You are on page 1of 5

Hydrometallurgy 151 (2015) 73–77

Contents lists available at ScienceDirect

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

Recovery of valuable metal ions from the spent lithium-ion battery using
aqueous mixture of mild organic acids as alternative to mineral acids
G.P. Nayaka a, J. Manjanna b,⁎, K.V. Pai a, R. Vadavi c, S.J. Keny d, V.S. Tripathi d
a
Dept. of Industrial Chemistry, Kuvempu University, Shankaraghatta 577 451, India
b
Dept. of Chemistry, Rani Channamma University, Belagavi 591 156, India
c
PG Dept. of Chemistry, B.K. College, Club Road, Belagavi 590 001, India
d
Chemistry Group, Bhabha Atomic Research Centre, Mumbai 400 085, India

a r t i c l e i n f o a b s t r a c t

Article history: A well characterized cathode material (LiCoO2) recovered from spent lithium-ion battery is dissolved in aqueous
Received 14 May 2014 mixture of citric acid (chelating agent) and ascorbic acid (reductant) at 80 °C. The dissolution proceeds with a
Received in revised form 7 November 2014 reductive-complexing mechanism, and complete dissolution occurs in about 6 h when stoichiometric amount
Accepted 13 November 2014
of C/A is used. The dissolution rate constants (k) are 3.1 × 10−3 min−1 for Li and 0.8 × 10−3 min−1 for Co ions
Available online 18 November 2014
as determined by ‘cubic rate law’ plots. The formation of Co(III)– to Co(II)–citrate during the dissolution is con-
Keywords:
firmed from the UV–Vis spectra. The dissolved solution was subjected for selective precipitation of cobalt as Co-
Lithium-ion battery oxalate and lithium as LiF using oxalic acid and NH4F, respectively. The present study has a merit when compared
Recovery of cobalt and lithium to literature reports as we make use of mild organic acids as alternatives to mineral acids.
Chemical extraction © 2014 Elsevier B.V. All rights reserved.
Citric acid
Oxalic acid
Ammonium fluoride

1. Introduction electrolytes stable at high potential with greater ionic conductivity


than polymer electrolytes.
Recovery of commercially important metal ions like Co and Li from LIBs are put to use in all personal computers, cellular phones, cam-
spent Li-ion batteries is advantageous considering the limited availabil- eras and many other modern-life appliances (Ra and Han, 2006; Nan
ity of these metal ions and environmental regulations. LiCoO2 as cath- et al., 2005; Lizuka et al., 2013; Wee, 2007; Swain et al., 2007) and re-
ode materials was discovered by the research group of Goodenough in cently in electric vehicles. Thus, LIB consumption is considerably high
1980 at Oxford University, which established a rechargeable cell within around the world. For instance, the household battery industry in the
the 4 V range using LiCoO2 as cathode and lithium metal as anode ma- USA is estimated to be a US $ 2.5 billion industry with annual sales of
terial (Mizushima et al., 1980). LIBs have been introduced to the market nearly 3 billion batteries. These batteries, also known as dry cells, are
by Sony Corp. in 1991 (Zou et al., 2013; Georgi-Maschler et al., 2012; used in over 900 million battery-operated devices. In Europe, 5 billion
Nishi, 2001). These batteries have substituted Ni–Cd and Ni–MH batte- units of batteries were produced in year 2000 (Bernardes et al., 2004).
ries in many applications due to their high energy density, low auto- In the USA and Europe, the consumption of batteries is estimated to
discharge rate, and excellent cycle life. Presently, lithium-intercalating be 8 billion units per year. In Japan, around 6 billion batteries were pro-
material, typically graphitic carbon is used as anode (negative elec- duced in 2004, while almost 1 billion units are consumed every year in
trode) and LiCoO2 is the most used cathode (positive electrode) materi- Brazil (Georgi-Maschler et al., 2012; Xu et al., 2008; Salgado et al., 2003;
al in LIB due to their good performance (Armstrong and Bruce, 1996; Gupta and Manthiram, 1996). World LIB production reached 500 mil-
Freitas and Garcia, 2007; Xin et al., 2009; Nan et al., 2005), although sev- lion units in 2000 and almost 4.6 billion in 2010. Therefore, equally
eral other oxides such as Ni, Mn-doped LiCoO2, LiMn2O4 and LiFePO4 large amounts of spent LIBs have to be handled after their lifetime.
have been investigated (Castillo et al., 2002). Also, significant improve- The increasing public concern on environmental sustainability in
ments in electrolyte (liquid and polymer) system are also underway. the last decade has resulted in stricter regulations worldwide, for the
For instance, Nanini-Maury et al. (2014) have reported new liquid adequate destination of hazardous residues from electronic wastage,
including LIBs. These regulations have impelled the society to look for
⁎ Corresponding author. technical solutions such as e-waste management and resource recycling
E-mail address: jmanjanna@rediffmail.com (J. Manjanna). techniques. Consequently, the spent LIB material has become an

http://dx.doi.org/10.1016/j.hydromet.2014.11.006
0304-386X/© 2014 Elsevier B.V. All rights reserved.
74 G.P. Nayaka et al. / Hydrometallurgy 151 (2015) 73–77

environmental burden. In China, consumer battery waste amounted from a typical spent LIB and the selective precipitation of constituent
to 200–500 t/year from 2002 to 2006 with significant amounts of metals are reported here. To the best of our knowledge, the present
metals, organic chemicals and plastics in the following proportions: study is the first report for the dissolution of LiCoO2 using ascorbic acid
5–20% Co, 5–10% Ni, 5–7% Li, ~ 15% organic chemicals and ~ 7% plas- as reductant.
tics. This composition varies slightly with different manufacturers
(Li et al., 2010a, 2010b, 2013; A.K. Jha et al., 2013; Sun and Qiu,
2. Experimental
2011). The use of cobalt in LIBs has grown from 700 to 1200 tpa
during the year 1995–2005, which is ~ 25% of cobalt demand global-
Several spent Li-ion batteries of BL-5CA (Nokia) series available in
ly. The other important metal, lithium, has a current demand of
the local market were collected. For safe handling, they were discharged
110,000 tpa and is expected to rise three fold within ten years
completely and then dismantled to separate the cathode and anode ma-
(M.K. Jha et al., 2013; Al-Thyabat et al., 2013; Dewulf et al., 2010;
terials coated on Al- and Cu-foil, respectively. The cathode material was
Georgi-Maschler et al., 2012). A typical LIB contains about 27.5%
scrubbed carefully and heated at 700 °C for 2 h to burn off the organics
LiCoO 2 , 24.5% steel/Ni, 14.5% Cu/Al, 16% carbon, 3.5% electrolyte
such as polyvinylidene fluoride (Li et al., 2010b). More than 90% of the
and 14% polymer (Li et al., 2013). Thus, toxic but commercially im-
this cathode material was found to be LiCoO2 based on X-ray diffraction,
portant metals and flammable compounds (LiBF4 and LiPF6 dissolved
scanning electron microscopy (SEM) and energy dispersive X-ray (EDX)
in organic solvent) present in these spent LIBs must be treated for re-
analysis. The remaining 5–10% is expected to be carbon residue formed
source recycling due to limited availability of these natural resources
due to organic burn off.
and also to meet the environmental regulations. It is impervious to
The above cathode material was subjected for chemical dissolution
note that there is hardly any natural resource of lithium in India.
using mild chemical formulation, an aqueous mixture of citric acid and
The existing methods for recycling spent LIBs are mainly pyro-
ascorbic acid (C–A). In a typical stoichiometric case, 20 mM metal ion
metallurgical and hydrometallurgical processes (Xu et al., 2008;
worth of LiCoO2 sample (0.2 g) was added to 100 ml water containing
Georgi-Maschler et al., 2012). In the pyrometallurgical processes, or-
100 mM of citric acid and 20 mM ascorbic acid (C–A) at 80 °C. The mix-
ganic electrolytes and binder are burnt off and make leaching of
ture was kept stirred for about 6 h. The pH of the formulation before and
valuable metals more easy (Li et al., 2010a; Freitas et al., 2010;
after dissolution was 2.05 and 2.66, respectively. The periodically col-
Wang et al., 2009; Chen et al., 2011). In the hydrometallurgical pro-
lected samples were syringe filtered (0.2 μm) and estimated for Co
cesses, the dismantled electrodes are first dissolved in concentrated
and Li ions using atomic absorption spectrometer (AAS). Also, the UV–
acids followed by recovery of metal ions by precipitation (Castillo et al.,
Vis absorption spectra of the formulation during dissolution (as it
2002; Shin et al., 2005; Ferreira et al., 2009; Li et al., 2010b; Espinosa
turned pink color) were recorded to known the complexation behavior
et al., 2004; Contestabile et al., 1999; Chagnes and Pospiech, 2013).
of Co ions. At the end of dissolution, the undissolved black carbon resi-
Therefore, recycling of constituent metal ions in LIBs has become inev-
due was separated by filtration.
itable towards the development of green technologies and to achieve
On adding stoichiometric amount of oxalic acid to the above dis-
some economical benefits.
solved solution, selective precipitation of Co as Co-oxalate occurred. It
There are many studies on the dissolution of active cathode materials
was separated by filtration, and NH4F (100 mM) was added to the fil-
from spent LIBs using strong acids viz., 2 M H2SO4 (A.K. Jha et al., 2013;
trate to precipitate Li as LiF. The recovery of Co and Li from the dissolved
Chen et al., 2011; Contestabile et al., 2001), 4 M HCl (Contestabile et al.,
solution was estimated by analyzing the filtrate solution for Co and Li by
2001) and 1 M HNO3 (Lee and Rhee, 2002) as leaching agents. The
AAS.
leaching efficiency of Co is highest in HCl when compared to H2SO4
and HNO3 at 80 °C. However, on adding H2O2 (5–20 vol.%) to H2SO4
and HNO3, the leaching efficiency was increased (Chen et al., 2011; 3. Results and discussion
Mantuano et al., 2006; Lee and Rhee, 2003) because H2O2 acts as a
reducing agent in acidic medium. Recently, a novel method to recycle Fig. 1 shows the XRD pattern of the cathodic active material obtained
mixed cathode materials for LIBs has been reported by Zou et al. from the spent LIBs. All the peaks could be indexed to rock-salt struc-
(2013). The authors have dissolved mixtures of LiCoO2, LiMn2O4, tured LiCoO2 (JCPDS 44-0145). The SEM/EDX analysis (not shown
Li(Ni0.33Mn0.33Co0.33)O2 and LiFePO4 in 4 M H2SO4 + 30 wt.% H2O2 at here) confirmed the residual carbon, 5–10 wt.%, present along with
70–80 °C for 2–3 h. The dissolved ions were precipitated as metal this cathode material due to organic burn off. The origin for such carbon
hydroxides at different pH and the Li as Li 2 CO 3 at 40 °C. Finally, might be from the acetylene black (used to ensure the electronic con-
they have succeeded in preparing the same active cathode material, ductivity in the cathodes) and polyvinylidene fluoride (used as a binder
Li(Ni0.33Mn0.33Co0.33)O2, by decomposing these metal hydroxides. to ensure the electrode cohesion).
There are few studies on the use of mild organic acids viz., 1.25 M citric
acid (Li et al., 2010b), 1.0 M oxalic acid (Sun and Qiu, 2012), 1.5 M
aspartic acid (Li et al., 2013) and 1.5 M malic acid (Li et al., 2010a) to
leach metal ions from spent LIBs. All of these reagents have not shown
complete dissolution of Co and Li ions, however, the dissolution was en-
hanced by adding 1–6 vol.% of H2O2 (reducing agent).
To develop environmentally benign process i.e. to avoid using the
mineral acids, we have formulated an aqueous mixture of mild organic
acids containing reductant (ascorbic acid) and chelating cum buffering
agent (citric acid). We believe such a mild formulation should be of
great help in the recovery of metal ions from spent LIBs and/or electron-
ic gadgets, especially during large scale treatment. In fact, it is shown
here that the leaching or chemical extraction efficiency is not compro-
mised with this mild formulation. This is because of the availability of
stoichiometric amounts of H+ ions from these reagents during dissolu-
tion. In the next step, we have succeeded in selective precipitation of co-
balt as Co-oxalate and lithium as LiF towards the recycling of active Fig. 1. Powder XRD patterns of the cathodic material (calcinated at 700 0C for 2 h) from a
cathode material. The details on dissolution behavior of cathode material spent LIB.
G.P. Nayaka et al. / Hydrometallurgy 151 (2015) 73–77 75

Fig. 2 shows the % Li and Co ions released from LiCoO 2 in C–A be due to reductive-complexing mechanism. The Co3+ (d6; r = 0.63 Ǻ)
(100 mM–20 mM) mixture at 80 °C over a period of 6 h. We see about in the oxide lattice can be easily ejected by reduction to Co2 + (d7;
80% leaching of Co ions whereas almost complete leaching of Li ions. r = 0.74 Ǻ), which is stabilized by complexation with chelating
Such a discrepancy could be ascribed to the nature of spent cathode ma- agent. In our study, we used ascorbic acid as reducing agent
terial used here. It is well known that a fraction of Li ions are irreversibly along with complexing/buffering agent, citric acid. Ascorbic acid
interacted with the anode (graphite) material during the life-time (H2A → 2H+ + 2e + A) is an environmentally benign weak reducing
charging/discharging process of LIB (Besenhard, 1999). Thus, it is plau- agent (E = 0.2 V vs. SHE), however, kinetically it has been proven to
sible that the oxide sample used here is Li-depleted to about 20%. In our be a strong reductant in the metal oxide dissolution studies (Manjanna
calculation, we have assumed nominal composition of LiCoO2 instead of et al., 2001). It gives two electrons to form a neutral dehydroascorbic
Li1 − xCoO2 and hence we have obtained 100% of Li and about 80% of Co. acid (A). Thus the dissolution follows a reductive-complexing mecha-
However, the presence of 5–10 wt.% carbon residue along with the sam- nism: Co3+}oxide + e (H2A) → Co2+}oxide → Co2+}(aq.) → Co(II)–citrate.
ple must be taken into account for knowing the actual amount of metal Such a heterogeneous reaction is expected to occur at the interface of
ions released here. A small fraction (b5 wt.%) of undissolved black res- solid (oxide lattice) and liquid. Once the lattice is disturbed, the sur-
idue was not found to contain any Co or Li ions. Based on these consid- rounding Co3+ as well as Li+ ions will also be leached to solution
erations, it is clear that the mild organic acid mixture of CA used here is through complexation. Both Li and Co ions are expected as citrate com-
able to dissolve the cathode material almost completely. Thus, it is pos- plexes because citrate is a strong complexing agent and available in
sible to overcome from the use of mineral acids (M.K. Jha et al., 2013; stoichiometric amount (assuming 1:2 metal-to-ligand ratio), unlike
Chen et al., 2011; Contestabile et al., 2001; Lee and Rhee, 2002, 2003) ascorbate. As the dissolution proceeds, the initially colorless solution
for recovery of valuable metal ions from spent LIB. turned pink color and the intensity increased with dissolution time.
We have determined the dissolution rate constant (k) by using a This is an indication of Co ion (3d metal) complexation preferably with
cubic rate law (Segal and Sellers, 1982), k t = (1 − f)1/3 where f — is citrate ligand. To our knowledge, there are no reports available on the
the fraction dissolved at different intervals of time, t. The inset of spectral features of Co(III)– and Co(II)–citrate complexes. Thus, we
Fig. 2 shows the kinetic plots and the corresponding rate constants. have recorded the UV–Vis spectra of the dissolution mixture.
For slow second-stage dissolution, k = 3.1 × 10− 3 min− 1 for Li and As shown in Fig. 3, it is clear that the absorbance of the solution in-
k = 0.8 × 10−3 min−1 for Co. Although the cubic rate law has been de- creased with dissolution time. The maximum absorption (λmax) around
rived for spherical particles, it is well demonstrated to be applicable 350 nm and 512 nm observed here is attributed to the Co(II)–citrate and
even for polydispersed particles (Manjanna and Venkateswaran, Co(III)–citrate, respectively. Initially, the absorption peak for Co(III)–
2002). We have reported the dissolution behavior only for stoichiomet- citrate alone is seen showing that Co(III) is dominant in the spent
ric (optimum) C/A ratio with respect to the cathode material used here cathode material (LiCoO2). After about an hour, the concentration
(LiCoO2). However, it was observed that rate of dissolution increases of Co(II)–citrate starts building-up due to the reduction of Co(III)–
with increase in C/A ratio and decreases with decrease in C/A ratio. to Co(II)–citrate by ascorbic acid. Thus, the intensity of Co(II)–citrate
We have not paid much attention on these kinetic aspects as the main increases rapidly when compared to Co(III)–citrate over the period
focus of this study is to recover the Co and Li though selective precipita- of dissolution. Nevertheless, the hump at 512 nm did not disappear
tion. In the real system application, it is essential to maintain the opti- because there was an increase in the entire UV–Vis region, due to
mum C/A ratio considering the post processing modalities in this the formation of Co(II)–citrate. In other words, the pull exerted by
valuable metal ion recovery. the increased absorbance at 350 nm undermines the decrease in ab-
On adding H2O2 to citric acid, the leaching efficiency of Co and Li in- sorbance at 512 nm. A similar behavior is seen during the oxidation
creased from 25% to 91% and 54% to 99%, respectively (Li et al., 2010b). of V(II) to V(III)–formate/NTA (Tripathi et al., 2004). Furthermore,
In similar studies, on adding H2O2 in malic acid (Li et al., 2010a) the one cannot rule out the fraction of Co(III)–citrate remaining because
leaching of Co increased from 37% to 93% and Li from 54% to 99%; some amount of ascorbic acid is also utilized for deoxygenation of
H2O2 in aspartic acid (Li et al., 2013) increased the leaching of Co and water in the dissolving medium.
Li from only 1% to 60%. In all these studies, 1–6 vol.% of H2O2 is used. The main reason for choosing the citric acid as complexing agent
Thus, it is beneficial to have a reducing agent along with organic acids. here is due to its weak stability constant with metal ions, unlike strong
The reason for enhanced dissolution in all these cases is presumed to metal complexes such as Co(III)–EDTA and (log K = 41.4) and Fe(III)–
EDTA (log K = 25.1) (Martell and Smith, 1989). For instance, the log K
of Co(II)–citrate is reported to be 6.187 (Armstrong and Bruce, 1996;
Lee and Reeder, 2006). Such a weak complexation is advantageous
here for the subsequent precipitate of Co(II) ions either as hydroxide

100 Li 350 nm
a (5 min), b (10 min), c (20 min),
0.4 j d (30 min), e (60 min) f (90 min),
80 Co g (120 min), h (180 min) i (240 min)
i and j (280 min)
Dissolution, %

0.3
Absorbance

1.0
60
5.2 x 10-3/ min h
0.8
0.83 x 10-3/ min 0.2
40 Co(II)-citrate Co(III)-citrate
(1-f)1/3

0.6 g
512 nm
3.1 x 10-3/ min
0.1
20
0.4 f
e
0.2
0 50 100 150 200 250 300 350 400
Dissolution time, min 0.0
0 350 400 450 500 550 600
0 50 100 150 200 250 300 350 400 Wavelength (nm)
Time, min
Fig. 3. UV–Visible spectra of the dissolved solution at different intervals of time showing
Fig. 2. Dissolution profiles of Co and Li. Inset shows cubic rate-law, (1 − f)1/3 vs. time, plots. the Co(II)–citrate complex.
76 G.P. Nayaka et al. / Hydrometallurgy 151 (2015) 73–77

from the spent Li-ion battery (BL-5CA Nokia series) was subjected for
dissolution in a mixture of citric acid (chelating agent) and ascorbic
acid (reductant) at 80 °C. Almost complete dissolution was achieved
via a reductive-complexation mechanism. Based on the cubic rate law,
the dissolution rate constants for the slow second stage were found to
be 3.1 × 10−3 min−1 for Li and 0.8 × 10−3 min−1 for Co ions. UV–Vis
spectra of pink colored dissolved solution confirm the build-up of
Co(II)–citrate (λmax ≈ 350 nm) through the reduction of Co(III)–citrate
(λmax ≈ 512 nm). The dissolved solution was subjected for selective
precipitation of cobalt as Co-oxalate and Li as LiF. There was a complete
removal of both Co and Li from the dissolved solution. Nevertheless, our
attempts are on to recover Li as Li2CO3 which is a suitable precursor for
obtaining the starting material.

Acknowledgment

One of the authors (G.P. Nayaka) gratefully acknowledges Kuvempu


University for financial support (Grant No.: 52/914).
Fig. 4. Powder X-ray diffraction pattern of CoC2O4 2H2O precipitate. Inset is the FTIR spec-
tra and photo of cobalt oxalate powder.
Appendix A. Supplementary data
or oxalate. And also can be passed through cation-exchange resin to pick
metal ions through dissociation (Ananthan et al., 2003). Supplementary data to this article can be found online at http://dx.
At the end of dissolution, on adding oxalic acid (0.63 g which is doi.org/10.1016/j.hydromet.2014.11.006.
worth of 100 mM) to the dissolved solution, the Co(II)–citrate ions
were precipitated as Co(II)–oxalate immediately at room temperature.
References
Such a solid product could be easily separated by filtration. It was
washed with excess water and dried in oven at ~90 °C for 24 h. Fig. 4 Al-Thyabat, S., Nakamura, T., Shibata, E., Iizuka, A., 2013. Adaptation of minerals process-
shows the XRD pattern and FTIR spectra (given as inset) of this solid ing operations for lithium-ion (LiBs) and nickel metal hydride (NiMH) batteries
recycling: critical review. Miner. Eng. 45, 4–17.
powder sample. The photo of this sample is also given in the inset of Ananthan, P., Venkateswaran, G., Manjanna, J., 2003. Enhanced dissolution of hematite in
Fig. 4. The XRD pattern is in good agreement with the reported pattern reductive-complexing formulation under regenerative mode. Chem. Eng. Sci. 58,
for the orthorhombic β-phase of pure crystalline CoC2O4 2H2O with 5103–5109.
Armstrong, A.R., Bruce, P.G., 1996. Synthesis of layered LiMnO2 as an electrode for re-
space group of Cccm (JCPDS file: 25-0250) (Chen et al., 2011). The chargeable lithium batteries. Nature 381, 499–500.
FTIR band ~3373 cm−1 has been assigned to O–H stretching vibration Bernardes, A.M., Espinosa, D.C.R., Tenorio, J.A.S., 2004. Recycling of batteries: a review of
and the band ~ 1623 cm−1 has been assigned to CO stretching. The current processes and technologies. J. Power Sources 130, 291–298.
Besenhard, J.O (Ed.), 1999. Handbook of Battery Materials. Wiley-VCH, New York. ISBN:
two peaks around 1359 cm−1 and 1316 cm−1 are due to the presence
3-527-29469-4, pp. 383–409.
of carbonyl group. These values are similar to the report by Sun and Castillo, S., Ansart, F., Laberty-Robert, C., Portal, J., 2002. Advances in the recovering of
Qiu (2012). The TG/DTA of this product (not shown here) further con- spent lithium battery compounds. J. Power Sources 112, 247–254.
Chagnes, A., Pospiech, B., 2013. A brief review on hydrometallurgical technologies for
firmed the high purity of Co-oxalate with a nominal composition of
recycling spent lithium-ion batteries. J. Chem. Technol. Biotechnol. 88, 1191–1199.
CoC2O4 2H2O. This can be readily used as a Co precursor for preparing Chen, L., Tang, X., Zhang, Y., Li, L., Zeng, Z., Zhang, Y., 2011. Process for the recovery of co-
the parent cathode material. balt oxalate from spent lithium-ion batteries. Hydrometallurgy 108, 80–86.
After separating the Co-oxalate by centrifugation and membrane fil- Contestabile, M., Panero, S., Scrosati, B., 1999. A laboratory-scale lithium battery recycling
process. J. Power Sources 83, 75–78.
tration (0.2 μm) unit, there was no measurable Co present in the filtrate. Contestabile, M., Panero, S., Scrosati, B., 2001. A laboratory-scale lithium-ion battery
Thus, a complete removal of Co is achieved here. In order to precipitate recycling process. J. Power Sources 92, 65–69.
the Li-ions, about 100 mM worth of NH4F was added to the above fil- Dewulf, J., Vorst, G.V.D., Denturck, K., Langenhove, H.V., 2010. Recycling rechargeable lith-
ium ion batteries: critical analysis of natural resource saving. Resour. Conserv. Recycl.
trate, which led to complete precipitation of LiF as confirmed from the 54, 229–234.
flame photometry. Although, Co-oxalate and LiF could be used as pre- Espinosa, D.C.R., Bernardes, A.M., Tenório, J.A.S., 2004. An overview on the current pro-
cursors to synthesize LiCoO2 for reuse in LIB, we are not in favor of cesses for the recycling of batteries. J. Power Sources 135, 311–319.
Ferreira, D.A., Prados, L.M.Z., Majuste, D., Mansur, M.B., 2009. Hydrometallurgical separa-
using LiF for the well-known reason of fluoride contamination. In fact, tion of aluminium, cobalt, copper and lithium from spent Li-ion batteries. J. Power
we tried to precipitate Li as Li2CO3, however, there was no precipitation Sources 187, 238–246.
when dry ice (CO2) or Na2CO3 was added to the dissolved solution ei- Freitas, M.B.J.G., Garcia, E.M., 2007. Electrochemical recycling of cobalt from cathodes of
spent lithium-ion batteries. J. Power Sources 171, 953–959.
ther before or after Co removal. Contestabile et al. (1999), Nan et al.
Freitas, M.B.J.G., Celante, V.G., Pietre, M.K., 2010. Electrochemical recovery of cobalt and
(2005) and Freitas and Garcia (2007) have reported the precipitation copper from spent Li-ion batteries as multilayer deposits. J. Power Sources 195,
of Li as Li2CO3 by passing CO2 gas or Na2CO3 in the dissolved medium 3309–3315.
Georgi-Maschler, T., Friedrich, B., Weyhe, R., Heegn, H., Rutz, M., 2012. Development of a
wherein only mineral acids (HCl, HNO3 or H2SO4) are used and not
recycling process for Li-ion batteries. J. Power Sources 207, 173–182.
the chelating agents. Thus, it is essential that free Li ions are available Gupta, R., Manthiram, A., 1996. Chemical extraction of lithium from layered LiCoO2. J.
for precipitation. However, in our study, such free Li ions are not avail- Solid State Chem. 121, 483–491.
able as we have used complexing agent (citric acid) in the dissolution Jha, A.K., Jha, M.K., Kumari, A., Sahu, S.K., Kumar, V., Pandey, B.D., 2013a. Selective separa-
tion and recovery of cobalt from leach liquor of discarded Li-ion batteries using
formulation. Therefore our efforts are on to oxidize/decompose the cit- thiophinic extractant. Sep. Purif. Technol. 104, 160–166.
rate ions and then use dry ice to form Li2CO3. Jha, M.K., Kumari, A., Jha, A.K., Kumar, V., Hait, J., Pandey, B.D., 2013b. Recovery of lithium
and cobalt from waste lithium ion batteries of mobile phone. Waste Manag. 33,
1890–1897.
4. Conclusions Lee, Y.J., Reeder, R.J., 2006. The role of citrate and phthalate during Co(II) coprecipitation
with calcite. Geochim. Cosmochim. Acta 70, 2253–2263.
Aqueous mixtures of mild organic acids are used here as an alterna- Lee, C.K., Rhee, K.I., 2002. Preparation of LiCoO2 from spent lithium-ion batteries. J. Power
Sources 109, 17–21.
tive to mineral acids for the recovery of valuable metal ions from spent Lee, C.K., Rhee, K.I., 2003. Reductive leaching of cathodic active materials from lithium ion
Li-ion batteries. As a typical case, the cathode material, LiCoO2, collected battery wastes. Hydrometallurgy 68, 5–10.
G.P. Nayaka et al. / Hydrometallurgy 151 (2015) 73–77 77

Li, L., Ge, J., Chen, R., Wu, F., Chen, S., Zhang, X., 2010a. Environmental friendly leaching Salgado, A.L., Veloso, A.M.O., Pereira, D.D., Gontijo, G.S., Salum, A., Mansur, M.B., 2003. Re-
reagent for cobalt and lithium recovery from spent lithium-ion batteries. Waste covery of zinc and manganese from spent alkaline batteries by liquid–liquid extrac-
Manag. 30, 2615–2621. tion with Cyanex 272. J. Power Sources 115, 367–373.
Li, L., Ge, J., Wu, F., Chen, R., Chen, S., Wu, B., 2010b. Recovery of cobalt and lithium from Segal, M.G., Sellers, R.M., 1982. Kinetics of metal oxide dissolution: reductive dissolution
spent lithium ion batteries using organic citric acid as leachant. J. Hazard. Mater. 176, of nickel ferrite by tris(picolinato) vanadium (II). J. Chem. Soc. Faraday Trans. 78,
288–293. 1149–1164.
Li, L., Dunn, J.B., Zhang, X.X., Gaines, L., Chen, R.J., Wu, F., Amine, K., 2013. Recovery of Shin, S.M., Kim, N.H., Sohn, J.S., Yang, D.H., Kim, Y.H., 2005. Development of a metal recov-
metals from spent lithium-ion batteries with organic acids as leaching reagents and ery process from Li-ion battery wastes. Hydrometallurgy 79, 172–181.
environmental assessment. J. Power Sources 233, 180–189. Sun, L., Qiu, K., 2011. Vacuum pyrolysis and hydrometallurgical process for the re-
Lizuka, A., Yamashita, Y., Nagasawa, H., Yamasaki, A., Yanagisawa, Y., 2013. Separation of covery of valuable metals from spent lithium-ion batteries. J. Hazard. Mater.
lithium and cobalt from waste lithium-ion batteries via bipolar membrane electrodi- 194, 378–384.
alysis coupled with chelation. Sep. Purif. Technol. 113, 33–41. Sun, L., Qiu, K., 2012. Organic oxalate as leachant and precipitant for the recovery of valu-
Manjanna, J., Venkateswaran, G., 2002. Preparation and kinetic considerations for the dis- able metals from spent lithium-ion batteries. Waste Manag. 32, 1575–1582.
solution of Cr-substituted iron oxides in reductive-complexing formulations. Can. J. Swain, B., Jeong, J., Lee, J.C., Lee, G.H., Sohn, J.S., 2007. Hydrometallurgical process for re-
Chem. Eng. 80, 882–896. covery of cobalt from waste cathodic active material generated during manufacturing
Manjanna, J., Venkateswaran, G., Sherigara, B.S., Nayak, P.V., 2001. Dissolution studies of of lithium ion batteries. J. Power Sources 167, 536–544.
chromium substituted iron oxides in reductive-complexing agent mixtures. Hydro- Tripathi, V.S., Manjanna, J., Venkateswaran, G., Gokhale, B.K., Balaji, V., 2004. Electrolytic
metallurgy 60, 155–165. preparation of vanadium(II) formate in pilot-plant scale using stainless steel mesh
Mantuano, D.P., Dorella, G., Elias, R.C.A., Mansur, M.B., 2006. Analysis of a hydrometallur- electrodes: dissolution of α-Fe2O3/Fe1.6Cr0.4O3 in an aqueous VII–NTA complex. Ind.
gical route to recover base metals from spent rechargeable batteries by liquid–liquid Eng. Chem. Res. 43, 5989–5995.
extraction with Cyanex 272. J. Power Sources 159, 1510–1518. Wang, R.C., Lin, Y.C., Wu, S.H., 2009. A novel recovery process of metal values from the
Martell, A.E., Smith, R.M., 1989. Critical Stability Constants. Plenum Press, New York, p. 98. cathode active materials of the lithium-ion secondary batteries. Hydrometallurgy
Mizushima, K., Jones, P.C., Wiseman, P.J., Goodenough, J.B., 1980. LixCoO2 (0 b x b −1): a 99, 194–201.
new cathode materials for batteries of high energy density. Mater. Res. Bull. 15, Wee, J.-H., 2007. A feasibility study on direct methanol fuel cells for laptop com-
783–789. puters based on a cost comparison with lithium-ion batteries. J. Power Sources
Nan, J., Han, D., Zuo, X., 2005. Recovery of metal values from spent lithium-ion batteries 173, 424–436.
with chemical deposition and solvent extraction. J. Power Sources 152, 278–284. Xin, B., Zhang, D., Zhang, X., Xia, Y., Wu, F., Chen, S., Li, L., 2009. Bioleaching mechanism of
Nanini-Maury, E., Światowska, J., Chagnes, A., Zanna, S., Tran-Van, P., Marcus, P., Cassir, M., Co and Li from spent lithium-ion battery by the mixed culture of acidophilic sulfur-
2014. Electrochemical behavior of sebaconitrile as a cosolvent in the formulation of elec- oxidizing and iron-oxidizing bacteria. Bioresour. Technol. 100, 6163–6169.
trolytes at high potentials for lithium-ion batteries. Electrochim. Acta 115, 223–233. Xu, J., Thomas, H.R., Francis, R.W., Lum, K.R., Wang, J., Liang, B., 2008. A review of processes
Nishi, Y., 2001. Lithium ion secondary batteries; past 10 years and the future. J. Power and technologies for the recycling of lithium-ion secondary batteries. J. Power
Sources 100, 101–106. Sources 177, 512–527.
Ra, D.I., Han, K.S., 2006. Used lithium ion rechargeable battery recycling using Etoile- Zou, H., Gratz, E., Apelian, D., Wang, Y., 2013. A novel method to recycle mixed cathode
Rebatt technology. J. Power Sources 163, 284–288. materials for lithium ion batteries. Green Chem. 15, 1183–1191.

You might also like