiency. In addition to
it was not an ideal leaching agent for this process. The leaching reaction being a weak acid, formic acid is a reducer of ions with high oxidative
of the LiCoO2 residue with a solution of citric acid can be represented potential, due to its aldehyde. Unlike other acids, formic acid can se-
by Eqs. (7)–(9) (Li et al., 2010b): lectively leach aluminum to obtain high purity aluminum, while nickel,
cobalt, and manganese can precipitate as hydroxides. The chemical
6H3 Cit(aq) + 2LiCoO2 (s) + H2 O2 (aq) reactions that could occur during the leaching process with formic acid
= 2Li+ (aq) + 6H2 Cit− (aq) + 2Co2 + (aq) + 4H2 O + O2 (g) (7) are described by Eqs. 10 and 11 (Gao et al., 2017):

2Al(s) + 6HCOOH(aq) → 2C3 H3 AlO6 (aq) + 3H2 (g) (10)


6H2 Cit− (aq) + 2LiCoO2 (s) + H2 O2 (aq)
= 2Li+ (aq) + 2Co2 + (aq) + 6HCit2 − (aq) + 4H2 O + O2 (g) (8) 6LiNi1/3Co1/3 Mn1/3O2 (s) + 21HCOOH (aq) → 2C2 H2 NiO4
(aq) + 2C2 H2 CoO4 (aq) + 2C2 H2 MnO4
6HCit2 − (aq) + 2LiCoO2 (s) + H2 O2 (aq)
(aq) + 6CHLiO2 (aq) + 3CO2 (g) + 12H2 O
= 2Li+ (aq) + 2Co2 + (aq) + 6Cit3 − (aq) + 4H2 O + O2 (g) (9)
(aq) (11)
Eqs. (7)–(9) suggest that the leaching efficiencies of Li and Co de-
pend on the concentration of H2O2. Moreover, Li dissolves more readily The leaching of Co may have decreased over time due to the

11
J. de Oliveira Demarco, et al. Hydrometallurgy 190 (2019) 105169

formation and precipitation of hydroxide. Furthermore, the solubility of spent LIBs. It was found that over 90% of these metals could be leached
Co in formic acid is relatively low, compared to the solubility of Li. using a solution of 2 M DL-malic acid, 6% (v/v) H2O2, S/L of 1:20 (m/
Hence, formic acid is not indicated as a leaching agent for Co when v), temperature of 95 °C, and leaching time of 60 min.
entire LIBs are comminuted in hammer mills, because a large number of The process for the recovery of metals from spent LIBs using DL-
other constituents are present, such as aluminum, which may have a malic acid could be considered economically and ecologically correct,
significant influence on the leaching of Co and Li. However, this acid avoiding negative impacts in the environment and recovering metals
can be used for the selective extraction of aluminum, consequently fa- with high added value that could be used in the manufacturing of new
cilitating the extraction of Co and Li (Gao et al., 2017). products.
The leaching reaction between LiCoO2 and malic acid is a multi-
phase process determined by the chemical reaction and by the transfer Acknowledgments
of ions in the solution. The presence of a reducing agent during the
leaching of LiCoO2 with malic acid can facilitate the direct reaction, Financial support for this work was provided by the Brazilian
since the Co3+ is reduced to Co2+. The chemical reaction can be re- agencies CAPES (Coordenação de Aperfeiçoamento de Pessoal de Nível
presented by Eqs. 12 and 13 (Li et al., 2010a): Superior), CNPq (Conselho Nacional de Desenvolvimento Científico e
Tecnológico), FAPERGS (Fundação de Amparo à Pesquisa do Estado do
2LiCoO2 (s) + 6C4 H6 O5 (aq) + H2 O2 → 4LiC4 H5 O5 (aq) + 2Co Rio Grande do Sul), and SDECT (Secretaria de Desenvolvimento
(C4 H5 O5 )2 (aq) + 4H2 O(l) + O2 (g) (12) Econômico, Ciência e Tecnologia do Rio Grande do Sul).

2LiCoO2 (s) + 6C4 H5 O− + 2+


5 (aq) + 2Li (aq) + 2Co (aq) + H2 O2 References
→ 2Li2 C4 H 4 O5 (aq) + 4CoC4 H 4 O5 (aq) + 4H2 O (l) + O2 (g) (13)
Al-Thyabat, S., Nakamura, T., Shibata, E., Iizuka, A., 2013. Adaptation of minerals pro-
The rates of both the chemical reaction and the ion transfer are cessing operations for lithium-ion (LiBs) and nickel metal hydride (NiMH) batteries
recycling: critical review. Miner. Eng. https://doi.org/10.1016/j.mineng.2012.12.
significantly affected by the temperature (Li et al., 2010a). At lower
005.
temperatures, the leaching is determined by the chemical reaction, for Argenta, A.B., Reis, C.M., Mello, G.P., Dotto, G.L., Tanabe, E.H., Bertuol, D.A., 2017.
which the rate increases with increase of the temperature. Hence, the Supercritical CO2extraction of indium present in liquid crystal displays from dis-
progress of the leaching is determined by the ionic transfer. At 95 °C, carded cell phones using organic acids. J. Supercrit. Fluids. https://doi.org/10.1016/
j.supflu.2016.10.014.
leaching using DL-malic acid for 60 min resulted in over 90% extraction Bahaloo-Horeh, N., Mousavi, S.M., 2017. Enhanced recovery of valuable metals from
of Co, Li, and Mn, showing the suitability of this acid for the recovery of spent lithium-ion batteries through optimization of organic acids produced by
metals from spent LIBs. Aspergillus niger. Waste Manag. https://doi.org/10.1016/j.wasman.2016.10.034.
Bertuol, D.A., Machado, C.M., Silva, M.L., Calgaro, C.O., Dotto, G.L., Tanabe, E.H., 2016.
DL-malic acid degrades more easily under aerobic and anaerobic Recovery of cobalt from spent lithium-ion batteries using supercritical carbon dioxide
conditions, compared to HCl, HNO3, and H2SO4. Furthermore, this acid extraction. Waste Manag. https://doi.org/10.1016/j.wasman.2016.03.009.
can be recycled and reused for subsequent leaching (Li et al., 2010a; Chen, X., Zhou, T., 2014. Hydrometallurgical process for the recovery of metal values
from spent lithium-ion batteries in citric acid media. Waste Manag. Res. https://doi.
Bahaloo-Horeh and Mousavi, 2017). Therefore, the process using DL- org/10.1177/0734242X14557380.
malic acid to recover metals from spent LIBs can be considered eco- Chen, X., Chen, Y., Zhou, T., Liu, D., Hu, H., Fan, S., 2015. Hydrometallurgical recovery of
nomically and ecologically advantageous, avoiding adverse impacts in metal values from sulfuric acid leaching liquor of spent lithium-ion batteries. Waste
Manag. https://doi.org/10.1016/j.wasman.2014.12.023.
the environment and recovering metals with high added value that can
Chen, X., Ma, H., Luo, C., Zhou, T., 2017. Recovery of valuable metals from waste cathode
be used in the manufacturing of new products. If not treated appro- materials of spent lithium-ion batteries using mild phosphoric acid. J. Hazard. Mater.
priately, spent LIBs are classified as hazardous waste that can cause https://doi.org/10.1016/j.jhazmat.2016.12.021.
Chen, X., Guo, C., Ma, H., Li, J., Zhou, T., Cao, L., Kang, D., 2018a. Organic reductants
harm to the environment, animals, and human health (Ordoñez et al.,
based leaching: a sustainable process for the recovery of valuable metals from spent
2016). lithium-ion batteries. Waste Manag. https://doi.org/10.1016/j.wasman.2018.01.
021.
Chen, X., Cao, L., Kang, D., Li, J., Zhou, T., Ma, H., 2018b. Recovery of valuable metals
4. Conclusions from mixed types of spent lithium-ion batteries. Part II: selective extraction of li-
thium. Waste Manag. https://doi.org/10.1016/j.wasman.2018.09.013.
The characterization results showed that heat treatment at 700 °C Deboer, M.A., Lammertsma, K., 2013. Scarcity of rare earth elements. ChemSusChem.
https://doi.org/10.1002/cssc.201200794.
for 2 h was effective for degradation of the graphite and polyvinylidene
Diaz, F., Wang, Y., Weyhe, R., Friedrich, B., 2019. Gas generation measurement and
fluoride components of the LIBs, making leaching of the metals easier evaluation during mechanical processing and thermal treatment of spent Li-ion
and more efficient. batteries. Waste Manag. https://doi.org/10.1016/j.wasman.2018.11.029.
Dutta, D., Kumari, A., Panda, R., Jha, S., Gupta, D., Goel, S., Jha, M.K., 2018. Close loop
Since manual separation of the different components would be im-
separation process for the recovery of Co, Cu, Mn, Fe and Li from spent lithium-ion
practicable on a large scale, mechanical processes using whole batteries batteries. Sep. Purif. Technol. https://doi.org/10.1016/j.seppur.2018.02.022.
and segregation of the components by granulometric separation were Ferreira, D.A., Prados, L.M.Z., Majuste, D., Mansur, M.B., 2009. Hydrometallurgical se-
adopted. Particle size separation increased the selectivity towards the paration of aluminum, cobalt, copper, and lithium from spent Li-ion batteries. J.
Power Sources. https://doi.org/10.1016/j.jpowsour.2008.10.077.
active materials, while also reducing the costs associated with different Freitas, M.B.J.G., Celante, V.G., Pietre, M.K., 2010. Electrochemical recovery of cobalt
separation stages. and copper from spent Li-ion batteries as multilayer deposits. J. Power Sources.
Evaluation was made of the effects of the parameters S/L ratio, https://doi.org/10.1016/j.jpowsour.2009.11.131.
Gao, W., Zhang, X., Zheng, X., Lin, X., Cao, H., Zhang, Y., Sun, Z., 2017. Lithium car-
H2O2 volume, temperature, and leaching time. S/L of 1:20 was most bonate recovery from cathode scrap of spent lithium-ion battery: a closed-loop pro-
suitable for recovering Co, Li, and Mn from the LiCoO2 powder using cess. Environ. Sci. Technol. https://doi.org/10.1021/acs.est.6b03320.
DL-malic, citric, and formic acids. The best H2O2 volumes were 1% with Gao, W., Liu, C., Cao, H., Zheng, X., Lin, X., Wang, H., Zhang, Y., Sun, Z., 2018a.
Comprehensive evaluation on effective leaching of critical metals from spent lithium-
citric and formic acids and 6% with DL-malic acid. The best leaching ion batteries. Waste Manag. https://doi.org/10.1016/j.wasman.2018.02.023.
temperatures were 65 °C, using citric and formic acids, and 95 °C, using Gao, W., Song, J., Cao, H., Lin, X., Zhang, X., Zheng, X., Zhang, Y., Sun, Z., 2018b.
DL-malic acid. Under these conditions, leaching for 60 min resulted in Selective recovery of valuable metals from spent lithium-ion batteries – process de-
velopment and kinetics evaluation. J. Clean. Prod. https://doi.org/10.1016/j.jclepro.
Co extraction of 90.57, 20.67, and 43.39%, Li extraction of 93.22,
2018.01.040.
63.25, and 82.18%, and Mn extraction of 99.53, 18.82, and 19.09%, Genieser, R., Ferrari, S., Loveridge, M., Beattie, S.D., Beanland, R., Amari, H., West, G.,
using DL-malic, formic, and citric acids, respectively. Based on these Bhagat, R., 2018. Lithium-ion batteries (NMC/graphite) cycling at 80 °C: different
electrolytes and related degradation mechanism. J. Power Sources. https://doi.org/
results and considering the need to minimize the consumption of energy
10.1016/j.jpowsour.2017.11.014.
and reagents, while achieving satisfactory leaching efficiency, the best Georgi-Maschler, T., Friedrich, B., Weyhe, R., Heegn, H., Rutz, M., 2012. Development of
conditions were determined for the leaching of Co, Li, and Mn from a recycling process for Li-ion batteries. J. Power Sources. https://doi.org/10.1016/j.

12
J. de Oliveira Demarco, et al. Hydrometallurgy 190 (2019) 105169

jpowsour.2012.01.152. org/10.1016/S0378-7753(03)00257-X.
Golmohammadzadeh, R., Rashchi, F., Vahidi, E., 2017. Recovery of lithium and cobalt Saeki, S., Lee, J., Zhang, Q., Saito, F., 2004. Co-grinding LiCoO2 with PVC and water
from spent lithium-ion batteries using organic acids: process optimization and kinetic leaching of metal chlorides formed in ground product. International Journal of
aspects. Waste Manag. https://doi.org/10.1016/j.wasman.2017.03.037. Mineral Processing. https://doi.org/10.1016/j.minpro.2004.08.002.
Golmohammadzadeh, R., Faraji, F., Rashchi, F., 2018. Recovery of lithium and cobalt Shafeeyan, M.S., Daud, W.M.A.W., Houshmand, A., Arami-Niya, A., 2011. Ammonia
from spent lithium-ion batteries (LIBs) using organic acids as leaching reagents: a modification of activated carbon to enhance carbon dioxide adsorption: effect of pre-
review. Resour. Conserv. Recycl. https://doi.org/10.1016/j.resconrec.2018.04.024. oxidation. Appl. Surf. Sci. https://doi.org/10.1016/j.apsusc.2010.11.127.
Guo, Y., Li, F., Zhu, H., Li, G., Huang, J., He, W., 2016. Leaching lithium from the anode Shin, S.M., Kim, N.H., Sohn, J.S., Yang, D.H., Kim, Y.H., 2005. Development of a metal
electrode materials of spent lithium-ion batteries by hydrochloric acid (HCl). Waste recovery process from Li-ion battery wastes. Hydrometallurgy. https://doi.org/10.
Manag. https://doi.org/10.1016/j.wasman.2015.11.036. 1016/j.hydromet.2005.06.004.
Hanisch, C., Loellhoeffel, T., Diekmann, J., Markley, K.J., Haselrieder, W., Kwade, A., Silva, R.G., Afonso, J.C., Mahler, C.F., 2018. Lixiviação ácida de baterias íon-lítio.
2015. Recycling of lithium-ion batteries: a novel method to separate coating and foil Química Nova. https://doi.org/10.21577/0100-4042.20170207.
of electrodes. J. Clean. Prod. https://doi.org/10.1016/j.jclepro.2015.08.026. Silveira, A.V.M., Santana, M.P., Tanabe, E.H., Bertuol, D.A., 2017. Recovery of valuable
Huang, Y., Han, G., Liu, J., Chai, W., Wang, W., Yang, S., Su, S., 2016. A stepwise recovery materials from spent lithium-ion batteries using electrostatic separation. Int. J. Miner.
of metals from hybrid cathodes of spent Li-ion batteries with leaching-flotation- Process. https://doi.org/10.1016/j.minpro.2017.11.003.
precipitation process. J. Power Sources. https://doi.org/10.1016/j.jpowsour.2016. Sun, L., Qiu, K., 2011. Vacuum pyrolysis and hydrometallurgical process for the recovery
06.072. of valuable metals from spent lithium-ion batteries. J. Hazard. Mater. https://doi.
Joulié, M., Laucournet, R., Billy, E., 2014. Hydrometallurgical process for the recovery of org/10.1016/j.jhazmat.2011.07.114.
high value metals from spent lithium nickel cobalt aluminum oxide based lithium-ion Sun, C., Xu, L., Chen, X., Qiu, T., Zhou, T., 2017. Sustainable recovery of valuable metals
batteries. J. Power Sources. https://doi.org/10.1016/j.jpowsour.2013.08.128. from spent lithium-ion batteries using DL-malic acid: leaching and kinetics aspect.
Ku, H., Jung, Y., Jo, M., Park, S., Kim, S., Yang, D., Rhee, K., An, E.M., Sohn, J., Kwon, K., Waste Manag. Res. https://doi.org/10.1177/0734242X17744273.
2016. Recycling of spent lithium-ion battery cathode materials by ammoniacal Swain, B., Jeong, J., Lee, J.C., Lee, G.H., Sohn, J.S., 2007. Hydrometallurgical process for
leaching. J. Hazard. Mater. https://doi.org/10.1016/j.jhazmat.2016.03.062. recovery of cobalt from waste cathodic active material generated during manu-
Kurajica, S., Tkalčec, E., Gržeta, B., Iveković, D., Mandić, V., Popović, J., Kranzelić, D., facturing of lithium-ion batteries. J. Power Sources. https://doi.org/10.1016/j.
2011. Evolution of structural and optical properties in the course of thermal evolu- jpowsour.2007.02.046.
tion of sol-gel derived cobalt-doped gahnite. J. Alloys Compd. https://doi.org/10. Swart, P., Dewulf, J., Biernaux, A., 2014. Resource demand for the production of different
1016/j.jallcom.2010.12.099. cathode materials for lithium-ion batteries. J. Clean. Prod. https://doi.org/10.1016/j.
Lee, C.K., Rhee, K.I., 2002. Preparation of LiCoO2 from spent lithium-ion batteries. J. jclepro.2014.01.056.
Power Sources. https://doi.org/10.1016/S0378-7753(02)00037-X. U.S. Geological Survey, 2011. 2009 Minerals Yearbook: Lithium. USGS, USA.
Lee, C.K., Rhee, K.I., 2003. Reductive leaching of cathodic active materials from lithium- Wang, Q., Ping, P., Zhao, X., Chu, G., Sun, J., Chen, C., 2012. Thermal runaway caused
ion battery wastes. Hydrometallurgy. https://doi.org/10.1016/S0304-386X(02) fire and explosion of lithium-ion battery. J. Power Sources. https://doi.org/10.1016/
00167-6. j.jpowsour.2012.02.038.
Li, L., Ge, J., Chen, R., Wu, F., Chen, S., Zhang, X., 2010a. Environmental friendly Wang, M.M., Zhang, C.C., Zhang, F.S., 2016. An environmental benign process for cobalt
leaching reagent for cobalt and lithium recovery from spent lithium-ion batteries. and lithium recovery from spent lithium-ion batteries by mechanochemical approach.
Waste Manag. https://doi.org/10.1016/j.wasman.2010.08.008. Waste Manag. https://doi.org/10.1016/j.wasman.2016.03.006.
Li, L., Ge, J., Wu, F., Chen, R., Chen, S., Wu, B., 2010b. Recovery of cobalt and lithium Xiao, J., Li, J., Xu, Z., 2017. Recycling metals from lithium-ion battery by mechanical
from spent lithium-ion batteries using organic citric acid as leachant. J. Hazard. separation and vacuum metallurgy. J. Hazard. Mater. https://doi.org/10.1016/j.
Mater. https://doi.org/10.1016/j.jhazmat.2009.11.026. jhazmat.2017.05.024.
Li, L., Lu, J., Ren, Y., Zhang, X.X., Chen, R.J., Wu, F., Amine, K., 2012. Ascorbic-acid- Xin, Y., Guo, X., Chen, S., Wang, J., Wu, F., Xin, B., 2016. Bioleaching of valuable metals
assisted recovery of cobalt and lithium from spent Li-ion batteries. J. Power Sources. Li, Co, Ni and Mn from spent electric vehicle Li-ion batteries for the purpose of re-
https://doi.org/10.1016/j.jpowsour.2012.06.068. covery. J. Clean. Prod. https://doi.org/10.1016/j.jclepro.2016.01.001.
Li, L., Zhai, L., Zhang, X., Lu, J., Chen, R., Wu, F., Amine, K., 2014. Recovery of valuable Yang, H., Zhuang, G.V., Ross, P.N., 2006. Thermal stability of LiPF6 salt and Li-ion bat-
metals from spent lithium-ion batteries by ultrasonic-assisted leaching process. J. tery electrolytes containing LiPF6. J. Power Sources. https://doi.org/10.1109/CISP.
Power Sources. https://doi.org/10.1016/j.jpowsour.2014.04.013. 2008.381.
Li, L., Fan, E., Guan, Y., Zhang, X., Xue, Q., Wei, L., Wu, F., Chen, R., 2017. Sustainable Yang, Y., Huang, G., Xu, S., He, Y., Liu, X., 2016. Thermal treatment process for the
recovery of cathode materials from spent lithium-ion batteries using lactic acid recovery of valuable metals from spent lithium-ion batteries. Hydrometallurgy.
leaching system. ACS Sustain. Chem. Eng. https://doi.org/10.1021/acssuschemeng. https://doi.org/10.1016/j.hydromet.2015.09.025.
7b00571. Yao, L., Feng, Y., Xi, G., 2015. A new method for the synthesis of LiNi 1/3 co 1/3 Mn 1/3 O 2
Meshram, P., Pandey, B.D., Mankhand, T.R., 2015. Hydrometallurgical processing of from waste lithium-ion batteries. RSC Adv. https://doi.org/10.1039/C4RA16390G.
spent lithium-ion batteries (LIBs) in the presence of a reducing agent with emphasis Yu, W., Sun, W., Xu, C., Wang, C., Jia, Y., Qin, X., Xie, C., Yu, S., Xian, M., 2018. Effective
on kinetics of leaching. Chem. Eng. J. https://doi.org/10.1016/j.cej.2015.06.071. adsorption toward p-aminobenzoic acid from aqueous solution by a DL-malic acid
Natarajan, S., Boricha, A.B., Bajaj, H.C., 2018. Recovery of value-added products from modified hyper-crosslinked resin: Equilibria and kinetics. J. Taiwan Inst. Chem. Eng.
cathode and anode material of spent lithium-ion batteries. Waste Manag. https://doi. 89, 105–112. https://doi.org/10.1016/j.jtice.2018.04.022.
org/10.1016/j.wasman.2018.04.032. Zhang, T., He, Y., Wang, F., Ge, L., Zhu, X., Li, H., 2014. Chemical and process miner-
Nayaka, G.P., Pai, K.V., Santhosh, G., Manjanna, J., 2016. Dissolution of cathode active alogical characterizations of spent lithium-ion batteries: An approach by multi-ana-
material of spent Li-ion batteries using tartaric acid and ascorbic acid mixture to lytical techniques. Waste Manag. https://doi.org/10.1016/j.wasman.2014.01.002.
recover Co. Hydrometallurgy. https://doi.org/10.1016/j.hydromet.2016.01.026. Zhang, X., Cao, H., Xie, Y., Ning, P., An, H., You, H., Nawaz, F., 2015. A closed-loop
Ordoñez, J., Gago, E.J., Girard, A., 2016. Processes and technologies for the recycling and process for recycling LiNi1/3Co1/3Mn1/3O2 from the cathode scraps of lithium-ion
recovery of spent lithium-ion batteries. Renew. Sust. Energ. Rev. https://doi.org/10. batteries: process optimization and kinetics analysis. Sep. Pur. Technol. https://doi.
1016/j.rser.2015.12.363. org/10.1016/j.seppur.2015.07.003.
Pant, D., Dolker, T., 2017. Green and facile method for the recovery of spent lithium Zhang, X., Gao, B., Creamer, A.E., Cao, C., Li, Y., 2017. Adsorption of VOCs onto en-
nickel manganese cobalt oxide (NMC) based lithium-ion batteries. Waste Manag. gineered carbon materials: a review. J. Hazard. Mater. https://doi.org/10.1016/j.
https://doi.org/10.1016/j.wasman.2016.09.039. jhazmat.2017.05.013.
Paulino, J.F., Busnardo, N.G., Afonso, J.C., 2008. Recovery of valuable elements from Zhang, J., Hu, J., Zhang, W., Chen, Y., Wang, C., 2018. Efficient and economical recovery
spent Li-batteries. J. Hazard. Mater. https://doi.org/10.1016/j.jhazmat.2007.10.048. of lithium, cobalt, nickel, manganese from cathode scrap of spent lithium-ion bat-
Pinna, E.G., Ruiz, M.C., Ojeda, M.W., Rodriguez, M.H., 2017. Cathodes of spent Li-ion teries. J. Clean. Prod. 204, 437–446. https://doi.org/10.1016/j.jclepro.2018.09.033.
batteries: Dissolution with phosphoric acid and recovery of lithium and cobalt from Zheng, Y., Song, W., Mo, W.T., Zhou, L., Liu, J.W., 2018. Lithium fluoride recovery from
leach liquors. Hydrometallurgy. https://doi.org/10.1016/j.hydromet.2016.10.024. cathode material of spent lithium-ion battery. RSC Adv. 8, 8990–8998. https://doi.
Ravdel, B., Abraham, K.M., Gitzendanner, R., DiCarlo, J., Lucht, B., Campion, C., 2003. org/10.1039/c8ra00061a.
Thermal stability of lithium-ion battery electrolytes. J. Power Sources. https://doi.

13

You might also like