Next Article in Journal
Experimental Study on Proppant Backflow and Fiber Sand Control in Vertical Fracture Based on the Visual Diversion Chamber Simulation
Previous Article in Journal
Upscaling of Toluene Oxidation Using Water-Sprinkled Pulsed Corona Discharge and Photocatalysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fabrication of Spinel-Type H4Ti5O12 Ion Sieve for Lithium Recovery from Aqueous Resources: Adsorption Performance and Mechanism

1
School of Environmental and Municipal Engineering, Qingdao University of Technology, Qingdao 266520, China
2
SINOPEC Research Institute of Dalian Petroleum and Petrochemicals Co., Ltd., National Engineering Research Center for Industrial Wastewater Harmlessness and Resource Utilization, Dalian 116045, China
*
Author to whom correspondence should be addressed.
Processes 2025, 13(9), 2981; https://doi.org/10.3390/pr13092981
Submission received: 29 August 2025 / Revised: 16 September 2025 / Accepted: 16 September 2025 / Published: 18 September 2025
(This article belongs to the Section Chemical Processes and Systems)

Abstract

Lithium (Li) ion sieve is considered to have great potential in the selective extraction of Li+ from complex Li+-containing brine owing to its cost-effectiveness, excellent adsorption performance, and environmental friendliness. Nevertheless, the defects of complex regulation and control of technological parameters in the preparation process of Li ion sieve and poor recycling efficiency limit its application. In this study, spinel-type H4Ti5O12 ion sieves (HTO) were successfully prepared through a high-temperature solid-state method for recovering Li+ from aqueous resources. Through the experiment of optimizing the key preparation process parameters of HTO, it was found that the optimum preparation conditions were as follows: lithium ion source of CH3COOLi‧H2O, calcination temperature of 800 °C, and acid (HCl) washing concentration of 0.3 mol/L. The uptake of Li+ by HTO aligned with the pseudo-second-order kinetic model, which was a chemical adsorption process controlled by reversible Li–H ion exchange reaction. HTO exhibited extremely high regeneration cycle characteristics, and after five cycles, it retained 96.06% of its initial adsorption capacity. The present work highlighted that spinel-type HTO has high industrial application potential in the field of Li+ recovery from oilfield brine.

1. Introduction

Lithium (Li), a crucial metal for global advancement, is valued for low density (0.534 g/cm3), strong fatigue resistance, and low redox potential (−3.04 V). These properties make it indispensable in key industries, including energy storage, ceramics, glass, grease, refrigerant and nuclear applications [1]. Lithium reserves are chiefly distributed between geological ores and liquid sources, notably salt lakes and certain types of industrial wastewater. Contemporary lithium production derives roughly 70% of its output from salt lake resources [2]. As we all know, aqueous resources exceed ore deposits in availability and the latter are more expensive for Li+ exploitation [3]. Therefore, it is essential to extract Li+ from aqueous resources, but this method still faces challenges such as difficult adjustment of preparation process parameters, presence of multiple competing ions (notably Na+, K+, Ca2+, Mg2+ and other coexisting cations), and poor recycling efficiency that limit its application [4]. Up to now, the extraction methods of Li+ resources mainly include adsorption [5,6], precipitation [7], extraction [8], and membrane separation [9]. Among them, adsorption has emerged as the most viable technique for Li+ recovery because of its environmental friendliness, recyclability, cost-effectiveness, and high adsorption selectivity [3,10,11].
Li ion sieves, which exhibit molecular memory properties enabling selective Li+ adsorption, have emerged as promising adsorbents for Li+ recovery [12]. These substances are primarily classified as either manganese-type or titanium-type Li+ sieves [13]. Tian et al. [14] successfully synthesized H1.33Mn1.67O4 for recovering Li+ from shale gas flowback produced water and the results showed that the adsorption capacity of H1.33Mn1.67O4 towards Li+ was 13.27 mg/g. Although manganese-based Li+ sieves are easy to prepare and have good adsorption behavior towards Li+, there are risks such as manganese dissolution which may pollute water resources in industrial production. Titanium-derived Li+ sieves exhibit several beneficial characteristics due to their strong Ti–O linkages: structural permanence, highly selective adsorption, outstanding acid tolerance, and excellent cyclic stability [15], which can overcome the disadvantages of manganese-based Li+ sieves. The spinel structure of Li4Ti5O12 is more stable than that of layered Li2TiO3 [16]. Li et al. [17] effectively fabricated Yolk-shell C@Li4Ti5O12 microspheres, with experimental results demonstrating a peak Li+ adsorption capacity of 28.46 mg/g. The adsorption behavior was accurately modeled using both Freundlich isotherm and pseudo-second-order kinetic equations. However, the defect in complex regulation and control of technological parameters in the preparation process of Li ion sieve and poor recycling efficiency hinder the development of H4Ti5O12 (HTO). Therefore, there is an urgent need to enhance the titanium-based Li+ sieves adsorption performance towards Li+ through optimization of the purity, morphology, and surface properties of HTO during its preparation process. Lithium source, calcination temperature, and acid washing concentration are three important parameters in the preparation of HTO. At present, LiOH and Li2CO3 are mostly used as Li sources for the preparation of HTO [18,19,20,21,22]. But the regulation law of the selection of Li sources on the microstructure of HTO has not been systematically explored, which makes the selection of Li sources lack reference. In addition, the mechanism of the joint regulation of Li source, calcination temperature, and acid washing concentration on the microstructure and adsorption performance of HTO is still unclear.
In the present work, spinel-structured Li+ sieve (HTO) was synthesized via a high-temperature solid-state approach, and the effects of key technological conditions such as lithium source, calcination temperature, and acid washing concentration on the crystal purity, morphological characteristics, and adsorption performance of the HTO were investigated. The adsorption capacity, adsorption selectivity, and cycle performance of the spinel-type HTO in simulated Li+-containing oilfield brine were also determined. Additionally, the adsorption mechanism of the spinel-type HTO was further discussed. This study will provide theoretical reference for the optimal preparation of spinel-type Li+ sieve and technical support for the separation and recovery of high-value Li resources in aqueous resources.

2. Materials and Methods

2.1. Reagents

Titanium dioxide (TiO2, anatase type), lithium carbonate (Li2CO3), lithium hydroxide monohydrate (LiOH‧H2O), and lithium acetate dihydrate (CH3COOLi‧2H2O) were all analytically pure and obtained from Shanghai Aladdin Biochemical Technology Co., Ltd., Shanghai, China. Hydrochloric acid (HCl), lithium chloride (LiCl), sodium chloride (NaCl), potassium chloride (KCl), magnesium chloride (MgCl2), and calcium chloride (CaCl2) were all analytically pure and provided by National Medicine Group Chemical Reagent Co., Ltd., Shanghai, China.

2.2. Preparation of HTO

The HTO was fabricated using a high-temperature solid-state method [19]. Firstly, the lithium source (Li2CO3 or LiOH‧H2O or CH3COOLi‧H2O) and the titanium source (nano-TiO2) were fully mixed according to the ratio of Ti/Li of 5:4.2, and an appropriate amount of anhydrous ethanol was added. Next, the obtained mixture was placed into an agate grinding bowl and ground for 40 min to achieve an even mixture, and placed in a 50 °C oven and dried for 12 h to remove residual ethanol. Then, the dried sample was ground into powder again and passed through a 200-mesh sieve, then placed into a corundum crucible for high-temperature calcination in a tube furnace (SK2-2-12TPA2 tubular furnace, Shanghai Zhuode Technology Co., Ltd., Shanghai, China) for 10 h. Once cooled to room temperature, the resulting powder was ground and sieved through a 200-mesh sieve to obtain Li4Ti5O12 (LTO), the precursor of the spinel-type H4Ti5O12 lithium ion sieve. Subsequently, the precursor of LTO was placed into a certain concentration of hydrochloric acid according to the solid–liquid ratio of 100 mL/g for 24 h of pickling, followed by repeated washing with deionized water until a neutral pH was achieved. Finally, the obtained product was oven-dried at 50 °C for 12 h to obtain the spinel-type H4Ti5O12 (HTO). According to different lithium ion sources, the HTOs were named as HTO-Li2CO3, HTO-LiOH, and HTO-LiAc, respectively.

2.3. Characterization

The crystal structures of the HTO before and after adsorption of Li+ were analyzed by an X-ray diffractometer (XRD, Rigaku Smart lab SE, Akishima, Japan). The morphologies of the samples were characterized by a scanning electron microscope (SEM, Zeiss Sigma500, Oberkochen, Germany). The functional group of the materials was recorded on a Fourier transform infrared spectrometer (FT-IR, ThermoScientific Nicolet, iS50, Waltham, MA, USA). The surface area and pore distribution of the samples were examined by a 3Fle multifunctional nitrogen adsorption/desorption analyzer (Mike, Atlanta, GA, USA). The element distribution and valence state of the materials were measured with an X-ray photoelectron spectroscope (XPS, Thermo Fisher ESCALAB 250Xi, Waltham, MA, USA).

2.4. Li+ Adsorption Experiments

Batch adsorption tests were conducted by using 0.1 g of HTO adsorbent in 100 mL of Li+-containing solution (pH = 12) at 25 °C, and the experiments were performed in an XMTD-4000 constant temperature heating magnetic stirrer (Yongguangming Medical Instrument Co., Ltd. Beijing, China) with shaking speed of 500 r/min. After adsorption for a desired time, a 1 mL aliquot of the supernatant was collected and passed through a 0.45 μm pore-size membrane, and then diluted to a suitable multiple for Li+ concentration determination. All the experiments were conducted in triplicate. The Li+ concentration was quantified using a LICAP PRO ICP-OES instrument (Thermo Fisher Scientific, Waltham, MA, USA). Calculation of the HTO adsorption capacity was performed with Equation (1):
q t = ( C 0 C t ) × V m
where qt is the HTO Li+ adsorption capacity at time t, mg/g; C0 denotes the initial Li+ concentration, mg/L; Ct represents Li+ concentration at time t, mg/L; V indicates the volume of Li+ solution, L; and m is the mass of HTO adsorbent, g.

2.5. Adsorption Kinetics

Adsorption kinetic tests were conducted by immersing 0.1 g of HTO into 100 mL Li+-containing solution (initial concentration: 1000 mg/L) and sampling at desired time intervals (0–48 h) to examine the Li+ concentration in the solution. Three kinetic models (pseudo-first-order, pseudo-second-order, and Evolich) were applied to analyze the HTO Li+ adsorption data and their equations are written as Equations (2)–(4) [23]:
ln ( q e q t ) = ln q e k 1 2.303 t
t q t = 1 q e 2 k 2 + t q e
q t = 1 β ln α β + 1 β ln t
where qe (mg/g) is the equilibrium Li+ uptake by HTO; qt (mg/g) is the instantaneous adsorption capacity of HTO towards Li+; k1 (min−1) and k2 (mg/kg/min−1) denote rate constants for pseudo-first-order and pseudo-second-order kinetics, respectively; and t represents the adsorption time (min).
q t = k d i t 1 / 2 + C i
where kdi (mg/(g‧min1/2)) denotes the intra-particle diffusion rate constant and Ci (mg/g) represents the boundary layer thickness [24,25].

2.6. Adsorption Isotherms

Adsorption isotherm experiments were carried out by immersing 100 mg of HTO into 100 mL Li+ solution (100–500 mg/L) for 24 h. The data were fitted using Langmuir (Equation (6)) and Freundlich (Equation (7)) models:
C e q e = 1 K L q max + C e q max
ln q e = ln K F + 1 n ln C e
where KL (L/mg) denotes the Langmuir equilibrium constant; Ce (mg/g) is the Li+ equilibrium concentration; qmax (mg/g) is the HTO maximum adsorption capacity at equilibrium; and KF (mg/g) and n correspond to the Freundlich coefficient and heterogeneity factor, respectively.

3. Results and Discussion

3.1. Effect of Key Preparation Process Parameters on the Microstructure and Adsorption Behavior of HTO

3.1.1. Effect of Lithium Source

The morphology of HTO-Li2CO3, HTO-LiOH, and HTO-LiAc was observed by SEM and the results are presented in Figure 1. HTO-Li2CO3 (Figure 1A) had a regular cubic shape, compact structure, and serious particle agglomeration. The particle size of HTO-Li2CO3 was about 600–800 nm (Figure 1A). The morphology of each grain of HTO-LiOH was similar (Figure 1B), and the particle agglomeration was also serious (particle size: 400–600 nm). Compared with HTO-Li2CO3 and HTO-LiOH, the particle size of HTO-LiAc (300–400 nm) was smaller, and its hierarchy was clearer (Figure 1C).
The specific surface area properties and pore-size distributions of HTO-Li2CO3, HTO-LiOH and HTO-LiAc were analyzed via N2 adsorption/desorption isotherms, as illustrated in Figure 1D–I. Regarding the specific surface area, the three samples displayed these sequences: HTO-LiAc (20.1775 m2/g) > HTO-LiOH (18.2940 m2/g) > HTO-Li2CO3 (16.9275 m2/g). The pore size of HTO-based adsorbent with different lithium sources shows that their pore size ranges belonged to the mesoporous range, and their average pore size was about 7 nm. Adsorption capacity is augmented by higher surface area through increased exposure of active sites [26]. Regarding adsorption efficiency, the materials in Figure 1J showed this progression: HTO-LiAc > HTO-LiOH > HTO-Li2CO3. In addition, the decomposition temperature of HTO-LiAc may be lower, and the gas production during its calcination process was higher, which made it easier to create a loose and porous structure and endowed it with more active sites. To sum up, among the three common Li sources, the best Li source for preparing spinel-type HTO is CH3COOLi‧H2O.

3.1.2. The Effect of Calcination Temperature

During the synthesis of spinel-type HTO by a high-temperature solid-state method, calcination temperature is a crucial control factor. LTO was prepared by calcination at different temperatures of 700 °C, 800 °C and 900 °C with CH3COOLi‧H2O as lithium source and nano-TiO2 as titanium source. The XRD patterns of LTO-700, LTO-800, and LTO-900 are shown in Figure 2A. Compared with the standard PDF card of spinel Li4Ti5O12 crystal (PDF#49-0207), it can be seen that spinel LTO was formed at the above three calcination temperatures. Impurity peaks were detected in LTO-700 and LTO-900 at 25.27° and 27.43°, respectively. LTO-800 had the highest purity and the best crystallinity, which indicated that other impurity crystal forms were easily produced when the temperature was too high or too low. On the whole, the peak values of LTO-800 were higher than those of LTO-700 and LTO-900, which suggested that the LTO crystal generated at the calcination temperature of 800 °C was more orderly and had a more stable structure.
The spinel-type HTOs were further obtained by eluting the LTO precursor prepared at 700 °C, 800 °C, and 900 °C with 0.3 mol/L HCl. The SEM images of HTOs were shown in Figure 2B–D. The morphology of HTO-700 appeared irregular and disordered, suggesting that the reaction was incomplete. The shape and size of HTO-800 were uniform, and the agglomeration phenomenon was light. The agglomeration of HTO-900 was serious. Based on the above analysis, the optimum calcination temperature is 800 °C.

3.1.3. Effect of Acid Washing Concentration

The acid washing step is another key step in the process of preparing spinel-type HTO by the high-temperature solid-state method. Following successful synthesis of the LTO-800 precursor, the Li+ in the precursor was removed via 24 h of HCl treatment at room temperature to obtain HTO, leaving holes with sizes matching the Li+. The effects of acid washing concentration (0.1–0.5 mol/L) on Li+ elution rate and Ti4+ dissolution loss are shown in Figure 3. According to Figure 3, at 0.3 mol/L HCl, the Li+ elution rate of LTO-800 was 94.11% which is much higher than that of HTO obtained with 0.1 mol/L HCl (75.84%). Additionally, the Ti4+ loss rate of LTO-800 precursor was only 0.39%, which was much lower than that of HTO obtained with 0.5 mol/L HCl (0.70%). It was attributed that the high HCl concentration would destroy the structure of LTO precursor. Usually, the ideal Li ion sieves should have small Ti4+ dissolution and large Li+ extraction to increase the adsorption behavior [20,22]. The following experiments employed the optimal 0.3 mol/L HCl to prepare the spinel-type HTO from LTO-800.

3.2. Adsorption Studies

3.2.1. Adsorption Kinetics

As a critical indicator of adsorption efficiency, the adsorption rate directs adsorbent development and enhancement [26,27]. The relationship between adsorption time and Li+ removal by HTO is illustrated in Figure 4A. The Li+ adsorption capacity of HTO exhibited a time-dependent increase, reaching maximum uptake after 24 h. Three kinetic models (pseudo-first-order, pseudo-second-order, and Evolich) were applied to analyze the HTO Li+ adsorption data, with fitting results presented in Figure 4B–D and Table 1.
As shown in Figure 4B–D and Table 1, the correlation coefficient (R2) values of pseudo-second-order (R2 = 0.997) were higher than that of pseudo-first-order (R2 = 0.985) and Evolich kinetic model (R2 = 0.980), indicating that the pseudo-second-order kinetics provided the most accurate fit to the Li+ adsorption data. These results suggest that Li+ adsorption by HTO was predominantly chemisorption-driven, mediated by Li+–H+ ion exchange.
To further study the diffusion mechanism of Li+ adsorbed by HTO, experimental data were fitted using the intra-particle diffusion model (Equation (5)) [28], with fitting results presented in Figure 4E and Table 2.
According to Figure 4E and Table 2, qt against t1/2 presented three-phase curves and the obtained Ci values were not zero, indicating that besides intra-particle diffusion, there were other diffusion processes involved in the process of Li+ adsorbed by HTO. The uptake of Li+ by HTO involved three stages including film diffusion, intra-particle diffusion, and chemisorption of pore surface stages. The order of diffusion rate constants (kdi) calculated by the intra-particle diffusion model in the above three stages were in the following order: kd1 > kd2 > kd3. As the adsorption reaction progressed, the reduction in Li+ solution concentration led to greater mass transfer resistance and a corresponding decline in adsorption rate. Meanwhile, with the progress of the reaction, the boundary layer effect in the system was enhanced, and the Ci values were increased (C1 < C2 < C3) [29].

3.2.2. Adsorption Isotherms

The Langmuir (Equation (6)) and Freundlich (Equation (7)) isotherm models [30,31] were employed to elaborate the adsorption process of Li+ onto HTO and the fitting results are shown in Figure 5A,B and Table 3.
The Freundlich isotherm model (R2 = 0.9858) demonstrated superior fit to the HTO adsorption data for Li+ compared to the Langmuir model (R2 = 0.9741), as evidenced by the higher R2 value shown in Figure 5 and Table 4. Freundlich isotherm analysis revealed an n value exceeding 1, indicating favorable multilayer adsorption of Li+ by HTO through chemical interactions. The comparison of HTO and other adsorbents in recent publications is shown in Table 4. It can be seen that the adsorption capacity of HTO is at a moderate level. Considering the long time required for HTO to reach adsorption equilibrium, further efforts should be made to improve the performance of adsorbents in further exploration.

3.2.3. Adsorption Selectivity

Given that in actual oilfield produced water, Li+ usually coexists with Na+, K+, Ca2+ and Mg2+, which will reduce its adsorption capacity because of the competition of other coexisting ions for adsorption sites [20]. The HTO ion selectivity was tested by mixing 0.1 g of HTO with 100 mL simulation solution containing Li+, Na+, K+, Ca2+ and Mg2+. Detailed information regarding the simulated oilfield produced water’s cation content (Li+, Na+, K+, Ca2+ and Mg2+) is exhibited in Table 5. After the adsorption reaction was completed, metal ion concentrations were quantified using LICAP PRO ICP-OES (Thermo Fisher Scientific, USA). The selectivity of HTO towards Li+ was evaluated by the coefficient (Kd) and separation factor ( α L i M ) using Equations (8) and (9) [21]:
K d = ( C 0 C e ) V C e m
α L i M = K d L i K d M
where Ce is the equilibrium concentration of Li+, mg/L; and α L i M represents Li to M ion (Na, K, Ca and Mg) separation factors.
As evidenced in Table 5, Li+ exhibited superior Kd value (173.39 mL/g) over competing ions (Na+: 2.56, K+: 2.41, Ca2+: 1.75, Mg2+: 8.03 mL/g). High competing ion concentrations decreased the HTO Li+ adsorption efficiency compared with that in the single Li+ solution, but the decrease was not significant. The differences in Kd values of different metal ions could result from size differences, as Na+ (102 pm), K+ (138 pm), and Ca2+ (100 pm) all possessed larger ionic radii than Li+ (76 pm) [19]. As a result, the large cations cannot overcome the steric hindrance and diffuse into the pore channel of HTO, but only anchor on the outer surface of HTO. In particular, the Kd and α L i M g values of Mg2+ were 8.03 mL/g and 21.60, respectively. The similar sizes but differing hydration energies (Li+: −515 kJ/mol; Mg2+: −1922 kJ/mol) explained why Mg2+ could not readily exchange with H+ in HTO [18]. The α L i M values between Li+ and Na+, K+, Ca2+ and Mg2+ were all greater than 1, which further indicated that the spinel-type HTO had relatively high adsorption selectivity. Therefore, the spinel-type HTO is a potential candidate for recovering Li+ in actual Li+-containing oilfield produced water.

3.2.4. Recyclability

The regeneration and reuse performance of adsorbent is of great significance to evaluate its practical application in industry, because it will directly affect the cost of adsorbent [21]. In the present work, the recyclability of HTO was investigated and the experimental results are illustrated in Figure 6. The HTO regeneration rate decreased marginally with successive cycles, and the regeneration rate was 96.06% after five times of reuse (Figure 6A). The Li+ removal rate in the process of HTO acid washing has been maintained at a high level, with an average Li+ removal rate of 93.65% (Figure 6B). At the same time, the acid washing process resulted in low Ti4+ dissolution (average: 0.42% per cycle) (Figure 6B). Thus, HTO with excellent reusability (high regeneration rate) and chemical stability (low Ti4+ dissolution rate) can be recycled for Li+ recovery.

3.3. Adsorption Mechanism

FT-IR and XRD analyses were conducted on HTO samples pre- and post-Li+ adsorption to elucidate the adsorption mechanism, with results displayed in Figure 7. The adsorption peak around 3410 cm−1 corresponding to –OH did not change obviously before and after the adsorption reaction, which demonstrated that –OH was not involved in the HTO adsorption of Li+ ions (Figure 7A). The peak at 638.1 cm−1 in HTO was assigned to the vibration of Ti–O–Ti skeleton, and it was blue-shifted by 14.4 cm−1 after Li+ adsorption, suggesting that the crystal structure of HTO was ordered after Li+ was embedded to replace H+, that is, the crystal lattice shrank. The unchanging position of the Ti–O bending vibration peak (477.6 cm−1) before and after adsorption indicates that the HTO spinel framework remained structurally intact during Li+ incorporation. In the XRD pattern of HTO-Li+, the diffraction peaks at 18.37°, 35.61°, 43.28°, 47.39°, 62.87°, and 66.11° corresponded to the reflections of (111), (311), (400), (331), (440), and (531) crystal planes, respectively, which was consistent with the structure of spinel Li4Ti5O12 (PDF No. 49-0207). Compared with HTO, the intensity of all diffraction peaks in XRD pattern of HTO-Li+ obtained after adsorption of Li+ was significantly increased (Figure 7B), which indicated that the embedding of Li+ in the adsorption process repaired the lattice defects caused by protonation and improved the order of crystal structure. After adsorption of Li+, the distance between crystal planes of HTO was reduced, and the lattice constant a (a = b = c) was reduced from 8.4006 nm before adsorption to 8.3549 nm, which indicated that the crystal had undergone uniform changes during adsorption, and the ion exchange mechanism was verified. Elemental analysis of HTO and HTO-Li+ (at%) determined by XPS showed that the content of Li was enhanced after the adsorption reaction (Table 6), also confirming the successful adsorption of Li+.
The above change confirmed the reversible structural transformation ability of HTO, which was the basis of high recyclability of ion sieve. On the other hand, the XRD peak position of HTO-Li+ after adsorption was slightly shifted to the right compared with HTO, indicating that Li+ intercalation replaced H+ in the adsorption process. This change reflected that there was a highly reversible ion exchange in the adsorption process of Li+ by HTO.

4. Conclusions

The synthesis of spinel-structured HTO in this work employed a high-temperature solid-state approach, and the optimum preparation conditions were as follows: lithium ion source of CH3COOLi‧H2O, calcination temperature of 800 °C, and acid (HCl) washing concentration of 0.3 mol/L. The obtained HTO with good recyclability had high adsorption selectivity for Li+ with the coexisting metal ions including Na+, K+, Ca2+ and Mg2+ in the simulated oilfield produced water. Combined adsorption tests and characterization analysis showed that Li+ uptake by HTO primarily occurred through ion exchange, and the adsorption process was a multi-molecular layer chemical adsorption reaction. The present work establishes novel guidelines for synthesizing, utilizing, and understanding the adsorption mechanisms of spinel-structured HTO, and provides a new solution for the recovery of Li+ in oilfield brine.

Author Contributions

Methodology, W.M. and H.H.; Formal analysis, W.M., H.H. and G.Z.; Investigation, W.M. and H.H.; Resources, W.M. and H.H.; Data curation, W.M.; Writing—original draft, W.M.; Writing—review & editing, H.H., G.Z., X.W., Q.K. and X.S.; Visualization, G.Z.; Supervision, G.Z.; Project administration, G.Z.; Funding acquisition, G.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was founded by National Key Research and Development Program of China (2024YFC3713800), and the Qingchuang Science and Technology Program of Shandong Province, China (2022KJ164).

Data Availability Statement

The data presented in this study are available on request from the corresponding author. (The data are not publicly available due to privacy or ethical restrictions.).

Conflicts of Interest

Author Guangjin Zhu and Xueqing Wang were employed by the company SINOPEC Research Institute of Dalian Petroleum and Petrochemicals Co., Ltd., Dalian, China. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Ding, W.; Zhang, J.; Liu, Y.; Guo, Y.; Deng, T.; Yu, X. Synthesis of granulated H4Mn5O12/chitosan with improved stability by a novel cross-linking strategy for lithium adsorption from aqueous solutions. Chem. Eng. J. 2021, 426, 131689. [Google Scholar] [CrossRef]
  2. Zhang, X.; Wu, X.; Gan, J.; Yu, X.; Wang, H.; Ou, R. Building block design of thermally regenerable metal-organic framework composites for highly selective lithium adsorption. Chem. Eng. J. 2024, 499, 156352. [Google Scholar] [CrossRef]
  3. Gao, A.; Sun, Z.; Li, S.; Hou, X.; Li, H.; Wang, C.; Wu, W. Self-assembled layered lithium manganese oxide shows ultra-large adsorption capacity and high selectivity for lithium. Chem. Eng. J. 2023, 471, 144287. [Google Scholar] [CrossRef]
  4. Yu, H.; Naidu, G.; Zhang, C.; Wang, C.; Razmjou, A.; Han, D.S.; He, T.; Shon, H. Metal-based adsorbents for lithium recovery from aqueous resources. Desalination 2022, 539, 115951. [Google Scholar] [CrossRef]
  5. Bao, L.; Zhang, J.; Tang, W.; Sun, S. Synthesis and adsorption properties of metal oxide-coated lithium ion-sieve from salt lake brine. Desalination 2023, 546, 116196. [Google Scholar] [CrossRef]
  6. Fu, X.; Niu, Z.; Peng, C.; Han, H.; Sun, W.; Yue, T. Quantitative synergistic adsorption affinity of Ca(II) and sodium oleate to predict the surface reactivity of hematite and quartz. Sep. Purif. Technol. 2025, 360, 131196. [Google Scholar] [CrossRef]
  7. Wang, J.; Yue, X.; Wang, P.; Yu, T.; Du, X.; Hao, X.; Abudula, A.; Guan, G. Electrochemical technologies for lithium recovery from liquid resources: A review. Renew. Sust. Energy Rev. 2022, 154, 111813. [Google Scholar] [CrossRef]
  8. Zhang, L.; Lijuan, L.; Ji, L.; Song, X. Lithium recovery from effluent of spent lithium battery recycling process using solvent extraction. J. Hazard. Mater. 2020, 398, 122840. [Google Scholar] [CrossRef]
  9. Chen, Z.; Du, J.; Shi, J. Mxene-based lithium-ion sieve polymer membrane for sustainable lithium adsorption. Sep. Purif. Technol. 2025, 354, 129316. [Google Scholar] [CrossRef]
  10. Zhong, J.; Lin, S.; Yu, J. Li+ adsorption performance and mechanism using lithium/aluminum layered double hydroxides in low grade brines. Desalination 2021, 505, 114983. [Google Scholar] [CrossRef]
  11. Kong, Q.; Shi, X.; Ma, W.; Zhang, F.; Yu, T.; Zhao, F.; Zhao, D.; Wei, C. Strategies to improve the adsorption properties of graphene-based adsorbent towards heavy metal ions and their compound pollutants: A review. J. Hazard. Mater. 2021, 415, 125690. [Google Scholar] [CrossRef]
  12. Xu, J.; Chen, P. Preparation and characterization of lithium-ion sieve with attapulgite. Desalination 2024, 571, 117111. [Google Scholar] [CrossRef]
  13. Wang, Q.; Yang, X.; Li, C.; Zhao, C.; Huang, Y.; Qiu, S.; Qin, Y.; Shi, C. Titanium-based lithium ion sieve adsorbent H2TiO3 with enhanced Li+ adsorption properties by magnetic fe doping. Sep. Purif. Technol. 2024, 346, 127455. [Google Scholar] [CrossRef]
  14. Tian, L.; Liu, Y.; Tang, P.; Yang, Y.; Wang, X.; Chen, T.; Bai, Y.; Tiraferri, A.; Liu, B. Lithium extraction from shale gas flowback and produced water using H1.33Mn1.67O4 adsorbent. Resour. Conserv. Recy. 2022, 185, 106476. [Google Scholar] [CrossRef]
  15. Sujoto, V.S.H.; Prasetya, A.; Petrus, H.T.B.M.; Astuti, W.; Jenie, S.N.A.; Anggara, F.; Utama, A.P.; Kencana, A.Y.; Singkuang, D.E.S.; Sutijan, A.G.A.S. Advancing lithium extraction: A comprehensive review of titanium-based lithium-ion sieve utilization in geothermal brine. J. Sustain. Metall. 2024, 10, 1959–1982. [Google Scholar] [CrossRef]
  16. Wu, Q.; Ding, Z.; Wang, C.; Chen, Z.; Sui, K.; Liu, Y.; Qi, P. Hollow laminar Li4Ti5O12 nanofibers with polycrystalline property facilitate super-high and ultrafast extraction of lithium ions. Chem. Eng. J. 2024, 498, 155531. [Google Scholar] [CrossRef]
  17. Li, N.; Lu, D.; Zhang, J.; Wang, L. Yolk-shell structured composite for fast and selective lithium ion sieving. J. Colloid Interface Sci. 2018, 520, 33–40. [Google Scholar] [CrossRef] [PubMed]
  18. Yang, C.; Zhang, S.; Lu, H.; Tang, K.; Yang, J.; Shi, L.; Xiong, S.; Tang, A.; Jiang, J. Preparation of H4Ti5O12-hydroxyethyl cellulose composite for rapid adsorption of lithium in aqueous solution. Sep. Purif. Technol. 2025, 362, 131852. [Google Scholar] [CrossRef]
  19. Xiao, X.; Li, J.; Qiu, K.; Chen, M.; Zhang, X. Dual surfactants-assisted adsorption of lithium ions in liquid lithium resources on a superhydrophilic spinel-type H4Ti5O12 ion sieve. Sep. Purif. Technol. 2024, 330, 125479. [Google Scholar] [CrossRef]
  20. Zhao, B.; Qian, Z.; Qiao, Y.; Li, J.; Wu, Z.; Liu, Z. The Li(H2O)n dehydration behavior influences the Li+ ion adsorption on H4Ti5O12 with different facets exposed. Chem. Eng. J. 2023, 451, 138870. [Google Scholar] [CrossRef]
  21. Zhao, B.; Qian, Z.; Guo, M.; Wu, Z.; Liu, Z. The performance and mechanism of recovering lithium on H4Ti5O12 adsorbents influenced by (110) and (111) facets exposed. Chem. Eng. J. 2021, 414, 128729. [Google Scholar] [CrossRef]
  22. Zhao, B.; Guo, M.; Qian, F.; Qian, Z.; Xu, N.; Wu, Z.; Liu, Z. Hydrothermal synthesis and adsorption behavior of H4Ti5O12 nanorods along as lithium ion-sieves. RSC Adv. 2020, 10, 35153–35163. [Google Scholar] [CrossRef] [PubMed]
  23. Lv, M.; Yan, L.; Liu, C.; Su, C.; Zhou, Q.; Zhang, X.; Lan, Y.; Zheng, Y.; Lai, L.; Liu, X.; et al. Non-covalent functionalized graphene oxide (GO) adsorbent with an organic gelator for co-adsorption of dye, endocrine-disruptor, pharmaceutical and metal ion. Chem. Eng. J. 2018, 349, 791–799. [Google Scholar] [CrossRef]
  24. Wu, F.; Tseng, R.; Juang, R. Initial behavior of intraparticle diffusion model used in the description of adsorption kinetics. Chem. Eng. J. 2009, 153, 1–8. [Google Scholar] [CrossRef]
  25. Lei, S.; Fan, X.; Kexin, L.; Wulong, L.; Zhanxiong, L. Flexible MOF@fabrics composites: The in-situ growth of metal-organic frameworks on cotton and selectively adsorption of gold(III) from aqueous solution. Sep. Purif. Technol. 2025, 357, 129921. [Google Scholar] [CrossRef]
  26. Kong, Q.; Wang, P.; Song, B.; Lan, Y.; Ma, W.; Shi, X.; Xiao, L.; Zhu, G.; Wang, P.; Lian, J. Sludge-derived alginate-like extracellular polymers (ALE) for preparation of Fe-ALE and FeCaMg-ALE: Application to the adsorption of phosphate. Int. J. Biol. Macromol. 2024, 279, 134995. [Google Scholar] [CrossRef]
  27. Yang, K.; Yan, L.; Yang, Y.; Yu, S.; Shan, R.; Yu, H.; Zhu, B.; Du, B. Adsorptive removal of phosphate by Mg-al and Zn-Al layered double hydroxides: Kinetics, isotherms and mechanisms. Sep. Purif. Technol. 2014, 124, 36–42. [Google Scholar] [CrossRef]
  28. Chen, F.; Liang, W.; Qin, X.; Jiang, L.; Zhang, Y.; Fang, S.; Luo, D. Preparation and recycled simultaneous adsorption of methylene blue and Cu2+ co-pollutants over carbon layer encapsulated Fe3O4 /graphene oxide nanocomposites rich in amino and thiol groups. Colloids Surf. A—Physicochem. Eng. Asp. 2021, 625, 126913. [Google Scholar] [CrossRef]
  29. Ahsan, M.A.; Jabbari, V.; Islam, M.T.; Turley, R.S.; Dominguez, N.; Kim, H.; Castro, E.; Hernandez-Viezcas, J.A.; Curry, M.L.; Lopez, J.; et al. Sustainable synthesis and remarkable adsorption capacity of MOF/graphene oxide and MOF/CNT based hybrid nanocomposites for the removal of bisphenol a from water. Sci. Total Environ. 2019, 673, 306–317. [Google Scholar] [CrossRef]
  30. Cheng, X.; Huang, X.; Wang, X.; Zhao, B.; Chen, A.; Sun, D. Phosphate adsorption from sewage sludge filtrate using zinc-aluminum layered double hydroxides. J. Hazard. Mater. 2009, 169, 958–964. [Google Scholar] [CrossRef]
  31. Robati, D.; Mirza, B.; Rajabi, M.; Moradi, O.; Tyagi, I.; Agarwal, S.; Gupta, V.K. Removal of hazardous dyes-BR 12 and methyl orange using graphene oxide as an adsorbent from aqueous phase. Chem. Eng. J. 2016, 284, 687–697. [Google Scholar] [CrossRef]
  32. Long, X.; Liu, C.; Yang, Y.; Dong, B.; Chen, H.; Zhang, L.; Pan, J.; Zhou, C. Microwave-assisted hydrothermal synthesis of lithium ion-sieve for adsorption of lithium ion in coal gangue leaching solution. Colloids Surf. A 2024, 682, 132819. [Google Scholar] [CrossRef]
Figure 1. The SEM images of HTO-Li2CO3 (A), HTO-LiOH (B), and HTO-LiAc (C); BET surface area and pore size of HTO-Li2CO3 (D,E), HTO-LiOH (F,G), and HTO-LiAc (H,I), and the adsorption performance of HTO adsorbent (J).
Figure 1. The SEM images of HTO-Li2CO3 (A), HTO-LiOH (B), and HTO-LiAc (C); BET surface area and pore size of HTO-Li2CO3 (D,E), HTO-LiOH (F,G), and HTO-LiAc (H,I), and the adsorption performance of HTO adsorbent (J).
Processes 13 02981 g001
Figure 2. The XRD patterns of LTO precursor (A) and SEM images of HTO-700 (B), HTO-800 (C), and HTO-900 (D).
Figure 2. The XRD patterns of LTO precursor (A) and SEM images of HTO-700 (B), HTO-800 (C), and HTO-900 (D).
Processes 13 02981 g002
Figure 3. The effects of acid washing concentration on Li+ elution rate and Ti4+ dissolution loss.
Figure 3. The effects of acid washing concentration on Li+ elution rate and Ti4+ dissolution loss.
Processes 13 02981 g003
Figure 4. Variation of the adsorption capacity of Li+ with time (A), corresponding pseudo-first-order (B), pseudo-second-order kinetic model (C), Evolich kinetic model (D) and intra-particle diffusion model (E) (Adsorption conditions: initial Li+ concentration = 1000 mg/L, temperature = 25 °C, pH = 12).
Figure 4. Variation of the adsorption capacity of Li+ with time (A), corresponding pseudo-first-order (B), pseudo-second-order kinetic model (C), Evolich kinetic model (D) and intra-particle diffusion model (E) (Adsorption conditions: initial Li+ concentration = 1000 mg/L, temperature = 25 °C, pH = 12).
Processes 13 02981 g004
Figure 5. Fitting models of HTO adsorption in Li+ solutions at varying concentrations with Langmuir (A) and Freundlich (B) models (Conditions: contact time = 48 h, temperature = 25 °C, pH = 12).
Figure 5. Fitting models of HTO adsorption in Li+ solutions at varying concentrations with Langmuir (A) and Freundlich (B) models (Conditions: contact time = 48 h, temperature = 25 °C, pH = 12).
Processes 13 02981 g005
Figure 6. The recycling experiments of HTO (A) and the acid elution delithiation rate and titanium dissolution loss rate (B) (Adsorption conditions: adsorbent dosage = 0.5 g, contact time = 48 h, temperature = 25 °C, pH = 12).
Figure 6. The recycling experiments of HTO (A) and the acid elution delithiation rate and titanium dissolution loss rate (B) (Adsorption conditions: adsorbent dosage = 0.5 g, contact time = 48 h, temperature = 25 °C, pH = 12).
Processes 13 02981 g006
Figure 7. FT-IR spectra (A) and XRD patterns (B) of HTO before and after Li+ adsorption.
Figure 7. FT-IR spectra (A) and XRD patterns (B) of HTO before and after Li+ adsorption.
Processes 13 02981 g007
Table 1. Adsorption kinetic model rate constants (25 °C).
Table 1. Adsorption kinetic model rate constants (25 °C).
Adsorption
Kinetic
Model
Pseudo-First-Order Kinetic ModelPseudo-Second-Order Kinetic ModelEvolich Kinetic Model
k1/min−1R2k2/g‧(mg‧min)−1R2α/mg·(g·min)−1β/g·mg−1R2
Li+0.1850.9850.0460.9970.1218.880.980
Table 2. Rate constants for the intra-particle diffusion model.
Table 2. Rate constants for the intra-particle diffusion model.
ParametersValues
k1 (mg/(g·min0.5))12.46
C1 (mg/g)−5.62
R20.9953
k2 (mg/(g·min0.5))6.99
C2 (mg/g)3.61
R20.9497
k3 (mg/(g·min0.5))2.62
C3 (mg/g)19.61
R20.8158
Table 3. Adsorption isotherm model fitting parameters (25 °C).
Table 3. Adsorption isotherm model fitting parameters (25 °C).
Adsorption Isotherm Model ParametersLi+
Langmuirqm/mg·g−138.46
KL/L·mg−10.0021
R20.9741
FreundlichKF/(mg·g−1)(mg·L−1)−1/n0.2639
n1.42
R20.9858
Table 4. Comparison of the adsorption performance with other reported adsorbents.
Table 4. Comparison of the adsorption performance with other reported adsorbents.
AdsorbentC0 (mg/L)pHqe (mg/g)Adsorption TimeRef.
HTO-HEC-601001227.91[18]
HTO-170 °C-12 h1001230.95[18]
HTO nanosheets166.61318.754[32]
Yolk-shell structured HTO347.112.7320.106[17]
HTO-OS374.81318.814[20]
HTO-NS374.81328.894[20]
HTO10001238.4624Present study
Table 5. Adsorption selectivity of HTO in simulated oilfield produced water containing Li+, Na+, K+, Ca2+ and Mg2+ (Conditions: temperature = 25 °C, pH = 12).
Table 5. Adsorption selectivity of HTO in simulated oilfield produced water containing Li+, Na+, K+, Ca2+ and Mg2+ (Conditions: temperature = 25 °C, pH = 12).
Metal IonsLi+Na+K+Ca2+Mg2+
Ionic radius (pm)7610213810072
C0 (mg/L)186.3725,772.697389.414397.642981.92
Ce (mg/L)158.8325,706.897371.654389.972958.17
Kd (mL/g)173.39 2.562.411.758.03
α M L i 167.7471.9799.2421.60
Table 6. Elemental analysis of HTO and HTO-Li+ (at%) determined by XPS.
Table 6. Elemental analysis of HTO and HTO-Li+ (at%) determined by XPS.
Sample/ElementC1sO1sLi1sTi2pCl2p
HTO17.9752.1210.1619.350.39
HTO-Li+22.6543.8415.6313.30.57
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ma, W.; Huang, H.; Zhu, G.; Wang, X.; Kong, Q.; Shi, X. Fabrication of Spinel-Type H4Ti5O12 Ion Sieve for Lithium Recovery from Aqueous Resources: Adsorption Performance and Mechanism. Processes 2025, 13, 2981. https://doi.org/10.3390/pr13092981

AMA Style

Ma W, Huang H, Zhu G, Wang X, Kong Q, Shi X. Fabrication of Spinel-Type H4Ti5O12 Ion Sieve for Lithium Recovery from Aqueous Resources: Adsorption Performance and Mechanism. Processes. 2025; 13(9):2981. https://doi.org/10.3390/pr13092981

Chicago/Turabian Style

Ma, Weiwei, Hongrong Huang, Guangjin Zhu, Xueqing Wang, Qiaoping Kong, and Xueqing Shi. 2025. "Fabrication of Spinel-Type H4Ti5O12 Ion Sieve for Lithium Recovery from Aqueous Resources: Adsorption Performance and Mechanism" Processes 13, no. 9: 2981. https://doi.org/10.3390/pr13092981

APA Style

Ma, W., Huang, H., Zhu, G., Wang, X., Kong, Q., & Shi, X. (2025). Fabrication of Spinel-Type H4Ti5O12 Ion Sieve for Lithium Recovery from Aqueous Resources: Adsorption Performance and Mechanism. Processes, 13(9), 2981. https://doi.org/10.3390/pr13092981

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop