You are on page 1of 10

Chemical Engineering Journal 410 (2021) 128320

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Titanium-based ion sieve with enhanced post-separation ability for high


performance lithium recovery from geothermal water
Shangqing Chen, Zishen Chen, Zhenwei Wei, Jiayin Hu *, Yafei Guo, Tianlong Deng *
Tianjin Key Laboratory of Brine Chemical Engineering and Resource Eco-utilization, College of Chemical Engineering and Materials Science, Tianjin University of Science
and Technology, Tianjin 300457, PR China

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

Keywords: The improvement of post-separation performance of powdery lithium ion sieve with the guarantee of high
Lithium recovery adsorption capacity and fast kinetics remains to be a huge challenge. Herein, a novel granular and porous
Porous ion sieve titanium-based lithium ion sieve (PIS) was developed using agar-assisted strategy and used for lithium recovery
Agar
from geothermal water. The agar acted as both spherality-shaping and sacrificial porogenic agent. As a result, PIS
Geothermal water
had a 64-fold larger diameter (2.8 mm) than powdery ion sieve (IS), which could be easily separated by filtration
and then reused with a steady performance. Due to its loose microstructures and rich porosity, PIS showed high
adsorption capacity (25.8 mg/g) and rapid kinetic (equilibrium time 6 h) in geothermal water. In addition, the
separation factors of competitive ions related to Li+ were 1162.3, 273.7 and 328.5 for Na+, K+ and Ca2+,
respectively, exhibiting benign selectivity in natural geothermal water. Thus, this granular PIS with enhanced
post-separation ability, high capacity and fast kinetic could be considered as a promising candidate for lithium
recovery from geothermal water. Moreover, this simple and green method could be easily popularized to prepare
other porous adsorbents.

1. Introduction adsorption [24–26]. Among which, precipitation and solvent extraction


are usually used for lithium recovery from brines with high Li+ con­
Lithium is rewarded as the fascinating energy material advancing the centration and low Mg/Li ratio. Membrane and electrochemistry process
world and facilitating the development of national economy [1–8]. It is could selectively separate divalent Mg2+ and monovalent Li+, but they
expected that the market demand for lithium will continue to grow in usually suffer from poor recovery efficiency and high energy input [27].
the further, which provides huge challenge for the issue of long-term Thus, regarding the simple process and high selectivity, adsorption has
production of lithium. With the increasing depletion of high-grade been accepted as an ideal choice for low-grade lithium recovery.
lithium ores, lithium recovery from water resources such as seawater Lithium ion sieve (LIS) is recognized as a series of selective adsor­
and salt lake brines has become an inevitable trend [9,10]. Geothermal bents with high ion-sieving functionality and lithium uptake capacity,
water, as a new-type liquid lithium storehouse, has emerged as a which is capable of separating lithium from complex water system
promising candidate for lithium recovery owing to its low Mg/Li ratio [28–30]. Based on the chemical elements, LIS could be classified into
and total dissolved solids [10]. However, as far as we know, the re­ manganese-based lithium ion sieve (Mn-LIS) and lithium titanium-based
searches focused on lithium adsorption from geothermal water is lithium ion sieve (Ti-LIS) [31]. Given the serious Mn loss of Mn-LIS
extremely insufficient for the Li+ concentrations in geothermal water are during acid-eluting process, Ti-LIS is more chemically stable. While
usually low with the range of 20 ~ 50 mg/L [11–13]. Therefore, it is the ultrafine morphology of Ti-LIS produces poor liquidity, insufficient
highly desirable but challenging to develop economically viable and permeability and low recycling efficiency in industrial applications,
environmental friendly techniques for lithium recovery from such low leading to serious problems in post-separation. To address this issue, an
Li+ concentration aqueous solutions. effective strategy is to immobilize Ti-LIS powder onto binding materials,
There are several methods for lithium recovery from aqueous solu­ such as chitosan [32,33], polyvinyl alcohol (PVA) [34], poly­
tion, including precipitation [14], solvent extraction [15,16], mem­ acrylonitrile (PAN) [35,36] and polyvinyl chloride (PVC) [37] etc.
brane process [17–20], electrochemistry [21–23] and ion-exchange Limjuco et al. prepared a H2TiO3 (HTO)/PVA composite foam and its

* Corresponding authors.
E-mail addresses: hujiayin@tust.edu.cn (J. Hu), tldeng@tust.edu.cn (T. Deng).

https://doi.org/10.1016/j.cej.2020.128320
Received 21 October 2020; Received in revised form 3 December 2020; Accepted 25 December 2020
Available online 2 January 2021
1385-8947/© 2021 Elsevier B.V. All rights reserved.
S. Chen et al. Chemical Engineering Journal 410 (2021) 128320

uptake of lithium was 12.5 mg/g at 12 h [38]. In our previous work, a in absolute ethyl alcohol through vigorous stirring and ultrasonic pro­
porous PVC/HTO composite was synthesized for lithium adsorption cess [38]. Then, it was dried at 323 K at an oven. The obtained mixture
from geothermal water and obtained a capacity of 11.35 mg/g at 328.15 was loaded at a ceramic boat and roasted at 1123 K at a heating rate of 5
K and 12 h [11]. Although the stability and recycle performance of Ti- K/min at N2 atmosphere. Subsequently, it was naturally cooled to obtain
LIS are improved after granulation, there are still two major problems Li2TiO3 precursor (LTO). The Li2TiO3 precursor was placed in 0.25 mol/
(i) the decreased adsorption capacity after the addition of binding ma­ L HCl (m/V = 1 g/L) at 333 K for 12 h. After stirring, the slurry was
terials and (ii) retarded adsorption kinetics caused by the compact centrifuged, washed and dried at 333 K to give H2TiO3 (HTO).
structure and large mass transfer resistance. For example, the powder
HTO could reach adsorption equilibrium at 2 h with an uptake capacity 2.2.2. Preparation of IS and PIS
of 30 mg/g, but it prolonged to 24 h after granulation and the adsorption The preparation process of IS and PIS are shown in Fig. 1. Firstly,
capacity also suffered a sharp decrease [11]. From this perspective, HTO powder and agar with different mass ratios (1:1, 2:1, 3:1, 4:1 and
novel strategy for the preparation of high performance Ti-LIS is of great 5:1) were loaded at 15 mL deionized water, then it was heated to 373 K
significance. with simultaneously stirring until the agar was completely molten. The
Agar, a polysaccharide composed of agarose and agaropectin [39], slurry was immediately dropped into cold simethicone via an injector to
shows good potential to be spherality-shaping for the synthesis of obtained white spheres. After stood for 24 h, the spheres was cleaned by
granular materials. Herein, we present a novel agar-assisted strategy to petroleum ether for several times and further dried to obtain a series of
prepare a granular Ti-LIS with enhanced post-separation ability and ion sieves (IS, denoted as IS-1 ~ 5). Moreover, the corresponding PIS
used for lithium recovery from geothermal water. The major objective were acquired after the calcination of IS spheres at 823 K for 1 h in air.
was to prepare a granular adsorbent to improve the post-separation
ability of powdery Ti-LIS with the guarantee of high capacity and fast 2.3. Characterization
kinetic. Therefore, the current study is aimed to (i) prepare granular
agar-HTO composite ion sieve (IS) via the glomeration functionality of XRD patterns were collected by powder X-ray diffraction (XRD,
agar; (ii) synthesis of porous ion sieve sphere (PIS) comprised of MSAL XD-3) using Cu Kα radiation over a 2θ range of 5-75 and X-ray

chemical stability and uniform mesoporous structure; (iii) analyse the power of 36 kV/20 mA at a scan. The thermogravimetric analysis (TGA)
physiochemical properties and structural configuration of PIS by SEM, was carried out in a TG-DSC instrument (Seteram Labsys, France). The
TEM, BET, XPS, XRD, TG-DSC; (iv) evaluate the adsorption performance surface morphologies of IS and PIS was demonstrated by scanning
of lithium and optimize adsorption parameters using batch experiments; electron microscopy (SEM, SU1510, HITCH, Japan). TEM was recorded
(v) investigate the potential application of PIS in lithium recovery in real by FEI TalosF200x. Pore structure data were collected and calculated by
geothermal water. As a result, this novel granular and porous lithium ion N2 adsorption–desorption apparatus (Autosorb iQ). X-ray photoelectron
sieve prepared using agar as dual spherality-shaping and sacrificial spectroscopy (XPS) was carried out using a K-Alpha (Thermo Scientific)
porogenic agent exhibited good post-separation ability and stability, equipped with a microfocused monochromator X-ray source.
high adsorption capacity, fast kinetics and benign selectivity in lithium
recovery from geothermal water. This porous ion sieve sphere is fresh,
facilely prepared and effective. This simple and green preparation 2.4. Lithium adsorption
method could be easily popularized in the fabrication of granular porous
adsorbents. It’s noteworthy that the temperature of geothermal water is rela­
tively high [34]. In order to simulate the adsorption process in
geothermal water, the adsorption temperature is controlled at 333 K
2. Experiment section
unless otherwise stated in this work.
2.1. Materials
2.4.1. Adsorption isotherm
Adsorption isotherm was determined by shaking a desired amount of
Lithium carbonate (Li2CO3, AR), lithium chloride (LiCl, AR), hy­
adsorbents and lithium solution ranging from 0 to 500 mg/L in a poly­
drochloric acid (HCl, AR) and sodium hydroxide (NaOH, AR) were ob­
tetrafluoroethylene bottle at pH 12 and 333 K. After shaking for 10 h, the
tained from Sinopham Chemical Reagent Co., Ltd. Agar (BR, high gel
supernatant was obtained by filtration and the concentration of residual
strength 1000–1200 g/cm2), titanium dioxide (TiO2, AR, anatase),
Li+ was analyzed by ICP-OES, and the adsorption capacities at different
simethicone (AR), and petroleum ether (AR) were purchased from
initial Li+ concentrations were calculated.
Macklin Biochemical Co., Ltd.

2.4.2. Adsorption kinetic


2.2. Preparation of lithium ion sieves The time profile as well as adsorption kinetic during adsorption
process was obtained by determining the residual Li+ concentrations at
2.2.1. Preparation of ion sieves precursor different adsorption time with a shaking speed of 200 rpm. Then, the
Li2CO3 and TiO2 with a molar ratio of Li/Ti = 2:1 were well dispersed adsorption capacities at different contact time were calculated.

Fig. 1. Schematic diagram of the synthesis of IS and PIS.

2
S. Chen et al. Chemical Engineering Journal 410 (2021) 128320

2.5. Application in lithium recovery from geothermal water 3. Results and discussion

Geothermal water sampled from a bore well in Tibet autonomous 3.1. Characterization
region, China was used in this work. After adsorption, the adsorbents
were eluted by HCl solution at 333 K for 12 h and then subject to another The XRD patterns of LTO, HTO, IS and PIS are shown in Fig. 2a. The
adsorption process. Concerning the dissolution loss of adsorbents, the diffraction peaks at 18.46◦ , 20.45◦ , 35.8◦ , 43.5◦ , 47.4◦ , 57.5◦ , 63.5◦ and
concentrations of Ti element in eluents were determined. 66.7◦ attributed to the (0 0 2), (1 1 0), (− 1 3 1), (− 1 3 3), (− 2 0 4), (0 0 6),
(− 2 0 6) and (0 6 2) lattice planes in LTO precursor [40–42]. The pattern
2.6. Data analysis agreed well with the standard card (Li2TiO3, JCPDS-033–0831), indi­
cating the successful formation of LTO. When LTO was delithiation
The adsorption and desorption properties were evaluated by through acid treatment, the crystal planes of (− 1 3 3), (0 0 6), (− 2 0 6)
adsorption efficiency (E, %), adsorption capacity (qe, mg/g), distribution and (0 6 2) disappeared due to the replacement of H between the Ti-O
coefficient (Kd, mL/g), separation factor (SF), desorption efficiency (DE, layers. After granulated by agar method, a new peak was found in IS
%) and dissolution loss rate (DL, %). The calculated equations were at about 19◦ , which was attributed to the characteristic peak of agar. It
listed as below. was noteworthy that the pure agar was of low degree of crystallinity,
thus the agar peak was not obvious. While in the XRD pattern of PIS, the
C0 − Ce
E(\%) = ) = × 100 (1) agar peak disappeared and the pattern was similar with that of HTO,
C0
indicating the decomposition of agar in thermal process.
(C0 − Ce )V1 To further investigate the thermal decomposition process, the ther­
qe (mg/g) = (2) mogravimetric analyses of agar, HTO and IS were carried out and the
m1
results are shown in Fig. 2b. It was found that the weight of HTO did not
(C0 − Ce) 1000V1 changed notably in calculation, indicating its stable structure. The
Kd (mL/g) = × (3) weight loss of agar went through two steps: from room temperature to
Ce m1
473 K, the loss was about 10.5%, which may ascribe to the release of
Kd (Li) H2O and impurities; from 473 K to 823 K, the weight loss increased
SF = (4)
Kd (M) sharply to 76.8% due to the decomposition of major agar in the form of
CO2 and H2O [43]. As for IS, the weight loss was only about 23.5%
DE(\%) = ) =
C1 × V2
× 100 (5) because of the low mass fraction of agar in IS. In this case, we concluded
m2 that the thermal formation of PIS was accompanied by agar
decomposition.
C 2 × V2
DL(\%) = ) = × 100 (6) The release of CO2 and H2O vapor in the calcination process would
m3 generate abundant pore channels, which was further confirmed by SEM
and TEM images. Fig. 3 showed the digital photos and SEM images of IS
where Co and Ce (mg/L) are the concentrations of Li+ in the solutions at
and PIS spheres. The granular PIS spheres had high degree of sphericility
initial and equilibrium. C1 and C1 (mg/L) refer to the concentrations of
with an average diameter of 2.8 mm, which was 64-fold larger than that
Li+ and Ti4+ in the eluent after stripping. V1 and V2 represents the
of powdery Ti-LIS (43.7 μm) [11]. Thus, PIS could be easily separated by
volumes of lithium solution and eluent. m1 and m2 are the mass of
simple filtration. Comparing to the micromorphologies of original IS, the
adsorbent and lithium-laden adsorbent, and m3 is the mass of Ti element
thermal-treated PIS sphere was corrugated, which could provide more
in the lithium-laden adsorbent, respectively. M represents different
contact areas for Li+, facilitating the adsorption process. Moreover, PIS
metallic ions co-existed with lithium in geothermal water.
spheres had loose microstructures and rich porosity, which was
demonstrated by TEM results (Fig. S1) and N2 adsorption–desorption
isotherms (Fig. 4). The isotherms of PIS exhibited the typical type IV

Fig. 2. Powder X-ray diffraction patterns (a) and thermogravimetric analysis (b).

3
S. Chen et al. Chemical Engineering Journal 410 (2021) 128320

Fig. 3. Digital photos (a) and SEM images (b) of IS and PIS spheres.

Fig. 4. The N2 adsorption-desorption isotherm (a) and pore size distribution curve (b) of IS and PIS.

isotherms with hysteresis loops. The specific surface areas of IS and PIS structure. The time profiles vs. adsorption efficiency of IS-4 and PIS-4
were 6.94 and 102.78 m2/g, which convincingly confirmed the gener­ were further determined. It was found that the adsorption kinetic of
ation of porous structure in the thermal process. The pore size distri­ PIS was faster than that of IS, resulting in a higher adsorption efficiency
bution curve (Fig. 4b) showed the pore diameters of PIS were mainly in the same adsorption time (Fig. 5b). This results were resulted from the
distributed at 3–5 nm, demonstrating its uniform mesoporous structure. porous structure of PIS, which greatly facilitate Li+ to diffusion into the
inner adsorbent to be captured. Taking into account of adsorption effi­
3.2. Adsorption performance evaluation ciency and adsorption kinetics, PIS-4 was used for further exploration in
this work.
3.2.1. Effect of HTO concentration
The HTO concentration greatly influenced the lithium adsorption 3.2.2. Effect of adsorption conditions
performance of IS and PIS. In this case, IS and PIS with different mass The effects of adsorption conditions, such as pH value, temperature,
ratios of HTO:agar (denoted as IS-1 ~ 5 and PIS-1 ~ 5) were prepared Li + concentration and time on lithium adsorption performance of PIS-4
and the lithium adsorption performance were evaluated in batch ex­ were investigated, and the results are shown in Fig. 6. As shown in
periments. As Fig. 5a showed, the lithium adsorption efficiency Fig. 6a, it can be seen that the adsorption efficiencies and distribution
increased with the increasing concentration of HTO and reached a coefficients were relative low in neutral and alkalescent conditions.
plateau at HTO:agar of 4:1 in both IS and PIS. Obviously, the adsorption While the alkalinity became stronger, the Li+ adsorption efficiency and
performance of PIS was better than that of IS because of its mesoporous distribution coefficient increased sharply. The optimal adsorption

4
S. Chen et al. Chemical Engineering Journal 410 (2021) 128320

Fig. 5. (a) Lithium adsorption efficiency of IS and PIS with different mass ratios of HTO: agar, conditions: C0 = 25 mg/L, m/V = 2 g/L, pH 12, T = 333 K, t = 8 h and
(b) lithium adsorption kinetics of IS-4 and PIS-4, conditions: C0 = 25 mg/L, m/V = 2 g/L, pH 12, T = 333 K, t = 10 h.

Fig. 6. Effect of (a) pH values, (b) temperature, (c) Li+ concentration and (d) time on adsorption performance for lithium. conditions: (a) C0 = 25 mg/L, m/V = 2 g/L,
pH 6–13, T = 333 K, t = 8 h; (b) C0 = 25 mg/L, m/V = 2 g/L, pH 12, T = 293–353 K, t = 8 h; (c) C0 = 25–500 mg/L, m/V = 2 g/L, pH 12, T = 333 K, t = 8 h; (d) C0 =
25 mg/L, m/V = 2 g/L, pH 12, T = 333 K, t = 10 h.

efficiency of 98.4% and distribution coefficient of 30750 mL/g were this case, the adsorption process was realized by ion exchange between
obtained at pH 12. However, once pH value exceeded 12 and the ion H+ and Li+ in the solution. The generated H+ would rapidly bond to
strength was excessive, the adsorption capacity gradually decreased. In OH− , thereby promoting the ion exchange process and then enhancing

5
S. Chen et al. Chemical Engineering Journal 410 (2021) 128320

the Li+ adsorption performance [44]. Table 1


The increasing temperature was conducive to the Li+ adsorption Adsorption isotherm and kinetics parameters of lithium adsorption process by
(Fig. 6b), indicating the adsorption by PIS-4 was an endothermic pro­ PIS-4.
cess. With the increasing Li+ concentration in solution, the adsorption Adsorption isotherms
efficiency decreased markedly from 98.4% to 14.6% (Fig. 6c), but the Langmuir model Freundlich model
adsorption capacity increased from 12.3 mg/g to 36.6 mg/g, showing
PIS-4 Qmax (mg/g) b (L/mg) R2 KF (mg/g) n R2
superior lithium adsorption ability of PIS-4. When it comes to time
profile, the adsorption proceeded rapidly within the first 4 h and almost 34.23 0.119 0.993 41.43 6.75 0.954
unchanged in further contact time (Fig. 6d). The results demonstrated Adsorption kinetics
the high capacity and fast kinetic of PIS-4 in lithium recovery from pseudo-first-model pseudo-second-model
aqueous solutions. − 1 2
qe, cal k1 (h ) R qe, cal k2 (mg⋅g− 1⋅h) R2

3.2.3. Adsorption isotherms 12.78 0.556 0.942 12.64 0.622 0.958


Langmuir and Freundlich isotherm equations were used to evaluate
the experimental data in this work, and their expressions were listed in
Eqs. (7) and (8) [45,46]. ln(qe − qt ) = lnqe − k1 t (9)
Ce 1 Ce
= + (7) t 1 t
qe bqm qm = ( 2) + (10)
qt k2 qe qe

lnqe = lnKF +
1
lnCe (8) In which qe and qt (mg/g) are the removal capacity of cesium ions at
n equilibrium and at time t, k1 (h− 1) and k2 (mg⋅g− 1⋅h) denote the pseudo-
first-order and pseudo-second-order constants, respectively.
where qe and qm (mg/g) are the equilibrium and maximum adsorption
The higher R2 value of pseudo-second-order (0.958, Table 1) sug­
capacity, Ce (mg/L) is the equilibrium concentration. b is the Langmuir
gested the ion-exchange chemisorption process. In addition, the rate
coefficient, KF and 1/n refer to the constants related to the adsorption
constant k2 was larger than that of some reported lithium adsorbents,
capacity and the adsorption intensity.
such as HTO/PVA [38] and PVC-HTO [11], further indicating a fast
The experimental data and isotherm fitting curves are shown in
adsorption process of PIS-4, which is important for practical applica­
Fig. 7a, and the adsorption isotherm parameters are listed in Table 1. It
tions. This fast adsorption kinetics were benefited from the abundant
can be seen that the correlation coefficient R2 of Langmuir isotherm
pore channels and large specific surface areas generated during calci­
equation (0.993) was higher than that of Freundlich isotherm equation
nation process (Fig. 4a and Fig. S1), which could provide more contact
(0.954), indicating the adsorption process was more likely to be
areas for Li+, greatly facilitating Li+ to diffusion into the inner adsorbent
described by Langmuir isotherm. Based on the Langmuir equation, the
to be captured.
maximum Li+ adsorption capacity was calculated to be 34.23 mg/g. In
the thermal process, the decomposition of agar led to a higher HTO
3.2.5. Adsorption thermodynamics
concentration in PIS-4. Benefiting from its fluffy and porous structures,
Thermodynamic parameters could be obtained from the linear fitting
the adsorption capacity of PIS-4 was outstanding among other reported
of lnKd as a function of 1000/T (Fig. 8), including enthalpy (ΔH0,
adsorbents [11].
kJ⋅mol− 1), entropy (ΔS0, J⋅mol− 1⋅K− 1) and Gibbs free energy (ΔG0,
kJ⋅mol− 1). The values can be calculated using Eqs. (11) and (12) [49].
3.2.4. Adsorption kinetics
The adsorption capacity on PIS-4 as a function of adsorption time is ΔS0 ΔH 0 1
lnKd = − ⋅ (11)
presented in Fig. 7b. Moreover, the adsorption kinetic data were fitted R R T
using pseudo-first-order model (Eq. (9)) and pseudo-second-order model
(Eq. (10)) [47,48]. ΔG0 = ΔH 0 − TΔS0 (12)

Fig. 7. (a) Adsorption isotherm curves and (b) kinetic curves of PIS-4 for lithium adsorption.

6
S. Chen et al. Chemical Engineering Journal 410 (2021) 128320

Table 3
Lithium selective recovery from geothermal water using PIS-4.a.
Ions Ionic radius (pm) C0 (mg/L) E(%) Kd (mL/g) SF

Li
+
76 25.8 95.3 10241.7 1.0
Na+ 102 682.1 5.9 8.8 1162.3
K+ 138 137.9 4.8 37.4 273.7
Ca2+ 100 211.1 6.3 32.1 328.5
a
m/V = 2 g/L, pH 12, T = 333 K, t = 6 h.

hindering its imbedding into PIS-4 [27]. The overall results indicated the
PIS-4 could be readily used for highly selective recovery of Li+ from
geothermal water and even Na, K, Ca, Mg-rich sea water.

3.3.2. Desorption and regeneration


Followed by adsorption, the lithium-laden PIS-4 (PIS(Li)-4) was
collected by simple filtration and regenerated by HCl solution. In order
to obtain the optimal HCl concentration, the effect of HCl concentration
on desorption performance was firstly investigated. As Fig. 9a shows, the
Fig. 8. Thermodynamics fitting of lithium adsorption by PIS-4. desorption efficiency of Li+ gradually increased with increasing HCl
concentration and when HCl concentration was larger than 0.25 mol/L,
negligible increase was observed. However, the dissolution loss of Ti
where Kd is the equilibrium distribution coefficient, T(K) is the tem­ increased sharply in high HCl concentration. Even so, compared with the
perature, R refers to the universal gas constant with the value of 8.314 original LTO (4.2% in 0.5 mol/L HCl), the dissolution loss of Ti of PIS
J⋅mol− 1⋅K− 1, respectively. (Li)-4 (1.2%) decreased significantly, which indicated its improved
The positive value of ΔH0 and ΔS0 suggested an endothermic and stability. Taking Li+ desorption efficiency and dissolution loss into ac­
randomness process, and the negative value of ΔG0 indicate the count, 0.25 mol/L HCl solution was selected for desorption process. In
adsorption was spontaneous (Table 2). In this case, higher temperature this case, the Li+ desorption efficiency could reach 95.8% and the
is conducive for lithium adsorption by PIS-4, which is favorable to be overall recovery rate was 91.3% in a single adsorption–desorption cycle.
used in lithium recovery from geothermal water. Moreover, the morphologies of PIS and regenerated PIS-4 (denoted as
RPIS-4) were also determined in Fig. S2. It was found that the macro­
3.3. Application in lithium recovery from geothermal water scopic and microscopic morphologies of PIS-4 before and after regen­
erated (RPIS-4) were similar, indicating its excellent physical and
3.3.1. Lithium adsorption chemical stability. Whereafter, the reusability of PIS-4 was further
From above findings, PIS-4 was highly efficient in lithium recovery evaluated through five times recycle experiments. It can be seen from
from Li+ solutions. After that, PIS-4 was applied in geothermal water Fig. 9b that the adsorption efficiency and capacity decreased slightly,
containing competitive ions (Na+, K+ and Ca2+) for selective lithium but did not change notably relative to the pristine one. In addition, a
adsorption and the results are presented in Table 3. The initial concen­ twelve-times recycle experiment was also carried out in Fig. S3 and the
trations of Li+, Na+, K+ and Ca2+ in geothermal water were expressed as results showed the desorption efficiency and capacity did not change
C0, which were 25.8, 682.1, 137.9 and 211.1 mg/L, respectively. The pH notably after twelve-times recycle, indicating the long-term application
value of original geothermal water was 8.80 [11]. In order to carry out performance of PIS-4 for Li+ recovery from geothermal water. These
the Li+ absorption process at optimal pH value [40], the pH value was results demonstrated the long-term application potential for PIS-4 for
adjusted to 12 using NaOH solution. Although the concentrations of Li+ recovery from geothermal water.
competitive ions were much higher than Li+, the adsorption efficiency
and distribution coefficient of Li+ (95.3% and 10241.7) were extremely
3.4. Adsorption mechanism
higher than others (5.9% and 8.8 for Na+, 4.8% and 37.4 for K+, 6.3%
and 32.1 for Ca2+). The separation factors of competitive ions related to
Adsorption mechanism is a significant part in the design and appli­
Li+ were 1162.3, 273.7 and 328.5 for Na+, K+ and Ca2+, respectively,
cation of adsorbents. Thus, the XPS analyses of PIS-4 and PIS(Li)-4 were
and the results should be attributed to the ion sieving effect of PIS-4. In
carried out to give insight into the adsorption process. As shown in
order to further investigate the competitive effect of Mg2+ on Li+
Fig. 10a, the characteristic peaks of Ti 2p and O 1 s were identified in the
adsorption, binary solutions containing Li+ and Mg2+ ions were used.
survey XPS spectra, which appeared at 459.2 eV and 532.1 eV in PIS-4
The initial Li+ concentration was fixed at 25 mg/L, and that for Mg2+
and 458.8 eV and 532.2 eV in PIS(Li)-4, respectively. The 0.4 eV
was set as 50, 100 and 200 mg/L, which were much larger than Li+. The
displacement of Ti 2p could be attributed to the transformation of H-O-
adsorption efficiency and separation factors were shown in Table S1,
Ti to Li-O-Ti after Li+ adsorption. In this case, the chemical environment
and the results indicated the increasing Mg2+ concentration showed
of O atom changed, so that the binding energy of O 1 s increased for 0.1
negligible effect on the adsorption efficiency of Li+, and the adsorption
eV. These results suggested that the less electronegative Li+ was
efficiency of Mg2+ was much smaller than that of Li+. The separation
embedded into the ion cages [27], which was further verified by the
factors between Li+ and Mg2+ was in the range of 557.5 to 807.9, which
high resolution of Li 1 s of PIS(Li)-4 (Fig. 10b, inset). Therefore, the
exhibited benign separation performance. Although the size of Mg2+ is
adsorption mechanism of Li+ adsorption by PIS-4 was identified to be
similar with Li+, the hydration energy is much larger than that of Li+,
ion exchange, being in good agreement with the results of pseudo-
second-order model and previous reports [11].
Table 2
Thermodynamic parameters for lithium adsorption onto PIS-4.
3.5. Comparison with other adsorbents
ΔH0 (kJ⋅mol− 1) ΔS0 (J⋅mol− 1⋅K− 1) ΔG0 (kJ⋅mol− 1)

298 K 308 K 318 K 328 K In order to evaluate the potential application of PIS-4 in lithium re­
39.2 202.7 − 20.1 − 24.2 − 28.3 − 32.4
covery, a comparison of PIS-4 with other reported LIS based composite

7
S. Chen et al. Chemical Engineering Journal 410 (2021) 128320

Fig. 9. (a) Effect of HCl concentration on desorption performance and (b) reusability of PIS-4. conditions: (a) 0.1–0.5 mol/L HCl solutions, m/V = 1 g/L, T = 333 K, t
= 12 h; (b) geothermal water, m/V = 2 g/L, pH 12, T = 333 K, t = 6 h.

Fig. 10. XPS survey spectra of PIS-4 and PIS(Li)-4-4 (a, b) and high resolution of PIS(Li)-4 (inset).

adsorbents is conducted and presented in Table 4. The adsorption ca­ concentration and excessive competitive ions, it also exhibited a favor­
pacity of PIS-4 in pure Li+ solution was comparable and even higher able adsorption capacity. More importantly, its equilibrium time (4 h in
than those of major reported composite materials, such as HMO/Al2O3 Li+ solution and 6 h in geothermal water) were much shorter than those
[50], HMO/AL [51], HMO/CTS [52], C@Li4Ti5O12 [53] and HTO/PAN of state-of-the-art composite adsorbents (12–168 h). These results
[54]. When it was applied in geothermal water with low Li+ placed PIS-4 at the outstanding material for lithium recovery. Thus, PIS-

Table 4
Comparison of PIS-4 in this work with other reported Li+ adsorbents.
Adsorbents Sample Li+ concentration Adsorption capacity Equilibrium time (h) Ref.
(mg/L) (mg/g)

HMO/Al2O3 Seawater 29.8 6.2 72 [50]


HMO/AL LiCl solution 29.4 24 48 [51]
HMO/CTS Seawater 30 11.4 168 [52]
C@Li4Ti5O12 LiOH solution 1197.5 28.5 2 [53]
HTO/PAN Li+ solution 70 32 24 [54]
HTO/PVA LiCl/LiOH solution 7 12 12 [38]
PVC-HTO Geothermal water 25.8 11.35 12 [11]
PIS-4 Li+ solution – 34.23a 4 this work
PIS-4 Geothermal water 25.8 12.29 6 this work
a
calculated by Langmuir isotherm model.

8
S. Chen et al. Chemical Engineering Journal 410 (2021) 128320

4 could be regarded as a promising candidate for the recovery of Li+ development of the electric vehicle industry, Renew Sustain. Energy Rev. 16
(2012) 1735–1744.
from geothermal water.
[11] H. Lin, X. Yu, M. Li, J.i. Duo, Y. Guo, T. Deng, Synthesis of Polyporous Ion-Sieve
and Its Application for Selective Recovery of Lithium from Geothermal Water, ACS
4. Conclusions Appl. Mater. Interfaces 11 (2019) 26364–26372.
[12] M. Pasta, A. Battistel, F. La Mantia, Batteries for lithium recovery from brines,
Energy Environ. Sci. 5 (2012) 9487–9491.
In this work, a novel granular and porous titanium-based lithium ion [13] A. Kumar, H. Fukuda, T.A. Hatton, J.H. Lienhard V, Lithium Recovery from Oil and
sieve (PIS) was prepared via agar-assisted method and used for lithium Gas Produced Water: A Need for a Growing Energy Industry, ACS Energy Lett. 4
recovery from geothermal water. The agar could not only granulate (2019) 1471–1474.
[14] J.W. An, D.J. Kang, K.T. Tran, M.J. Kim, T. Lim, T. Tran, Recovery of lithium from
powdery ion sieve to improve its post-separation performance, but also Uyuni salar brine, Hydrometallurgy 117 (2012) 64–70.
generate uniform mesoporous structures. The porous structures greatly [15] X. Yu, X. Fan, Y. Guo, T. Deng, Recovery of lithium from underground brine by
facilitate Li+ to diffusion into the inner adsorbent to be captured. In this multistage centrifugal extraction using tri-isobutyl phosphate, Sep. Purif. Technol.
211 (2019) 790–798.
case, this easily separated PIS has high adsorption capacity (25.8 mg/g) [16] S.Q. Chen, D.L. Gao, X.P. Yu, Y.F. Guo, T.L. Deng, Thermokinetics of lithium
and fast kinetic (equilibrium time 6 h) in geothermal water. The sepa­ extraction with the novel extraction systems (tri-isobutyl phosphate + ionic liquid
ration factors of competitive ions related to Li+ were 1162.3, 273.7 and + kerosene), J. Chem. Thermodynamics 123 (2018) 79–85.
[17] G. Liu, Z. Zhao, L. He, Highly selective lithium recovery from high Mg/Li ratio
328.5 for Na+, K+ and Ca2+, respectively. Moreover, PIS could be brines, Desalination 474 (2020).
facilely eluted by HCl solution with a steady adsorption performance in [18] B.K. Pramanik, L.D. Nghiem, F.I. Hai, Extraction of strategically important
five recycle times. Therefore, PIS successfully improved the post- elements from brines: Constraints and opportunities, Water Res. 168 (2020).
[19] X. Li, Y. Mo, W. Qing, S. Shao, C.Y. Tang, J. Li, Membrane-based technologies for
separation performance of Ti-LIS with the guarantee of high capacity
lithium recovery from water lithium resources: A review, J. Membr. Sci. 591
and fast kinetic, which could be regarded as a promising candidate for (2019).
lithium recovery from geothermal water and even salty sea water. [20] Y.i. Guo, Y. Ying, Y. Mao, X. Peng, B. Chen, Polystyrene Sulfonate Threaded
Moreover, this simple and green method to prepare porous granular through a Metal-Organic Framework Membrane for Fast and Selective Lithium-Ion
Separation, Angew. Chem. Int. Ed. 55 (2016) 15120–15124.
materials could be easily popularized and expected to be used for the [21] M.S. Palagonia, D. Brogioli, F. La Mantia, Lithium recovery from diluted brine by
preparation of a variety of functionalized adsorbents for industrial means of electrochemical ion exchange in a flow-through-electrodes cell,
applications. Desalination 475 (2020).
[22] X.-B. Cheng, T.-Z. Hou, R. Zhang, H.-J. Peng, C.-Z. Zhao, J.-Q. Huang, Q. Zhang,
Dendrite-Free Lithium Deposition Induced by Uniformly Distributed Lithium Ions
Declaration of Competing Interest for Efficient Lithium Metal Batteries, Adv. Mater. 28 (2016) 2888–2895.
[23] A. Battistel, M.S. Palagonia, D. Brogioli, F. La Mantia, R. Trócoli, Electrochemical
Methods for Lithium Recovery: A Comprehensive and Critical Review, Adv. Mater.
The authors declare that they have no known competing financial 32 (2020) 1905440.
interests or personal relationships that could have appeared to influence [24] Y. Huang, R. Wang, Green recovery of lithium from water by a smart imprinted
the work reported in this paper. adsorbent with photo-controlled and selective properties, Chem. Eng. J. 378
(2019).
[25] A.L. Gao, X.J. Hou, Z.H. Sun, S.P. Li, H.Q. Li, J.B. Zhang, Lithium-desorption
Acknowledgements mechanism in LiMn2O4, Li1.33Mn1.67O4, and Li1.6Mn1.6O4 according to
precisely controlled acid treatment and density functional theory calculations, J.
Mater. Chem. A., 2019, 7, 20878-20890.
The supports from the National Natural Science Foundation of China
[26] M. Dou, M. Fyta, Lithium adsorption on 2D transition metal dichalcogenides:
(21773170, 22073068 and 21901183), the Major Special Project of towards a descriptor for machine learned materials design, J. Mater. Chem. A
Tibet Autonomous Region (XZ201801-GB-01), Innovation Project of (2020), https://doi.org/10.1039/D0TA04834H.
[27] B. Swain, Recovery and recycling of lithium: A review, Sep. Purif. Technol. 172
Excellent Doctorial Dissertation of Tianjin University of Science and
(2017) 388–403.
Technology (201909) and the Yangtze Scholars and Innovative Research [28] X. Xu, Y. Chen, P. Wan, K. Gasem, K. Wang, T. He, H. Adidharma, M. Fan,
Team of the Chinese University (IRT_17R81) are acknowledged. Extraction of lithium with functionalized lithium ion-sieves, Prog. Mater Sci. 84
(2016) 276–313.
[29] S. Wei, Y. Wei, T. Chen, C. Liu, Y. Tang, Porous lithium ion sieves nanofibers:
Appendix A. Supplementary data General synthesis strategy and highly selective recovery of lithium from brine
water, Chem. Eng. J. 379 (2020).
Supplementary data to this article can be found online at https://doi. [30] M.B. Bajestani, A. Moheb, M. Masigol, Simultaneous Optimization of Adsorption
Capacity and Stability of Hydrothermally Synthesized Spinel Ion Sieve Composite
org/10.1016/j.cej.2020.128320. Adsorbents for Selective Removal of Lithium from Aqueous Solutions, Ind. Eng.
Chem. Res. 58 (2019) 12207–12215.
References [31] D.H. Snydacker, V.I. Hegde, M. Aykol, C. Wolverton, Computational Discovery of
Li–M–O Ion Exchange Materials for Lithium Extraction from Brines, Chem. Mater.
30 (2018) 6961–6968.
[1] D. Draaisma, Lithium: A doctor, a drug, and a breakthrough, Nature 572 (2019)
[32] T. Ryu, Y. Haldorai, A. Rengaraj, J. Shin, H.-J. Hong, G.-W. Lee, Y.-K. Han, Y.
584–585.
S. Huh, K.-S. Chung, Recovery of Lithium Ions from Seawater Using a Continuous
[2] E. Pomerantseva, F. Bonaccorso, X. Feng, Y.i. Cui, Y. Gogotsi, Energy storage: The
Flow Adsorption Column Packed with Granulated Chitosan–Lithium Manganese
future enabled by nanomaterials, Science 366 (2019) 969.
Oxide, Ind. Eng. Chem. Res. 55 (2016) 7218–7225.
[3] X. Zhang, A. Wang, X. Liu, J. Luo, Dendrites in lithium metal anodes: suppression,
[33] H.-J. Hong, I.-S. Park, T. Ryu, J. Ryu, B.-G. Kim, K.-S. Chung, Granulation of
regulation, and elimination, Acc. Chem. Res. 52 (2019) 3223–3232.
Li1.33Mn1.67O4 (LMO) through the use of cross-linked chitosan for the effective
[4] Y. Shi, L.L. Peng, Y. Ding, Y. Zhao, G.H. Yu, Nanostructured conductive polymers
recovery of Li+ from seawater, Chem. Eng. J. 234 (2013) 16–22.
for advanced energy storage, Chem. Soc. Rev. 44 (2015) 6684–6696.
[34] G.M. Nisola, L.A. Limjuco, E.L. Vivas, C.P. Lawagon, M.J. Park, H.K. Shon,
[5] H. Zhao, X. Yu, J. Li, B.o. Li, H. Shao, L. Li, Y. Deng, Film-forming electrolyte
N. Mittal, I.W. Nah, H. Kim, W.-J. Chung, Macroporous flexible polyvinyl alcohol
additives for rechargeable lithium-ion batteries: progress and outlook, J. Mater.
lithium adsorbent foam composite prepared via surfactant blending and cryo-
Chem. A 7 (2019) 8700–8722.
desiccation, Chem. Eng. J. 280 (2015) 536–548.
[6] V. Do, Deepika, M.S. Kim, M.S. Kim, K.R. Lee, W.I. Cho, Carbon Nitride Phosphorus
[35] M.J. Park, G.M. Nisola, A.B. Beltran, R.E.C. Torrejos, J.G. Seo, S.-P. Lee, H. Kim,
as an Effective Lithium Polysulfide Adsorbent for Lithium–Sulfur Batteries, ACS
W.-J. Chung, Recyclable composite nanofiber adsorbent for Li+ recovery from
Appl. Mater. Interfaces 11 (2019) 11431–11441.
seawater desalination retentate, Chem. Eng. J. 254 (2014) 73–81.
[7] R. Fang, K.e. Chen, L. Yin, Z. Sun, F. Li, H.-M. Cheng, The Regulating Role of
[36] Q. Jia, J. Wang, R. Guo, Preparation and characterization of porous HMO/PAN
Carbon Nanotubes and Graphene in Lithium-Ion and Lithium-Sulfur Batteries, Adv.
composite adsorbent and its adsorption–desorption properties in brine, J Porous
Mater. 31 (2019).
Mater 26 (2019) 705–716.
[8] W. Zhang, Q. Wu, J. Huang, L. Fan, Z. Shen, Y.i. He, Q.i. Feng, G. Zhu, Y. Lu,
[37] G. Zhu, P. Wang, P.F. Qi, C.J. Gao, Adsorption and desorption properties of Li+ on
Colossal Granular Lithium Deposits Enabled by the Grain-Coarsening Effect for
PVC-H1.6Mn1.6O4 lithium ion-sieve membrane, Chem. Eng. J. 235 (2014) 340–348.
High-Efficiency Lithium Metal Full Batteries, Adv. Mater. 24 (2020), 2001740.
[38] L.A. Limjuco, G.M. Nisola, C.P. Lawagon, S.-P. Lee, J.G. Seo, H. Kim, W.-J. Chung,
[9] S.E. Kesler, P.W. Gruber, P.A. Medina, G.A. Keoleian, M.P. Everson, T.
H 2 TiO 3 composite adsorbent foam for efficient and continuous recovery of Li +
J. Wallington, Global lithium resources: relative importance of pegmatite, brine
from liquid resources, Colloids Surf., A 504 (2016) 267–279.
and other deposits, Ore Geol. Rev. 48 (2012) 55–69.
[10] C. Grosjean, P.H. Miranda, M. Perrin, P. Poggi, Assessment of world lithium
resources and consequences of their geographic distribution on the expected

9
S. Chen et al. Chemical Engineering Journal 410 (2021) 128320

[39] Y. Zhu, M. Liang, H. Li, H. Ni, L. Li, Q. Li, Z. Jiang, A mutant of Pseudoalteromonas [48] S.Q. Chen, J.Y. Hu, J. Shi, M.X. Wang, Y.F. Guo, M.L. Li, J. Duo, T.L. Deng,
carrageenovora arylsulfatase with enhanced enzyme activity and its potential Composite hydrogel particles encapsulated ammonium molybdophosphate for
application in improvement of the agar quality, Food Chem. 320 (2020) 126652. efficiently cesium selective removal and enrichment from wastewater, J. Hazard.
[40] X. Li, Y. Chao, L. Chen, W. Chen, J. Luo, C. Wang, P. Wu, H. Li, W. Zhu, Taming Mater. 371 (2019) 694–704.
wettability of lithium ion sieve via different TiO2 precursors for effective Li [49] G. Xiong, B.-B. Wang, L.-X. You, B.-Y. Ren, Y.-K. He, F.u. Ding, I. Dragutan,
recovery from aqueous lithium resources, Chem. Eng. J. 392 (2020) 123731. V. Dragutan, Y.-G. Sun, Hypervalent silicon-based, anionic porous organic
[41] L.Y. Zhang, R.F. Wang, H.Q. Song, H. Xie, H.F. Fan, P.G. Sun, L. Du, Numerical polymers with solid microsphere or hollow nanotube morphologies and
investigation of techno-economic multiobjective optimization of geothermal water exceptional capacity for selective adsorption of cationic dyes, J. Mater. Chem. A 7
reservoir development: A case study of China, Water 11 (2019). (2019) 393–404.
[42] X.W. Li, L.L. Chen, Y.H. Chao, W. Chen, J. Luo, J. Xiong, F.X. Zhu, X.Z. Chu, H. [50] H.J. Hong, T. Ryu, I.S. Park, M. Kim, J. Shin, B.G. Kim, K.S. Chung, Highly porous
M. Li, W.S. Zhu, Amorphous TiO2-derived large-capacity lithium ion sieve for and surface-expanded spinel hydrogen manganese oxide (HMO)/Al2O3 composite
lithium recovery, Chem. Eng. Technol. 43 (2020) 1784–1791. for effective lithium (Li) recovery from seawater, Chem. Eng. J. 337 (2018)
[43] H. Li, M. Qu, Y. Hu, Preparation of spherical Li4SiO4 pellets by novel agar method 455–461.
for high-temperature CO2 capture, Chem. Eng. J. 380 (2020) 122538. [51] P. Koilraj, S.M. Smith, Q. Yu, S. Ulrich, K. Sasaki, Encapsulation of powdery spinel-
[44] D. Gu, W. Sun, G. Han, Q. Cui, H. Wang, Lithium ion sieve synthesized via an type Li+ ion sieve derived from biongenic manganese oxide in alginate beads,
improved solid state method and adsorption performance for West Taijinar Salt Powder Technol. 301 (2016) 1201–1207.
Lake brine, Chem. Eng. J. 350 (2018) 474–483. [52] M. Moazeni, H. Hajipour, M. Askari, M. Nusheh, Hydrothermal Synthesis and
[45] I. Langmuir, The adsorption of gases on plane surface of glass, mica and platinum, characterization of titanium dioxide nanotubes as novel lithium adsorbents, Mater.
J. Am. Chem. Soc. 40 (1918) 1361–1403. Res. Bull. 61 (2015) 70–75.
[46] H.M.F. Freundlich, Über die adsorption in Lösungen, Z. Phys. Chem. 57A (1906) [53] M. Li, D.L. Lu, J.L. Zhang, L.Z. Wang, Yolk-shell structured composite for fast and
385–470. selective lithium ion sieving, J. Colloid Interf. Sci. 520 (2018) 33–40.
[47] P.A. Kumar, M. Ray, S. Chakraborty, Adsorption behaviour of trivalent chromium [54] C.P. Lawagon, G.M. Nisola, R.A.A.I. Cuevas, H. Kim, S.P. Lee, W.J. Chung,
on amine-based polymer aniline formaldehyde condensate, Chem. Eng. J. 149 Development of high capacity Li+ adsorptions from H2TiO3/polymer nanofiber
(2009) 340–347. composites: systematic polymer screening, characterization and evaluation, Ind.
Eng. Chem. Res. 70 (2019) 124–135.

10

You might also like