You are on page 1of 5

DOI: 10.1002/ente.

201700488

Electrochemical Lithium Recovery with a LiMn2O4–Zinc


Battery System using Zinc as a Negative Electrode
Seongsoo Kim,[a] Jaehan Lee,[a, b] Seoni Kim,[a] Seonghwan Kim,[a, b] and
Jeyong Yoon*[a, b]

Securing lithium resources is an emerging issue owing to the low price, large capacity, and stable redox potential, was pro-
widespread use of lithium-ion batteries. Recently, an electro- posed as an alternative negative electrode material. Using a
chemical lithium recovery method based on the principle of LiMn2O4–zinc (LMO-Zn) battery system, lithium was selec-
a lithium-ion battery using a lithium-capturing electrode and tively recovered with an energy consumption of 6.3 Wh mol@1
silver as a negative electrode was proposed to meet the in- of lithium recovered. Zinc was reversibly oxidized and re-
creased demand for lithium. However, the high cost of silver duced during the lithium-recovery process without side reac-
is one hindrance to facilitating the practical use of this elec- tions and weight loss.
trochemical lithium recovery. In this study, zinc, which has a

Introduction

Lithium consumption has dramatically increased worldwide Several candidate materials for replacing the silver elec-
owing to its versatile use in smartphones, electric automo- trode such as activated carbon, I@/I3@ redox couple, and
biles, and nuclear fusion reactors. The demand for lithium is nickel hexacyanoferrate (NiHCF) have been proposed.[5]
likely to continue its upward trend in the years ahead be- However, the capacity of activated carbon is fairly small
cause the electric vehicle market is predicted to grow. There- compared with that of a silver electrode.[5a] An I@/I3@ redox
fore, a stable supply of lithium resources has become a great couple appears to be inadequate in an aqueous system be-
concern, and global competition for securing lithium resour- cause an organic solvent is required as an electrolyte.[5b] The
ces has intensified.[1] systems with NiHCF appear to be energy-efficient battery-
Generally, lithium can be obtained from brines and miner- like systems, although the energy consumption varies de-
als. Of the two, lithium dissolved in brines accounts for more pending upon the choice of cathode materials such as LFP or
than 66 % of global lithium reserves, and at present most of LMO. However, it was noted that the cation concentration
the lithium is produced (about 83 %) from brines.[2] The most such as Na + , K + in the recovery solution needs to be proper-
conventional method for lithium production is lime soda ly adjusted all the time so that NiHCF does not react with
evaporation, which simply uses solar energy to evaporate Li + .[5c,d]
brine water. However, this method is time consuming and Herein, zinc, which is a low-cost, large-capacity material
strongly affected by climate factors such as rainfall, illumi- and stable in aqueous solution,[6] is proposed as an alterna-
nance, and humidity. Additionally, it requires a tremendously tive negative electrode material for an electrochemical lithi-
large site for its operation as well as a substantial amount of um recovery system.
chemicals and water to purify the lithium. These restraints
make it difficult to steadily supply lithium as well as cause
environmental problems.
As part of an effort to overcome the limitation of the con-
ventional method, Kanoh et al. first reported an electro- [a] S. Kim, Dr. J. Lee, Dr. S. Kim, Dr. S. Kim, Prof. Dr. J. Yoon
chemical lithium-recovery system using a l-MnO2/Pt elec- School of Chemical and Biological Engineering, Institute of Chemical Process-
trode, demonstrating the possibility of selective lithium re- es (ICP)
Seoul National University
covery from brines.[3] For a more efficient process, systems
World Class University (WCU) Program of Chemical Convergence for Energy
based on the principle of a lithium secondary batteries were & Environment (C2E2)
developed.[4] These systems include LiMn2O4 (LMO) or Gwanak-gu, Daehak-dong, Seoul 151-742 (Republic of Korea)
LiFePO4 (LFP) for lithium capture as a positive electrode E-mail: jeyong@snu.ac.kr
and silver as a negative electrode. Positive electrodes selec- [b] Dr. J. Lee, Dr. S. Kim, Prof. Dr. J. Yoon
Asian Insititute for Energy, Environment & Sustainability (AIEES)
tively recover lithium in a few hours, that is, lithium can be
Seoul National University
obtained consistently without any other purification process. Gwanak-gu, Daehak-dong, Seoul 151-742 (Republic of Korea)
However, the high price of silver is one of the obstacles to Supporting Information for this article can be found under:
achieving a large-scale system and practical application. https://doi.org/10.1002/ente.201700488.

Energy Technol. 2018, 6, 340 – 344 T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 340
Description of the electrochemical lithium recovery system with Results and Discussion
zinc
Figure 1 shows the results of the selective lithium recovery
Scheme 1 shows the two-step process of the electrochemical from artificial brine the chemical composition of which was
lithium recovery system with LMO and a zinc electrode intended to simulate that of the Salar de Atacama in Chile
(LMO-Zn system). The reactor of this system is divided into (refer to Table S2 in the Supporting Information) with the
two compartments by an anion exchange membrane (AEM). LMO-Zn system in which zinc is used as a negative elec-
The LMO electrode, which is used as a positive electrode, is trode. The suitability of zinc in the LMO-Zn system was con-
immersed in one compartment, and a zinc foil as a negative firmed with cyclic voltammetry (refer to Figure S1 in the
electrode is dipped into the other compartment, which is Supporting Information). As shown in Figure 1, lithium was
filled with a ZnCl2 solution with a concentration above 1 m selectively recovered from brine water and extracted into the
for reversible redox reactions (refer to Table S1 in the Sup- charging solution. On the other hand, other ions (Na + , K + ,
porting Information). and Mg2 + ) were not significantly recovered despite their
In the discharging step, the LMO electrode is filled with high concentration in the brine water. This result is attribut-
brine water (Scheme 1 a). Lithium ions in the brine are selec- ed to the preference of lithium insertion into the l-MnO2
tively captured by the l-MnO2 structure, and simultaneously spinel structure; insertion of lithium is structurally and ener-
the zinc metal is oxidized to zinc ions. These reactions occur getically favorable compared with other ions.[7] The lithium
spontaneously, that is, electrical energy is generated. The selectivity in the LMO-Zn system appears to be low in com-
overall reaction is as follows: parison with the previous study,[5d] and a good explanation
for this difference other than the difference of the solution
2 l-MnO2 þ Liþ þ ZnðsÞ Ð LiMn2 O4 þ Zn2þ þ e@ ð1Þ concentration between the two studies is not available at
present. The crossover of zinc ions through the anion-ex-
In the charging step, the LMO electrode is moved to the change membrane was 3 % (refer to Table S3 in the Support-
charging solution, which then releases Li + to obtain a highly ing Information for details).
pure lithium solution. At the same time, zinc ions are re-
duced to zinc metal (Scheme 1 b). This step is the energy-
consuming process [the reverse reaction of Eq. (1)]. The
charge neutrality of this system is balanced by the transfer of
anions through the AEM.

Figure 1. Selective lithium recovery with the LMO-Zn system in the presence
of other cations (Na + , K + , and Mg2 + ): the concentration changes in the
charging solution during 3 cycles (Atacama brine for lithium source water,
CaCl2 (0.1 m) for charging solution, current density: 0.5 mA cm@2).

Figure 2 shows the respective voltage profiles of the LMO


and zinc electrodes during the lithium recovery process. The
electrode potential of LMO varies with two plateaus whereas
lithium ions are intercalated and de-intercalated. On the
other hand, the electrode potential of zinc remains almost
constant during the discharging and charging steps with a
small overpotential. This result implies that it would be easy
to construct a two-electrode system for practical application
when zinc is used as a negative electrode because zinc can
Scheme 1. Diagrams of the electrochemical lithium recovery process with the
function as a potential reference for the system.
LMO-Zn system. (a) Discharging step: lithium ions are selectively captured
from the brine water, and zinc oxidation occurs. (b) Charging step: lithium Figure 3 shows the cell voltage of the LMO-Zn system
ions are released into the charging solution, and zinc reduction occurs. plotted in terms of the charge delivered during the lithium-

Energy Technol. 2018, 6, 340 – 344 T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 341
Figure 2. Respective voltage profiles of the LMO and zinc electrodes during Figure 4. Effect of the intensity of the current density on lithium selectivity in
the electrochemical lithium recovery process in the LMO-Zn system. the electrochemical lithium recovery process with the LMO-Zn system.

where cLi,C is the cationic mole fraction of Li + in the charg-


ing solution after the lithium recovery process, and cLi,B is
the cationic mole fraction of Li + in the brine water (Ataca-
ma brine). As shown in Figure 4, the lithium selectivity tends
to decrease with a higher current density. The lithium-selec-
tive coefficient is 14.4 for 0.5 mA cm@2, 9.2 for 0.75 mA cm@2,
and 6.6 for 1.0 mA cm@2. This result implies that as a higher
current density is applied, more impurity ions gain enough
energy for insertion and co-intercalate into the l-MnO2
structure.
To demonstrate the cyclability and stability of the LMO-
Zn system, 100 cycles of discharging–charging steps with a
current density of 1 mA cm@2 were repeated in the LMO-Zn
cell. Figure 5 shows the capacity retention of the LMO-Zn
cell during the cycles. After 100 cycles, the LMO-Zn cell re-
Figure 3. Voltage versus capacity profile during the electrochemical lithium re- tained 73 % of its initial capacity.
covery process in the LMO-Zn system. The area between the discharging- Capacity fading is an inevitable feature of the lithium-ion
charging curves shows the consumed energy during the process storage material owing to intercalation-induced stress and is
(6.3 Wh molLi@1).
significantly intensified when the system is exposed to
oxygen.[8] The high concentration of impurity ions in the
recovery process. The energy consumption per cycle was cal- electrolyte used in this study is presumed to be another
culated by the circular integration of the plot as follows:
I
W ¼ @ DEdq ð2Þ
c

where W is the energy (J), DE is the cell voltage (V), and q


is the delivered charge (C).
In terms of energy, 6.3 Wh is required to recover 1 mol of
lithium during the electrochemical lithium recovery process
with the LMO-Zn system (the mechanical energy consumed
for moving the LMO electrode was not considered).
Figure 4 shows how the lithium selectivity of the LMO is
affected by the applied current densities in the electrochemi-
cal lithium recovery process. The lithium-selective coefficient
(L) is defined as follows:
Figure 5. Capacity retention of the LMO-Zn cell during 100 cycles of the dis-
cLi;C
L¼ ð3Þ charging and charging step (Atacama brine for LMO side solution, ZnCl2
cLi;B (3.2 m) for zinc side, 1 mA cm@2).

Energy Technol. 2018, 6, 340 – 344 T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 342
factor for the decrease in capacity because the structural zinc-side solution was 1.8 m ZnCl2, which was determined experi-
stress is assumed to be enhanced by the insertion of other mentally for a favorable reversible redox reaction in discharg-
cations that have bigger ionic radii. The zinc electrode was ing–charging step. The voltage profiles of LMO and Zn during
the lithium recovery were measured simultaneously with two dif-
repeatedly oxidized and reduced during the 100 cycles and
ferent Ag/AgCl (KCl sat.) reference electrodes, which were lo-
maintained its mass (1675 mg). The zinc electrode was ana-
cated in a LMO-side compartment and a Zn-side compartment,
lyzed by X-ray diffraction to confirm whether the reversible respectively. The potential of LMO was measured with a three-
redox reaction between Zn2 + and Zn0 occurs (Figure S2 in electrode system, which was comprised of LMO as working elec-
the Supporting Information). Peaks for zinc metal were trode, Zn as counter electrode, and Ag/AgCl (KCl sat.) in the
clearly detected; as a result, the zinc electrode showed great LMO-side compartment as reference electrode. At the same
stability during long-term operation. time, the open circuit potential between Zn and another Ag/
AgCl reference in the Zn-side compartment was measured. Ion
concentrations in the charging solution were measured by ion
Conclusions chromatography (DX-120, DIONEX, USA). Cyclic voltammetry
was performed to confirm the redox properties of zinc with a po-
This study demonstrated the feasibility of an electrochemical tentiostat (PARSTAT 2273, Princeton Applied Research, USA)
lithium recovery system using zinc as a negative electrode, in a potential range from @0.8 V to @1.9 V and a scan rate of
which is an inexpensive and large-capacity material. In the 5 mV s@1. Zinc foil and Ag/AgCl (KCl sat.) were used as the
LiMn2O4–zinc (LMO-Zn) battery system, lithium was selec- working and reference electrode, respectively. The Ag/AgCl
tively recovered from brine water that contained high con- counter electrode was prepared with the same method used in
the previous study.[4b]
centrations of impurities with an energy consumption of
For the long-term stability test of the LMO-Zn system, capacity
6.3 Wh per 1 mol of lithium recovered. By storing a high con- retention was measured repeatedly up to 100 times between
centration of ZnCl2 in the confined compartment, the zinc 1.93 V and 1.38 V as Atacama brine was in the LMO-side com-
electrode was reversibly oxidized and reduced without side partment regardless of discharging–charging step (Atacama brine
reactions and electrode-potential variation. The stability of was not replaced even in the charging step) and 3.2 m ZnCl2 was
the LMO-Zn system was examined through 100 cycles of dis- in the zinc electrode side. This long-term stability test focused on
charging–charging, and the LMO-Zn cell retained 73 % of its whether the capacity retention is maintained. The high concen-
initial capacity without loss of zinc. tration of ZnCl2 (3.2 m) was chosen to minimize the osmotic pres-
sure between the two compartments. XRD patterns of the zinc
electrode after the long-term stability test of the LMO-Zn
system were obtained with the Rigaku D-MAX2500, Japan.
Experimental Section
The LMO-Zn system in this study consisted of the LMO elec-
trode, a zinc electrode, and an anion-exchange membrane Acknowledgments
(AEM; AMX, ASTOM Co., Japan). AEM was used in the
LMO-Zn system as the production cost of AEM continues to be This research was supported by a grant (code 17IFIP-
greatly lowered to 4.84 $ m@2 owing to its mass production and B065893-05) from the Industrial Facilities & Infrastructure
technological developments in recent years.[9] The distance be- Research Program funded by Ministry of Land, Infrastructure
tween the LMO and zinc was 5 cm. The LMO electrode was fab-
and Transport of Korean government and Basic Science Re-
ricated by mixing LiMn2O4 (Sigma–Aldrich, Korea), carbon
black (Super P, Timcal, Switzerland), and polytetrafluoroethylene search Program through the National Research Foundation of
(PTFE, Sigma–Aldrich, Korea) at a weight ratio of 8:1:1 with Korea (NRF) funded by the Ministry of Education (NRF-
ethanol. The resulting mixture was roll-pressed to 300 mm thick- 2016R1D1A1A02937469), and by the “Foundation of IN-
ness and dried in a vacuum oven at 80 8C for 12 h to evaporate NOPOLIS Daegu Project” funded by Ministry of Science,
the solvent. The dried electrode was cut into 2.5 cm by 4 cm ICT and Future Planning of the Korean government
shape and attached to a titanium foil with carbon paint (DAG-T- (2015DG0016).
502, Ted Pella, USA). Zinc foil (Alfa Aesar, Korea), which was
cut to the same dimensions as the LMO electrode without any
other fabrication processes, was used as the negative electrode. Conflict of interest
The properties of the AEM, which separates the reactor of the
LMO-Zn system into two compartments, are described in
The authors declare no conflict of interest.
Table S4 in the Supporting Information.
The LMO electrode was converted into the l-MnO2 phase with
an electrochemical method, the form of which is susceptible to
capture lithium ions in the brine. The aqueous solution on the Keywords: batteries · lithium manganese oxide · lithium
LMO electrode side was Atacama brine (50 mL) in the discharg- recovery · membranes · zinc
ing step so that the l-MnO2 is supposed to capture lithium ions
selectively, and in charging step, it was replaced with CaCl2
[1] a) P. W. Gruber, P. A. Medina, G. A. Keoleian, S. E. Kesler, M. P. Ev-
(0.1 m) for quantifying the amounts of lithium ions released from erson, T. J. Wallington, J. Ind. Ecol. 2011, 15, 760 – 775; b) W. Tahil,
the LMO electrode. The cycle of discharging–charging, which is Implications of Future PHEV Production for Lithium Demand. Mar-
the lithium recovery process, were repeated galvanostatically tainville, Meridian International Research, 2007; c) L. Legers, Meridian
with a battery cycler (WBCS3000, WonATech, Korea) with a International Research, May 29, 2008.
constant current of 0.5 mA cm@2 in the 1.38–1.93 V range. The [2] J.-M. Tarascon, Nat. Chem. 2010, 2, 510.

Energy Technol. 2018, 6, 340 – 344 T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 343
[3] a) H. Kanoh, K. Ooi, Y. Miyai, S. Katoh, Langmuir 1991, 7, 1841 – [7] a) R. T. Shannon, Acta Crystallogr. Sect. A 1976, 32, 751 – 767; b) D. W.
1842; b) H. Kanoh, K. Ooi, Y. Miyai, S. Katoh, Sep. Sci. Technol. 1993, Smith, J. Chem. Educ. 1977, 54, 540.
28, 643 – 651. [8] a) M. Y. Song, D. S. Ahn, H. R. Park, J. Power Sources 1999, 83, 57 –
[4] a) M. Pasta, A. Battistel, F. La Mantia, Energy Environ. Sci. 2012, 5, 60; b) P. He, J.-L. Liu, W.-J. Cui, J.-Y. Luo, Y.-Y. Xia, Electrochim.
9487 – 9491; b) J. Lee, S.-H. Yu, C. Kim, Y.-E. Sung, J. Yoon, Phys. Acta 2011, 56, 2351 – 2357.
Chem. Chem. Phys. 2013, 15, 7690 – 7695; c) R. Trlcoli, A. Battistel, [9] H. Lu, W. Zou, P. Chai, J. Wang, L. Bazinet, J. Cleaner Prod. 2016,
F. L. Mantia, Chem. Eur. J. 2014, 20, 9888 – 9891. 112, 3097 – 3105.
[5] a) S. Kim, J. Lee, J. S. Kang, K. Jo, S. Kim, Y.-E. Sung, J. Yoon, Che-
mosphere 2015, 125, 50 – 56; b) J.-S. Kim, Y.-H. Lee, S. Choi, J. Shin,
H.-C. Dinh, J. W. Choi, Environ. Sci. Technol. 2015, 49, 9415 – 9422;
c) R. Trlcoli, A. Battistel, F. La Mantia, ChemSusChem 2015, 8, 2514 –
2519; d) R. Trlcoli, C. Erinmwingbovo, F. La Mantia, ChemElectro- Manuscript received: July 18, 2017
Chem 2017, 4, 143 – 149. Revised manuscript received: August 21, 2017
[6] G. Toussaint, P. Stevens, L. Akrour, R. Rouget, F. Fourgeot, ECS Accepted manuscript online: August 30, 2017
Trans. 2010, 28, 25 – 34. Version of record online: January 17, 2018

Energy Technol. 2018, 6, 340 – 344 T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 344

You might also like