You are on page 1of 9

Journal of Environmental Management 310 (2022) 114705

Contents lists available at ScienceDirect

Journal of Environmental Management


journal homepage: www.elsevier.com/locate/jenvman

Research article

A fundamental study on selective extraction of Li+ with


dibenzo-14-crown-4 ether: Toward new technology development for
lithium recovery from brines
Yanhang Xiong a, Tao Ge a, Liang Xu a, b, *, Ling Wang a, Jindong He a, Xiaowei Zhou a,
Yongpan Tian a, b, Zhuo Zhao a, b, **
a
School of Metallurgical Engineering, Anhui University of Technology, Ma’anshan, 243032, PR China
b
Low-Carbon Research Institute, Anhui University of Technology, Ma’anshan, 243032, PR China

A R T I C L E I N F O A B S T R A C T

Keywords: The present study has proposed a selective Li+ extraction process using a novel extractant of dibenzo-14-crown-4
Supramolecular chemistry ether functionalized with an alkyl C16 chain (DB14C4–C16) synthesized based on the ion imprinting technology
Lithium recovery (IIT). Theoretical analysis of the possible complexes formed by DB14C4–C16 with Li+ and the competing ions of
Crown ether
Na+, K+, Ca2+ and Mg2+ was performed through density functional theory (DFT) modeling. The Gibbs free
Solvent extraction
Ion imprinting
energy change of the complexes of metal ions with DB14C4–C16 and water molecules were calculated to be
− 125.81 and − 166.01 kJ/mol for lithium, − 55.73 and − 117.77 kJ/mol for sodium, and − 196.02 and − 291.52
kJ/mol for magnesium, respectively. Furthermore, the solvent extraction experiments were carried out in both
single Li+ and multi-ions containing solutions, and the results delivered a good selectivity of DB14C4–C16 to­
wards Li+ over the competing ions, showing separation coefficients of 68.09 for Ca2+-Li+, 24.53 for K+-Li+, 16.32
for Na+-Li+, and 3.99 for Mg2+-Li+ under the optimal conditions. The experimental results are generally in
agreement with the theoretical calculations.

1. Introduction is followed by 19.3 million tons for Argentina, 9.6 million tons for Chile,
7.9 million tons for the United States, 6.4 million tons for Australia, and
Lithium is an important raw material for lithium-ion battery pro­ 5.1 million tons for China (USGS, 2021). Economical concentrations of
duction, of which the global demand has increased drastically in recent lithium are usually found in hard rock resources such as spodumene,
years due to the technological advancements in electric vehicles and lepidolite, petalite and amblygonite, and hydro-mineral resources of
electronic devices requiring batteries of high energy-density (Para­ brines and seawater. Among these primary resources, brines are the
nthaman et al., 2017; Yao et al., 2018; Zhao et al., 2021; Wu et al., largest lithium resources, accounting for about 70–80% of global lithium
2021). In addition, lithium is widely used in many other industries such deposits (Gu et al., 2018; Barandiarán, 2019; Liu et al., 2020, 2021).
as primary aluminum production, lubricating greases, ceramics and However, Li+ normally co-exists with other alkali metal and alkali earth
glass, owing to it’s a variety of attractive properties (Swain, 2017; Liu metal ions such as Na+, K+, Ca2+, and Mg2+ of much higher concen­
and Azimi, 2021; Kablov et al., 2021). As a result of the great role of trations compared with lithium in the brines. These co-existing ions,
lithium in energy, metallurgy, chemical industry, and economic sectors, particularly Mg2+, has quite similar chemical properties to that of Li+,
it has gained importance as a near-critical element (U.S. Department of which always results in technical difficulties for the selective extraction
Energy, 2011; Sener et al., 2021). of lithium from brines (Guo et al., 2018a, 2018b). Therefore, Mg/Li ratio
The average content of lithium in the earth’s crust has been esti­ (weight ratio of magnesium to lithium) is the key factor determining the
mated to be approximately 0.007% (Swain, 2017). The latest report on extraction process selection and operation difficulty. The conventional
identified lithium reserves worldwide is around 86 million tons, with solar evaporation-precipitation process has limited applicability for
Bolivia being the largest lithium resource country (21 million tons). This lithium extraction from brines once the Mg/Li ratio is above 6, however,

* Corresponding author. Hudong Road 59, Ma’anshan, Anhui Province, 243032, PR China.
** Corresponding author. Hudong Road 59, Ma’anshan, Anhui Province, 243032, PR China.
E-mail addresses: liangxu_ahut@163.com (L. Xu), nonferrous_ahut@163.com (Z. Zhao).

https://doi.org/10.1016/j.jenvman.2022.114705
Received 8 December 2021; Received in revised form 24 January 2022; Accepted 8 February 2022
Available online 22 February 2022
0301-4797/© 2022 Elsevier Ltd. All rights reserved.
Y. Xiong et al. Journal of Environmental Management 310 (2022) 114705

the majority of known brines worldwide generally have high Mg/Li comparison with the experimental results. The present study aims to
ratios of around 35 (Song et al., 2017; Liu et al., 2018, 2019). Moreover, develop an efficient and cost-effective solvent extraction system for
the other disadvantages of the solar evaporation-precipitation method lithium recovery from brines, and more importantly, to provide theo­
such as long-time operation (up to 12–18 months), poor extraction ef­ retical basis for the applications of supramolecular chemistry in envi­
ficiency, and large waste production also limit its industrial applications ronment, resource recovery and recycling.
(Kim et al., 2015; Zhao et al., 2019). Therefore, it will be of great sig­
nificance to develop more efficient and cost-effective lithium extraction 2. Experimental and computational methods
processes from brines serving the sustainable development of lithium
related industries. 2.1. Computational modeling
There have been numerous efforts to develop innovative technolo­
gies such as ion-sieve adsorption (Wang et al., 2017; Zhao et al., 2018), All the DFT calculations in the present study were conducted with
membrane separation (Li et al., 2017; Zhang et al., 2017), electro­ Gauss 16 and Multiwfn 3.6 program package (Lu and Chen, 2012; Cao
chemical approaches (He et al., 2018; Kim et al., 2018), solvent et al., 2015). Geometrical structure optimization of the complexes
extraction (Shi et al., 2016, 2018), as well as separation technologies formed of DB14C4–C16 with Li+ and its competing ions of Na+, K+,
based on photocatalytic nanocomposites (Yousefi et al., 2019a, b; You­ Ca2+, and Mg2+ was carried out with M062X/def2SVP, and the corre­
sefi et al., 2021a, b, c). Among them, solvent extraction is an established sponding single point energy calculations were performed at
process for efficient separation and extraction of a variety of metals from M062X/def2TZVP level (Ho et al., 2010; Golebiowski et al., 2010;
aqueous media owing to its simplicity and convenient handling (Kumar Achazi et al., 2016). The optimized complex structures were further
et al., 2010; Torrejos et al., 2015; Swain, 2016). For lithium recovery confirmed to be real minima according to the frequency calculations.
from brines, the selectivity of the ligand towards Li+ over the co-existing The influence of aqueous solvent was taken into account using solvation
ions of Na+, K+, Ca2+ and particularly Mg2+ is essential to the feasibility model based on density (SMD) (Marenich et al., 2009). VMD was uti­
and economic benefits of the solvent extraction process. Crown ether is a lized to achieve the visualization of the molecules and complex in­
sort of macrocyclic ligand possessing excellent complexing affinity to teractions. The energy decomposition analysis was conducted based on
metal ions, which is able to form complexes through ion-dipole inter­ symmetry-adapted perturbation theory (SAPT) with PSI4 modeling
action of the negatively charged oxygen atoms with metal ions (in program at the SAPT2+/aug-cc-pVDZ level (Turney et al., 2012; Parker
particular alkali metal and alkali earth metal ions) positioned in the et al., 2014; Parrish et al., 2017). The Gibbs free energy change of the
center of the crown ether ring. The stability of the formed complex is complexation process of DB14C4–C16 with the metal ions was evaluated
significantly affected by the relative size of the crown ether cavity to the according to Eq. (1).
target metal ion, which means the selectivity of crown ether towards the
target metal ion is mainly determined by the cavity-cation size ΔG ​ = ​ GDB14C4− C16− M ​ − ​ (GM ​ + ​ GDB14C4− C16 ) (1)
compatibility (Sun et al., 2020; Zhao et al., 2020). Li+ has an ionic
Where GDB14C4–C16, GDB14C4–C16-M and GM are the Gibbs free energies of
diameter of 1.36 Å, which matches best the cavity size of 14-crown-4
DB14C4–C16, DB14C4–C16-M and metal ion (M = Li+, Na+, K+, Ca2+,
(14C4, 1.2–1.5 Å) (Pedersen, 1967; Torrejos et al., 2015). Therefore,
and Mg2+), respectively.
14C4 and its derivatives are expected to have the highest selectivity to
Li+ among the know crown ethers. On the other hand, 14C4 exhibits
poor complexing affinity towards most of the competing ions such as 2.2. Synthesis of DB14C4–C16
Na+ (2.04 Å), K+ (2.76 Å), and Ca2+ (1.98 Å) because of the incom­
patible size of the crown ether cavity with these cations. Although Mg2+ The synthesis of DB14C4–C16 follows the route shown in Fig. 1. It is
(1.42 Å) has the nearly identical ion diameter to that of Li+ (1.36 Å), its a three-step synthesis process including the successive reactions of
complexation with 14C4 or 14C4 derivatives during the solvent nucleophilic substitution for 1,3-bis(2-hydroxyphenoxy)propane (in­
extraction process is possibly to be limited in terms of hard and soft acid termediate 1) formation, ring closure for hydroxy-DB14C4 (intermedi­
and base (HSAB) theory, steric hindrance, as well as hydration effect ate 2) generation, and alkylation with 1-bromohexadecane for the
(Rodriguez and Lisy, 2011; Swain, 2016; Zhao et al., 2020; Tian et al., production of DB14C4–C16. All the synthesis procedures were pro­
2020). ceeded under the protection of continuous high-purity N2 flow. The
In the present paper, a novel solvent extraction system using product of each step (intermediate 1, intermediate 2, and DB14C4–C16)
dibenzo-14-crown-4 ether functionalized with an alkyl C16 chain was dissolved in CH2Cl2, subsequently washed with deionized water,
(DB14C4–C16) as extractant for the selective recovery of lithium from and finally dehydrated with anhydrous Na2SO4 for complete
brines has been proposed. The two benzene rings are able to increase the purification.
rigidity of 14C4, retaining the cavity size and minimizing the energy loss
during the complexation process. The long lipophilic C16 alkyl chain 2.2.1. Synthesis of intermediate 1
could effectively decrease the dissolution loss of crown ether to the 2 g (18 mmol) catechol was dissolved in 20 ml anhydrous acetonitrile
aqueous phase and increase the extractant loading in the organic in a round-bottom flask, and 0.87 g (36 mmol) LiOH was introduced into
diluent, which could consequently improve the lithium extraction ca­ the system after the complete dissolution of catechol. Subsequently, 1,3-
pacity of the solvent extraction system (Torrejos et al., 2016). The dibromopropane was added dropwise in the flask over a period of 2 h
synthesis of DB14C4–C16 was based on the ion imprinting technology under magnetic stirring. After the completion of 1,3-dibromopropane
(IIT) using Li+ as the template. In this way, specific shape and size addition, the system temperature was increased to 100 ◦ C and the re­
memory between the imprinted crown ether cavity and Li+ is generated, action was proceeded for 20 h under refluxing. The system was cooled to
further enhancing the selectivity of DB14C4–C16 to Li+ over the room temperature after the termination of reflux reaction, and the so­
competing ions especially Mg2+. A series of solvent extraction experi­ lution pH was adjusted to 4–5 with HCl. The solvent of acetonitrile was
ments based on the synthesized DB14C4–C16 were carried to investigate removed by vacuum evaporation for the recovery of powdery interme­
the effects of extractant loading, phase ratio (O/A), temperature, solu­ diate 1 (45% yield).
tion pH and reaction time on the selective Li+ uptake process. Moreover,
the structural and thermodynamic properties of the possible complexes 2.2.2. Synthesis of intermediate 2
formed by the interactions of DB14C4–C16 with Li+ and the co-existing 1.95 g (7.5 mmol) intermediate 1, 0.36 g (15 mmol) LiOH and 180 ml
ions of Na+, K+, Ca2+ and Mg2+ were analyzed through density func­ deionized water were homogeneously mixed in a round-bottom flask at
tional theory (DFT) modeling. The theoretical calculations were in 95 ◦ C under magnetic stirring. The temperature was then decreased to

2
Y. Xiong et al. Journal of Environmental Management 310 (2022) 114705

Fig. 1. Synthesis route of DB14C4–C16.

50 ◦ C and 0.69 g (7.5 mmol) epichlorohydrin (ECH) was dropwise added coefficient (S) were used to evaluate the solvent extraction process, as
into the system over 3 h. After the ECH addition, the system was expressed in Eqs. (2)–(4).
continuously stirred at 50 ◦ C for 10 h under refluxing. The reflux reac­ ( )
E ​ = ​ Ci ​ − ​ Cf ​ / ​ Ci ​ × ​ 100% (2)
tion was repeated by adding another batch of 0.36 g (15 mmol) LiOH
and 0.69 g (7.5 mmol) ECH under the same experimental conditions to ( ) ( )
D ​ = ​ Ci ​ − ​ Cf ​ / ​ Cf ​ × ​ V aq ​ / ​ V org (3)
enhance the ring closure reaction producing hydroxy-DB14C4. After­
wards, the system was cooled to room temperature, and the powdery
S ​ = ​ DLi ​ / ​ DM (4)
intermediate 2 with a yield of 70% was obtained by liquid-solid
separation.
Where Ci and Cf are the metal ions concentrations in the aqueous phase
before and after extraction, respectively. Vaq and Vorg represent the
2.2.3. Synthesis of DB14C4–C16
volumes of the aqueous and organic phases, respectively.
100 ml anhydrous tetrahydrofuran (THF) and 0.29 g (12 mmol) NaH
Overall, the methodology of this study has been summarized, as seen
were introduced into a round-bottom flask, and 1.9 g (6 mmol) inter­
in Fig. 2.
mediate 2 dissolved in 25 ml anhydrous THF was subsequently added
into the flask drop by drop over a period of 1 h. The system was
3. Results and discussion
continuously stirred at room temperature for another 1.5 h after inter­
mediate 2 addition. Afterwards, 3.7 g (12 mmol) 1-bromohexadecane
3.1. Computational modeling
was dissolved in 25 ml anhydrous THF, which was then added into
the flask dropwise over 1 h, and the system was kept magnetically stirred
3.1.1. Geometrical structure optimization
for 16 h to guarantee the attachment of long lipophilic C16 alkyl chain
The optimized structures of free DB14C4–C16 and its complexes
on the crown ether. After the reaction, methanol was slowly added in the
formed with Li+, Na+, K+, Ca2+, and Mg2+ from the top and side views
system to remove the residue NaH, and the THF solvent was separated
are shown in Fig. 3. The complex structure is determined by the
through vacuum evaporation. In this way, Li+ imprinted DB14C4–C16
matching degree of the metal ion with crown ether. It is seen in Fig. 3
was obtained with a yield of 70%. The structure of the synthesized
that Li+ and Mg2+ lie well in the plane of oxygen atoms, because the
DB14C4–C16 was determined through nuclear magnetic resonance
ionic diameters of Li+ (1.36 Å) and Mg2+ (1.42 Å) fit best the cavity size
analysis (NMR, AVANCE 400, Bruker, Switzerland) and mass spectro­
of DB14C4–C16 (1.2–1.5 Å) among the five cations. On the other hand,
graphic analysis (GC-MS, 2010 PLUS, Shimadzu, Japan).
1 Ca2+, Na+, and K+ have larger ionic diameters of 1.98 Å, 2.04 Å, and
H NMR (400 MHz, CDCl3): δ (ppm) 0.8–0.9 (m, 3H, –CH3), 1.2 (s,
2.76 Å, respectively, showing poor size compatibility with the crown
26H, –CH2-), 1.6 (m, 2H, –CH2-), 2.2–2.3 (s, 2H, –CH2-), 3.6 (m, 2H,
ether. Therefore, these cations are not able to enter the cavity of
–CH2-), 4.0 (m, 1H, –CH–), 4.2 (m, 8H, –CH2-), 6.9–7.0 (m, 8H, ArH).
DB14C4–C16 during the complexation process, lying much farther
ESI-HRMS: m/z 541.389 for [M + H]+ (540.382 calculated for
beyond the plane of oxygen atoms compared with Li+ and Mg2+.
C34H52O5).

2.3. Solvent extraction experiments

2.3.1. Experimental procedure


Certain amount of the synthesized DB14C4–C16 was dissolved in
CH2Cl2 to serve as the organic phase in the solvent extraction experi­
ments. The solutions containing pure Li+ or in the presence of competing
ions such as Na+, K+, Ca2+ and Mg2+ were employed as the aqueous
samples with each cation concentration being 200 mg/L to simulate the
composition of low salinity brines, from which the selective extraction of
Li+ is exceedingly challenging. The solvent extraction experiments were
conducted in an air-bathing thermostatic vibrator with a speed of 300
rpm. The effect of varied conditions such as DB14C4–C16 concentration,
phase ratio (O/A), temperature, solution pH and reaction time were
systematically investigated. After extraction, the organic and aqueous
phases were completely separated with a separating funnel for further
characterizations.

2.3.2. Statistical analysis


The concentrations of metal ions in the solution before and after
extraction were analyzed using an inductively coupled plasma optical
emission spectrometer (ICP-OES, ICPS-7510 PLUS, Shimadzu, Japan).
Extraction efficiency (E), distribution coefficient (D), and separation Fig. 2. Schematic diagram of the methodology of the present study.

3
Y. Xiong et al. Journal of Environmental Management 310 (2022) 114705

Fig. 3. Optimized structures of DB14C4–C16 and its complexes with Li+, Na+, K+, Ca2+, and Mg2+.

The geometric parameters of the complexes between DB14C4–C16 occupied molecular orbital (HOMO) and lowest occupied molecular
and the metal ions are listed in Table S1. As the complexes are formed orbital (LUMO) energies as well as HOMO-LUMO gap have been
through the coordination of the metal ions with the oxygen atoms in the calculated based on the corresponding wave function documents
crown ring of DB14C4–C16, the focus is mainly placed on the bonds generated by Gauss 16. A higher HOMO-LUMO gap generally means a
between the donor oxygen atoms and the cations (M-O). A shorter M-O better stability of a complex (Pereira et al., 2016). Fig. 4 displays the
bond length generally represents a higher complexation affinity. The three-dimensional HOMO and LUMO plots of DB14C4–C16 before and
bond length of Li+ with oxygen atom (Li+-O) shows the lowest value of after complexation with Li+. The HOMO and LUMO of free DB14C4–C16
1.94–1.95 Å among the five cations, indicating the highest affinity of the was calculated to be about − 7.12 eV and 8.87 eV, respectively, resulting
donor oxygen atoms in DB14C4–C16 towards Li+ in the presence of the in a HOMO-LUMO gap of 7.99 eV, which is 0.17 eV higher than its
competing ions. The Mg2+-O bond length is slightly larger than that of complex formed with Li+. This indicates a slight decrease in stability by
Li+-O bond, as to be in the range of 2.02–2.04 Å. This is followed in the formation of DB14C4–C16–Li+ complex.
sequence by 2.25–2.26 Å for Na+-O, 2.49–2.54 Å for Ca2+-O, and The calculated frontier molecular orbital energies of the complexes
2.64–2.67 Å for K+-O, mainly due to the poor size-compatibility of these formed between DB14C4–C16 and the cations of Li+, Na+, K+, Ca2+, and
cations with the cavity of DB14C4–C16. Mg2+ are shown in Table S2. It is observed that the complexes con­
taining most of the cations such as K+, Na+ and Ca2+ exhibit relatively
3.1.2. Molecular orbital analysis low HOMO-LUMO gaps of 7.12, 7.49 and 7.58 eV, respectively, sug­
In order to clarify the stability change of DB14C4–C16 before and gesting poor stabilities of DB14C4–C16–K+, DB14C4–C16–Na+, and
after complexation with Li+, Na+, K+, Ca2+, and Mg2+, the highest DB14C4–C16–Ca2+ complexes. It is worth noting that the HOMO-LUMO

Fig. 4. Three-dimensional plots of HOMO and LUMO of DB14C4–C16 and DB14C4–C16–Li+.

4
Y. Xiong et al. Journal of Environmental Management 310 (2022) 114705

gap of DB14C4–C16–Mg2+ is higher than that of DB14C4–C16–Li+ interaction energies between metal ions and water become more nega­
(7.82 eV), as to be 8.20 eV, which indicates a stronger complexation tive with the increase of the number of water molecules. And more
affinity of DB14C4–C16 towards Mg2+ compared with Li+ from the importantly, the total energies of water-Mg2+ complexes are much
perspective of molecular orbital theory. However, the distortion of the higher than that of water-Li+ and water-Na+ complexes. This indicates a
crown ether during its complexation with metal ions will probably high hydration effect of Mg2+ that water molecules have a strong
change the molecular orbitals, which could not be taken into consider­ complexation affinity toward Mg2+ in the aqueous phase, which
ation in the molecular orbital analysis. This will limit its reference value significantly hinders the complexation of Mg2+ with DB14C4–C16
to some extent for the practical selective extraction of Li+ in the presence transferring to the organic phase. This is beneficial for the selective
of competing ions of Na+, K+, Ca2+, and particularly Mg2+. lithium extraction process (Xu et al., 2021; Li et al., 2022).
The thermodynamic properties of the complexes formed by metal
3.1.3. Interaction and energy analysis ions with DB14C4–C16 and water were further evaluated, and the cor­
The interactions between DB14C4–C16 and Li+ during the selective responding results are shown in Table S4. It is seen that the Gibbs free
lithium extraction process were theoretically analyzed based on the energies of DB14C4–C16–Li+, DB14C4–C16–Na+, and
2+
independent gradient model (IGM), and the corresponding results are DB14C4–C16–Mg are of negative values, indicating that the
shown in Fig. 5. It is clear to see that Li+ principally interacts with the complexation between DB14C4–C16 and Li+, Na+, and Mg2+ could
donor oxygen atoms in the crown ring. The blue and green colors occur spontaneously. Moreover, for the complexes formed by metal ions
represent strong electrostatic and van der Waals interactions, respec­ with either DB14C4–C16 or water, the ΔG values follow the order of
tively, indicating that strong electrostatic and van der Waals are the Mg2+ < Li+ < Na+, which exhibit the same trend as that of the total
main interactions between DB14C4–C16 and Li+. The alkyl C16 chain on interaction energy analysis. This means although DB14C4–C16–Mg2+
DB14C4–C16 was not participated in the complexation process. has the highest stability among the three crown ether complexes, the
The theoretical calculations of interaction energies of metal ions with effect of hydration will result in the largest resistance for the complex­
both the extractant (DB14C4–C16) and solvent (H2O) were further ation of Mg2+ with DB14C4–C16 in terms of Mg2+ migration from the
performed on the basis of the symmetry-adapted perturbation theory aqueous solution towards the organic phase. Therefore, it is difficult to
(SAPT). The interaction energy was quantitatively dissected into four exactly identify if Li+ could be selectively extracted by DB14C4–C16
components of electrostatics, exchange, induction, as well as dispersion, only from the point of view of theoretical evaluation. Further experi­
and the corresponding results are shown in Table S3. The focus was mental work is essentially needed to clarify the feasibility of selective
placed on Li+ and its major competing ions of Na+ and Mg2+, whereas Li+ extraction in the presence of competing ions such as Na+, K+, Ca2+,
K+ and Ca2+ which are known to be easily separated from Li+ were not and particularly Mg2+.
taken into account in the SAPT calculations.
As can be seen from Table S3, for the interactions of metal ions with 3.2. Solvent extraction of Li+
both DB14C4–C16 and H2O, the largest stabilizing contribution is
electrostatics among the four components, which is followed by induc­ 3.2.1. Extraction in single Li+ containing solution
tion and dispersion successively. The exchange terms are generally of The solvent extraction results in single Li+ containing solution are
positive values, representing the adverse contribution of exchange to the shown in Fig. 6. It is clearly seen that the lithium extraction proceeded
complexation process. For the interactions with crown ether, the total rapidly at the beginning and the uptake of Li+ became limited by further
energy of DB14C4–C16–Mg2+ was calculated to be − 1082.97 kJ/mol, prolonging the reaction time after 20 min, resulting in a highest lithium
followed by − 419.19 kJ/mol for DB14C4–C16–Li+, and − 268.67 kJ/ extraction efficiency of about 7%. In order to guarantee the sufficient
mol for DB14C4–C16–Na+. This means the stability of the complex contact between the organic and aqueous phases facilitating the Li+
formed by DB14C4–C16 with Mg2+ is higher than that with Li+ and Na+. extraction process, a reaction time of 30 min was employed in the
However, it is worth noting that the solvent extraction process involves following extraction experiments.
the transport of metal ions from the aqueous solution towards the The extraction efficiency of Li+ was slightly enhanced by increasing
organic phase, which means the extraction efficiency is not only domi­ O/A from 1:1 to 2:1, because a higher O/A generally results in larger
nated by the selectivity of crown ether to the target ions, but also by the concentration differences between the reaction interface and organic
effect of hydration (Saito et al., 2010; Mason et al., 2015; Lommelen bulk, which could facilitate the diffusions of the reactant (DB14C4–C16)
et al., 2020; Li and Binnemans, 2020). In other words, the complexation and product (DB14C4–C16–Li+) within the organic phase in terms of
of metal ion with crown ether at the interface of aqueous and organic kinetics (Xu et al., 2020). The lithium extraction was hardly enhanced
phases is basically a competition process between water molecules and by further increasing the O/A ratio, indicating that the O/A effect on
crown ether for the target metal ion. It is seen from Table S3 that the lithium extraction became inconspicuous at higher O/A levels. There­
fore, O/A of 2:1 is considered to be suitable for lithium extraction from
the point of view of process efficiency.
The extraction efficiency of Li+ slightly decreased by approximately
2% with the increase of temperature from 10 to 30 ◦ C, indicating that a
lower temperature is favorable for the extraction of lithium by
DB14C4–C16. The theoretical estimations made in section 3.1.3
(Table S4) delivers an enthalpy change (ΔH) of about − 162.55 kJ/mol
for the complexation between DB14C4–C16 and Li+. The negative ΔH
value indicates that the coordination of DB14C4–C16 with Li+ at the
organic-aqueous interface is an exothermic process, which is in consis­
tence with the experimental results. In general, the extraction of Li+ was
not significantly affected by temperature in the range of 10–30 ◦ C.
Therefore, a reaction temperature of 20 ◦ C (close to the room temper­
ature) was applied in the following extraction experiments for easier
operations.
It is observed from the effect of extractant loading in the organic
phase that the Li+ extraction efficiency increased linearly from below
Fig. 5. IGM image of DB14C4–C16–Li+. 1% to about 7% with the increase of DB14C4–C16 to Li+ mole ratio in

5
Y. Xiong et al. Journal of Environmental Management 310 (2022) 114705

Fig. 6. Effects of various parameters on Li+ extraction process.

the range of 0.1–1.0. A larger DB14C4–C16/Li+ ratio means a higher


DB14C4–C16 concentration in the organic phase at a fixed O/A ratio of Kex = [nDB14C4–C16– Li+]org / [Li+]aq [DB14C4–C16]norg (6)
2, which normally results in the enhancement of both the activity and
quantity of the crown ether extractant, leading to positive impacts on DLi = Kex [DB14C4–C16]norg (7)
the extraction of Li+. The extraction efficiency was slightly influenced by log DLi = log Kex + n log [DB14C4–C16]org (8)
further increasing the DB14C4–C16/Li+ ratio, indicating the sufficient
supply of the DB14C4–C16 extractant in the organic phase for the The logarithmic plot of DLi with respect to DB14C4–C16 concentra­
lithium extraction process. tion in the organic phase based on the data in Fig. 6 is shown in Fig. 7. A
Solution pH is one of the most important parameters influencing the good linear relationship of the plotting results is clearly evident with the
lithium extraction process. It can be observed from the dependence of correlation coefficient R2 of over 0.99. Therefore, based on Fig. 7, the
Li+ extraction efficiency on the solution pH that the extraction of Li+ slope of 1.059 indicates a 1:1 coordination of DB14C4–C16 with Li+
was significantly enhanced by increasing the solution pH from 1 to 7, during the lithium extraction process, and the y-intercept delivers the
indicating the adverse effect of hydrogen ions on the lithium extraction extraction constant (log Kex) of 4.548.
process. The presence of large amount of hydrogen ions in the aqueous
solution could hinder the uptake of lithium ions during the extraction
process because H+ will compete against Li+ to coordinate with the
crown ether at the reaction interface. Therefore, it is clear that the
complexation of DB14C4–C16 towards Li+ is preferentially occurring at
relatively high pH levels, indicating the excellent application prospects
of DB14C4–C16 in lithium uptake from brines.
It has been reported that the complexes of both 1:1 and 2:1 ligand to
metal stoichiometry are possibly formed through the interactions be­
tween crown ethers and univalent metal ions (Asfari et al., 1996; Roper
et al., 2008). In the present study, the complexation mechanism between
DB14C4–C16 and Li+ during the lithium extraction process was inves­
tigated according to a slope analysis based on Eq. (5) (Torrejos et al.,
2016; Roper et al., 2008), where n represents the coordination stoichi­
ometry of DB14C4–C16 with Li+. On this basis, the extraction constant
(Kex) can be expressed in Eq. (6), and Eq. (7) can be further generated by
the combination of Eq. (6) and Eq. (3) (section 2.3.2). By taking loga­
rithms on both sides of Eq. (7), a linear relationship between log DLi and
log [DB14C4–C16]org can be finally obtained, as shown in Eq. (8), where
the complexation stoichiometry (n) and extraction constant (log Kex) are
corresponded to the slope and the intercept at y-axis, respectively.

Li+ aq + nDB14C4–C16org ⇌ nDB14C4–C16–Li+ org (5) Fig. 7. Logarithmic plot of DLi versus DB14C4–C16 concentration.

6
Y. Xiong et al. Journal of Environmental Management 310 (2022) 114705

3.2.2. Extraction in the presence of competing ions 1:1 coordination mechanism. Afterwards, the generated
In order to further determine the selectivity of DB14C4–C16 towards DB14C4–C16–Li+ complexes transfer from the reaction interface back­
Li+, solvent extraction experiments were carried out in the presence of wards to the organic bulk for the completion of lithium extraction. In
Na+, K+, Ca2+, and Mg2+, which are the main competing ions in brines. summary, the extraction of Li+ follows four essential steps successively:
The experiments were performed under the optimal conditions obtained (I) Transports of DB14C4–C16 and nH2O–Li+ within the organic and
in single Li+ extraction experiments (section 3.2.1), and the corre­ aqueous phases, respectively, towards the reaction interface. (II)
sponding results are shown in Fig. 8. It is clearly seen that the lithium Dehydration of nH2O–Li+ for free Li+ ions liberation. (III) Complexation
extraction process was barely interfered by the co-existing metal ions. between DB14C4–C16 and Li+ forming DB14C4–C16–Li+ complex at the
The extraction efficiencies of the cations were in an order of E+ 2+
Li > EMg >
interface. (IV) Transport of DB14C4–C16–Li+ complex from the inter­
E+ 2+ face towards the organic bulk.
Na > EK > ECa , which is in good agreement with the aforementioned
+
From the viewpoint of the competing ions, as the ionic diameters of
theoretical calculation results. DB14C4–C16 generally showed excellent
Na+, K+, and Ca2+ are much larger than the cavity size of DB14C4–C16,
selectivity toward Li+ over the impurities with the separation co­
they can hardly form stable complexes with the crown ether at the two-
efficients of Na+-Li+, K+-Li+, and Ca2+-Li+ being 16.32, 24.53, and
phase interface entering the organic phase during the liquid-liquid
68.09, respectively, indicating the limited complexation capacity of
extraction process. Although Mg2+ has a much better size compati­
DB14C4–C16 toward these competing ions. This can be attributed to the
bility with DB14C4–C16 compared with Na+, K+, and Ca2+, its
poor size compatibility of the crown ether cavity (1.2–1.5 Å) with Ca2+
complexation affinity to water molecules is the highest among the cat­
(1.98 Å), Na+ (2.04 Å), and K+ (2.76 Å) that the cations can hardly
ions, keeping it in the aqueous phase in the form of nH2O–Mg2+ clusters
coordinate with the donor oxygen atoms in the plane of the crown ring
rather than entering the organic phase through the complexation with
forming stable complexes during the lithium extraction process (Ped­
DB14C4–C16. Therefore, the selective extraction of Li+ in the presence
ersen, 1967; Shannon, 1976; Torrejos et al., 2015). Moreover, it is
of Na+, K+, Ca2+, and Mg2+ can be achieved by using DB14C4–C16 as
interesting to note that DB14C4–C16 shows also selectivity towards Li+
the extractant.
over Mg2+, which is the most competing cation for lithium extraction,
though the Mg2+-Li+ separation coefficient is only approximately 4,
4. Conclusions
much smaller than that for the other competing ions. As theoretically
estimated, DB14C4–C16–Mg2+ is more stable than DB14C4–C16–Li+,
This study has proposed a novel approach for selective Li+ extraction
nevertheless, the hydration effect of Mg2+ is much stronger compared
using dibenzo-14-crown-4 ether functionalized with an alkyl C16 chain
with Li+. Therefore, the uptake of Li+ by DB14C4–C16 is practically
(DB14C4–C16) as extractant, which was synthesized based on the ion
prior to that of Mg2+ as a result of the combined interactions with both
imprinting technology (IIT). The extraction capacity and selectivity of
crown ether and water during the solvent extraction process.
DB14C4–C16 towards Li+ over the competing ions such as Na+, K+, Ca2+
and Mg2+ were theoretically evaluated in terms of geometrical structure
3.2.3. Mechanism of Li+ extraction
optimization, molecular orbital analysis, as well as interaction and en­
The solvent extraction of Li+ by DB14C4–C16 is a typical liquid-
ergy analysis. The results reveal that Li+ can be readily separated from
liquid reaction, and the reaction mechanism can be described in
Ca2+, K+ as well as Na+ based on the size effect, and its separation from
Fig. 9. According to the aforementioned theoretical evaluations, lithium
Mg2+ can be achieved from the viewpoint of hydration effect. The
exists in the aqueous solution mainly in the form of nH2O–Li+ rather
liquid-liquid extraction experiments were performed in both single Li+
than free Li+ ions due to the strong hydration effect. For the liquid-liquid
and multi-ions containing solutions. The effects of various conditions
extraction system, since the organic and aqueous phases are two
such as reaction time, organic to aqueous phase ratio, temperature,
immiscible liquids, the interactions between the extractant of
crown ether concentration, and solution pH on the Li+ extraction pro­
DB14C4–C16 and the target ions of Li+ will occur at the two-phase
cess were systematically investigated. The optimal conditions for
interface. Therefore, in the initial stage of the extraction process,
lithium extraction are 30 min for reaction time, 2 for O/A ratio, 20 ◦ C for
DB14C4–C16 in the organic phase and nH2O–Li+ cluster ions in the
temperature, 1.5 for DB14C4–C16/Li+ ratio, and 7 for solution pH.
aqueous solution will simultaneously transfer towards the two-phase
Moreover, it was found that DB14C4–C16 forms 1:1 stoichiometry
interface to facilitate the complexation of the crown ether with
complex with Li+ at the reaction interface, and shows good selectivity
lithium. As DB14C4–C16 is not able to directly coordinate with the
towards Li+ over the competing ions. The separation coefficients of
nH2O–Li+ cluster ions due to the size effect, dehydration of nH2O–Li+
Ca2+-Li+, K+-Li+, Na+-Li+, and Mg2+-Li+ were obtained to be 68.09,
will take place to liberate free Li+ ions, which will further enter the
24.53, 16.32, and 3.99, respectively. The experimental results are
crown ether cavity to form DB14C4–C16–Li+ complexes following the

Fig. 8. Results of solvent extraction in the presence of competing ions.

7
Y. Xiong et al. Journal of Environmental Management 310 (2022) 114705

Fig. 9. Reaction mechanism of lithium extraction by DB14C4–C16.

basically in consistence with the theoretical analysis. and 1H-NMR spectroscopy. Evidence for mononuclear and binuclear complexes from
X-ray diffraction. Modeling of the metal complexation. New J. Chem. 20, 1183.
Barandiarán, J., 2019. Lithium and development imaginaries in Chile, Argentina and
Credit author statement Bolivia. World Dev. 113, 381–391.
Cao, J., Ren, Q., Chen, F., Lu, T., 2015. Comparative study on the methods for predicting
the reactive site of nucleophilic reaction. Sci. China Chem. 58, 1845–1852.
Yanhang Xiong: Methodology, Validation, Formal analysis, Inves­ Golebiowski, J.M., Lamare, V., Ruiz-López, M.F., 2010. Rb(+)/Cs(+) selectivity of benzo
tigation, Data curation, Writing – original draft, Writing – review & and tribenzo derivatives of the 21C7 crown ether. A density functional study.
editing. Tao Ge: Methodology, Validation, Formal analysis, Investiga­ J. Comput. Chem. 23, 724–731.
Gu, D., Sun, W., Han, G., Cui, Q., Wang, H., 2018. Lithium ion sieve synthesized via an
tion, Data curation. Liang Xu: Conceptualization, Methodology, Vali­
improved solid state method and adsorption performance for West Taijinar Salt Lake
dation, Formal analysis, Investigation, Data curation, Writing – original brine. Chem. Eng. J. 350, 474–483.
draft, Writing – review & editing, Funding acquisition. Ling Wang: Guo, Z.Y., Ji, Z.Y., Chen, Q.B., Liu, J., Zhao, Y.Y., Li, F., Liu, Z.Y., Yuan, J.S., 2018a.
Prefractionation of LiCl from concentrated seawater/salt lake brines by
Validation, Investigation, Data curation. Jindong He: Formal analysis,
electrodialysis with monovalent selective ion exchange membranes. J. Clean. Prod.
Investigation, Data curation. Xiaowei Zhou: Methodology, Investiga­ 193, 338–350.
tion. Yongpan Tian: Conceptualization, Validation, Data curation. Guo, X., Hu, S., Wang, C., Duan, H., Xiang, X., 2018b. Highly efficient separation of
Zhuo Zhao: Conceptualization, Methodology, Validation, Formal anal­ magnesium and lithium and high-valued utilization of magnesium from salt lake
brine by a reaction-coupled separation technology. Ind. Eng. Chem. Res. 57,
ysis, Investigation, Resources, Data curation, Writing – review & editing, 6618–6626.
Supervision, Project administration, Funding acquisition. He, L., Xu, W., Song, Y., Luo, Y., Liu, X., Zhao, Z., 2018. Lithium-ion batteries: new
insights into the application of lithium-ion battery materials: selective extraction of
lithium from brines via a rocking-chair lithium-ion battery system (global challenges
Declaration of competing interest 2/2018). Glob. Chall. 2, 1870020.
Ho, J., Klamt, A., Coote, M.L., 2010. Comment on the correct use of continuum solvent
models. J. Phys. Chem. 114, 13442–13444.
The authors declare that they have no known competing financial Kablov, E.N., Antipov, V.V., Oglodkova, J.S., Oglodkov, M.S., 2021. Development and
interests or personal relationships that could have appeared to influence application prospects of aluminum-lithium alloys in aircraft and space technology.
Metallurgist 65, 72–81.
the work reported in this paper. Kim, S., Lee, J., Kang, J.S., Jo, K., Kim, S., Sung, Y.E., Yoon, J., 2015. Lithium recovery
from brine using a λ-MnO2/activated carbon hybrid supercapacitor system.
Acknowledgements Chemosphere 125, 50–56.
Kim, S., Lee, J., Kim, S., Kim, S., Yoon, J., 2018. Electrochemical lithium recovery with
aLiMn2O4-zinc battery system using zinc as a negative electrode. Energy Technol. 6,
This work was supported by the National Natural Science Foundation 340–344.
of China (No. 51904003, U1703130, 51704011), Anhui Provincial Kumar, V., Sahu, S.K., Pandey, B.D., 2010. Prospects for solvent extraction processes in
the Indian context for the recovery of base metals. A review. Hydrometallurgy 103,
Natural Science Foundation (No. 2108085J26), and China Postdoctoral 45–53.
Science Foundation (No. 2019M651466). Li, Q., Liu, H., He, B., Shi, W., Ji, Y., Cui, Z., Yan, F., Mohammad, Y., Li, J., 2022.
Ultrahigh-efficient separation of Mg2+/Li+ using an in-situ reconstructed positively
charged nanofiltration membrane under an electric field. J. Membr. Sci. 641,
Appendix A. Supplementary data 119880.
Li, W., Shi, C., Zhou, A., He, X., Sun, Y., Zhang, J., 2017. A positively charged composite
Supplementary data to this article can be found online at https://doi. nanofiltration membrane modified by EDTA for LiCl/MgCl2 separation. Separ. Purif.
Technol. 186, 233–242.
org/10.1016/j.jenvman.2022.114705. Li, Z., Binnemans, K., 2020. Hydration counteracts the separation of lanthanides by
solvent extraction. AIChE J. 66 (9), e16545.
Liu, G., Zhao, Z., Ghahreman, A., 2019. Novel approaches for lithium extraction from
References
salt-lake brines: a review. Hydrometallurgy 187, 81–100.
Liu, X., Zhong, M., Chen, X., Li, J., Zhao, Z., 2020. Enriching lithium and separating
Achazi, A.J., von Krbek, L.K.S., Schalley, C.A., Paulus, B., 2016. Theoretical and lithium to magnesium from sulfate type salt lake brine. Hydrometallurgy 192,
experimental investigation of crown/ammonium complexes in solution. J. Comput. 105247.
Chem. 37, 18–24. Liu, X., Zhong, M., Chen, X., Zhao, Z., 2018. Separating lithium and magnesium in brine
Asfari, Z., Naumann, C., Vicens, J., Nierlich, M., Thuery, P., Bressot, C., Lamare, V., by aluminum-based materials. Hydrometallurgy 176, 73–77.
Dozol, J.F., 1996. 1,3-Calix(4)-bis-crown-4. Cesium complexation studies by 133Cs-

8
Y. Xiong et al. Journal of Environmental Management 310 (2022) 114705

Liu, H., Azimi, G., 2021. Process analysis and study of factors affecting the lithium Swain, B., 2017. Recovery and recycling of lithium: a review. Separ. Purif. Technol. 172,
carbonate crystallization from sulfate media during lithium extraction. 388–403.
Hydrometallurgy 199, 105532. Tian, Y., Chen, W., Zhao, Z., Xu, L., Tong, B., 2020. Interaction and selectivity of 14-
Liu, J., Zhang, Y., Miao, Y., Yang, Y., Li, P., 2021. Alkaline resins enhancing Li+/H+ ion crown-4 derivatives with Li+, Na+, and Mg2+ metal ions. J. Mol. Model. 26, 67.
exchange for lithium recovery from brines using granular titanium-type lithium ion- Torrejos, R.E.C., Nisola, G.M., Park, M.J., Shon, H.K., Seo, J.G., Koo, S., Chung, W.J.,
sieves. Ind. Eng. Chem. Res. 60, 16457–16468. 2015. Synthesis and characterization of multi-walled carbon nanotubes-supported
Lommelen, R., Onghena, B., Binnemans, K., 2020. Cation effect of chloride salting agents dibenzo-14-crown-4 ether with proton ionizable carboxyl sidearm as Li+ adsorbents.
on transition metal ion hydration and solvent extraction by the basic extractant Chem. Eng. J. 264, 89–98.
methyltrioctylammonium chloride. Inorg. Chem. 59 (18), 13442–13452. Torrejos, R.E.C., Nisola, G.M., Song, H.S., Han, J.W., Chung, W.J., 2016. Liquid-liquid
Lu, T., Chen, F., 2012. Multiwfn: a multifunctional wavefunction analyzer. J. Comput. extraction of lithium using lipophilic dibenzo-14-crown-4 ether carboxylic acid in
Chem. 33, 580–592. hydrophobic room temperature ionic liquid. Hydrometallurgy 164, 362–371.
Mason, P.E., Ansell, S., Neilson, G.W., Rempe, S.B., 2015. Neutron scattering studies of Turney, J.M., Simmonett, A.C., Parrish, R.M., Hohenstein, E.G., Evangelista, F.A.,
the hydration structure of Li+. J. Phys. Chem. B 119, 2003–2009. Fermann, J.T., Mintz, B.J., Burns, L.A., Wilke, J.J., Abrams, M.L., 2012. Psi4: an
Marenich, A.V., Cramer, C.J., Truhlar, D.G., 2009. Universal solvation model based on open-source ab initioelectronic structure program. Wiley Interdiscip. Rev. Comput.
solute electron density and on a continuum model of the solvent defined by the bulk Mol. Sci. 2 (4), 556–565.
dielectric constant and atomic surface tensions. J. Phys. Chem. B 113, 6378–6396. U.S Geological Survey, 2021. Lithium Statistics and Information.
Paranthaman, M.P., Li, L., Luo, J., Hoke, T., Ucar, H., Moyer, B.A., Harrison, S., 2017. U.S. Department of Energy, 2011. Critical Materials Strategy, Technical Report.
Recovery of lithium from geothermal brine with lithium-aluminum layered double Wang, S., Li, P., Zhang, X., Zheng, S., Zhang, Y., 2017. Selective adsorption of lithium
hydroxide chloride sorbents. Environ. Sci. Technol. 51 (22), 13481–13486. from high Mg-containing brines using HxTiO3 ion sieve. Hydrometallurgy 174,
Parker, T.M., Burns, L.A., Parrish, R.M., Ryno, A.G., Sherrill, C.D., 2014. Levels of 21–28.
symmetry adapted perturbation theory (SAPT). I. Efficiency and performance for Wu, X., Cui, S., Liu, S., Li, G., Gao, X., 2021. From dendrites to hemispheres: changing
interaction energies. J. Chem. Phys. 140, 094106. lithium deposition by highly ordered charge transfer channels. ACS Appl. Mater.
Parrish, R.M., Burns, L.A., Smith, D.G.A., Simmonett, A.C., Deprince, A.E., Hohenstein, E. Interfaces 13, 6249–6256.
G., Bozkaya, U., Sokolov, A.Y., Remigio, R.D., Richard, R.M., Gonthier, J.F., Xu, L., Xiong, Y., Zhang, G., Zhang, F., Yang, Y., Hua, Z., Tian, Y., Zhao, Z., 2020. An
James, A.M., McAlexander, H.R., Kumar, A., Saitow, M., Wang, X., Pritchard, B.P., environmental-friendly process for recovery of tellurium and copper from copper
Verma, P., Schaefer, H.F., Patkowski, K., King, R.A., Valeev, E.F., Evangelista, F.A., telluride. J. Clean. Prod. 272, 122723.
Turney, J.M., Crawford, T.D., Sherrill, C.D., 2017. Psi4 1.1: anopen-source electronic Xu, F., Dai, L., Wu, Y., Xu, Z., 2021. Li+/Mg2+ separation by membrane separation: the
structure program emphasizing automation, advanced libraries, and role of the compensatory effect. J. Membr. Sci. 636, 119542.
interoperability. J. Chem. Theor. Comput. 13 (7), 3185–3197. Yao, Y., Zhu, M., Zhao, Z., Tong, B., Fan, Y., Hua, Z., 2018. Hydrometallurgical processes
Pedersen, C.J., 1967. Cyclic polyethers and their complexes with metal salts. J. Am. for recycling spent lithium-ion batteries: a critical review. ACS Sustain. Chem. Eng. 6
Chem. Soc. 89, 2495–2496. (11), 13611–13627.
Pereira, F., Xiao, K., Latino, D.A.R.S., Wu, C., Zhang, Q., Aires-De-Sousa, J., 2016. Yousefi, S.R., Amiri, O., Salavati-Niasari, M., 2019a. Control sonochemical parameter to
Machine learning methods to predict density functional theory B3LYP energies of prepare pure Zn0.35Fe2.65O4 nanostructures and study their photocatalytic activity.
HOMO and LUMO orbitals. J. Chem. Inf. Model. 57, 11–21. Ultrason. Sonochem. 58, 104619.
Rodriguez, J.D., Lisy, J.M., 2011. Probing lonophore selectivity in argon-tagged hydrated Yousefi, S.R., Masjedi-Arani, M., Morassaei, M.S., Salavati-Niasari, M., Moayedi, H.,
alkali metal ion-crown ether systems. J. Am. Chem. Soc. 133, 11136–11146. 2019b. Hydrothermal synthesis of DyMn2O5/Ba3Mn2O8 nanocomposite as a
Roper, E.D., Talanov, V.S., Butcher, R.J., Talanova, G.G., 2008. Selective recognition of potential hydrogen storage material. Int. J. Hydrogen Energy 44 (43), 24005–24016.
thallium(I) by 1,3-alternate calix[4]arene-bis(crown-6 ether): a new talent of the Yousefi, S.R., Ghanbari, M., Amiri, O., Marzhoseyni, Z., Mehdizadeh, P., Hajizadeh-
known ionophore. Supramol. Chem. 20 (1–2), 217–229. Oghaz, M., Salavati-Niasari, M., 2021a. Dy2BaCuO5/Ba4DyCu3O9.09 S-scheme
Saito, T., Sato, K., Miratsu, M., Matsubayashi, I., Inoue, M., Hasegawa, Y., 2010. Several heterojunction nanocomposite with enhanced photocatalytic and antibacterial
factors influencing the stability of extracted species in synergistic extraction of activities. J. Am. Ceram. Soc. 104, 2952–2965.
lanthanide(III) with pivaloyltrifluoroacetone and 2,2-bipyridyl. Anal. Sci. 26 (11), Yousefi, S.R., Alshamsi, H.A., Amiri, O., Salavati-Niasari, M., 2021b. Synthesis,
1187–1191. characterization and application of Co/Co3O4 nanocomposites as an effective
Sener, S.E.C., Thomas, V.M., Hogan, D.E., Maier, R.M., Carbajales-Dale, M., Barton, M.D., photocatalyst for discoloration of organic dye contaminants in wastewater and
Karanfil, T., Crittenden, J.C., Amy, G.L., 2021. Recovery of critical metals from antibacterial properties. J. Mol. Liq. 337, 116405.
aqueous sources. ACS Sustain. Chem. Eng. 9, 11616–11634. Yousefi, S.R., Sobhani, A., Alshamsi, H.A., Salavati-Niasari, M., 2021c. Green
Shannon, R.D., 1976. Revised effective ionic radii and systematic studies of interatomic sonochemical synthesis of BaDy2NiO5/Dy2O3 and BaDy2NiO5/NiO nanocomposites
distances in halides and chalcogenides. Acta Crystallogr. A 32 (5), 751–767. in the presence of core almond as a capping agent and their application as
Shi, C., Yan, J., Jiang, X., Wang, X., Jia, Y., 2016. Solvent extraction of lithium from photocatalysts for the removal of organic dyes in water. RSC Adv. 11, 11500.
aqueous solution using non-fluorinated functionalized ionic liquids as extraction Zhang, H.Z., Xu, Z.L., Ding, H., Tang, Y.J., 2017. Positively charged capillary
agents. Separ. Purif. Technol. 172, 473–479. nanofiltration membrane with high rejection for Mg2+ and Ca2+ and good separation
Shi, D., Cui, B., Li, L., Peng, X., Zhang, L., Zhang, Y., 2018. Lithium extraction from low- for Mg2+ and Li+. Desalination 420, 158–166.
grade salt lake brine with ultrahigh Mg/Li ratio using TBP-kerosene-FeCl3 system. Zhao, Q., Gao, J., Guo, Y., Cheng, F., 2018. Facile synthesis of magnetically recyclable
Separ. Purif. Technol. 211, 303–309. Fe-doped lithium ion sieve and its Li adsorption performance. Chem. Lett. 47,
Song, J.F., Nghiem, L.D., Li, X.M., He, T., 2017. Lithium extraction from Chinese salt-lake 1308–1310.
brines: opportunities, challenges, and future outlook. Environ. Sci-Wat. Res. 3 (4), Zhao, X., Yang, H., Wang, Y., Sha, Z., 2019. Review on the electrochemical extraction of
593–597. lithium from seawater/brine. J. Electroanal. Chem. 850, 113389.
Sun, Y., Zhu, M., Yao, Y., Wang, H., Tong, B., Zhao, Z., 2020. A novel approach for the Zhao, Z., Xiong, Y., Cheng, X., Hou, X., Yang, Y., Tian, Y., You, J., Xu, L., 2020.
selective extraction of Li+ from the leaching solution of spent lithium-ion batteries Adsorptive removal of trace thallium(I) from wastewater: a review and new
using benzo-15-crown-5 ether as extractant. Separ. Purif. Technol. 237, 116325. perspectives. J. Hazard Mater. 393, 122378.
Swain, B., 2016. Separation and purification of lithium by solvent extraction and Zhao, Y., Wu, Y., Liu, H., Chen, S., Bo, S., 2021. Accelerated growth of electrically
supported liquid membrane, analysis of their mechanism: a review. J. Chem. isolated lithium metal during battery cycling. ACS Appl. Mater. Interfaces 13,
Technol. Biotechnol. 91, 2549–2562. 35750–35758.

You might also like