Original Paper

Journal of Korea TAPPI. 30 April 2025. 29-45
https://doi.org/10.7584/JKTAPPI.2025.4.57.2.29

ABSTRACT


MAIN

  • 1. Introduction

  • 2. Materials and Methods

  •   2.1 Materials

  •   2.2 Preparation of PAA-CA/PAA hydrogels

  •   2.3 Characterization

  •   2.4 Adsorption

  •   2.5 Reusability

  • 3. Results and Discussion

  •   3.1 Characterization and comparison of PAA and PAA-CA hydrogels

  •   3.2 REEs enrichment by absorbents

  •   3.3 Adsorption mechanism

  • 4. Conclusions

1. Introduction

Rare earth elements (REEs) refer to a group of 17 elements, comprising the 15 lanthanides along with scandium and yttrium. These elements are widely recognized for their exceptional physicochemical properties, including outstanding optical,1) electrical,2) and magnetic characteristics.3) Due to these unique attributes, REEs have become indispensable in a broad range of industries, including renewable energy,4) military technology, aerospace engineering, electronics, and advanced manufacturing.5) As the global demand for high-performance materials continues to grow, REEs are increasingly regarded as strategic resources, making their sustainable supply and efficient utilization a critical concern.6)

However, the extraction, refining, and subsequent treatment of REE-containing ores generate substantial amounts of wastewater.7) This wastewater is typically characterized by low REE concentrations and the presence of diverse and complex ion species, making the efficient separation and recovery of REEs highly challenging.8) The inefficiency of conventional recovery methods not only leads to significant resource loss but also raises serious environmental concerns due to the potential release of toxic heavy metals and other pollutants into aquatic ecosystems.

To address these challenges, a variety of separation techniques have been developed, including ion exchange, chemical precipitation, electrochemical approaches, solvent extraction, and adsorption.9) Among these, adsorption has garnered considerable attention due to its cost-effectiveness, ease of operation, and compatibility with large-scale industrial applications. Unlike solvent extraction and electrochemical methods, which often involve high energy consumption and the use of hazardous chemicals, adsorption-based techniques offer a more environmentally friendly approach to REE recovery. Nevertheless, existing adsorption strategies still face significant limitations, particularly in terms of selectivity, adsorption capacity, and material recyclability. These drawbacks undermine the urgent need for the development of novel adsorbents with enhanced performance.10)

Porous adsorbents, such as aerogels,11) cryogels,12) foams,13) and xerogels,14) have emerged as promising candidates for REE separation due to their large surface area, tunable porosity, and high adsorption efficiency. Surface modifications can further enhance the selectivity of these materials, enabling them to target specific REEs over competing species.15) However, a major challenge in the fabrication of these porous materials is the difficulty in the precise control of pore structure, which is largely dictated by preparation techniques such as freeze-drying and supercritical drying.16) These conventional methods often result in inconsistent pore architectures, affecting the overall performance of the adsorbents.

High internal phase emulsions (HIPEs) offer a versatile and innovative strategy for the fabrication of highly porous materials with well-defined structures. HIPEs are emulsions in which the dispersed phase exceeds 74% of the total volume, which can lead to the formation of interconnected porous networks upon polymerization or gelation.17,18) The unique morphology of HIPE-templated materials arises from the accumulation of large oil droplets, with smaller droplets filling the interstitial voids, thereby generating a highly porous structure with exceptional adsorption capabilities. This templating approach provides precise control over pore size distribution, mechanical stability, and surface chemistry, making HIPE-derived porous materials highly attractive for REEs adsorption.19)

In addition to high porosity, selective adsorption is a crucial requirement for efficient REE recovery. Selectivity is largely determined by the presence of functional groups that interact specifically with REE ions. Functional groups such as hydroxyl (-OH), carboxyl (-COOH), and phosphate (-PO43-) groups have been widely explored for their ability to coordinate with REEs through strong chelation interactions.20) Among these, carboxyl groups are particularly attractive due to their ease of functionalization, strong binding affinity for REEs, and compatibility with a wide range of polymeric materials, especially biopolymers.21)

Poly(acrylic acid) (PAA) has emerged as a promising candidate for REE adsorption due to its abundance of carboxyl (-COOH) groups, which serve as highly effective coordination sites for REE binding.22) However, despite its strong affinity for REEs, PAA-based adsorbents often suffer from limited adsorption efficiency and insufficient structural stability, necessitating further optimization. One potential strategy to enhance the adsorption performance of PAA-based materials is the incorporation of citric acid (CA) as a functional modifier. CA, with its multiple carboxyl functional groups, can significantly increase the density of active binding sites and improve the overall binding affinity for REEs. By introducing CA into PAA-functionalized porous hydrogels, the adsorption selectivity and capacity can be greatly enhanced, making these materials more effective for practical applications in REE recovery.

To achieve the goal of designing absorbents with high adsorption capacity, high selectivity, and good reusability, the HIPEs-templating method was integrated with the PAA-CA crosslinking approach. To test their feasibility, the adsorption properties were studied for light and heavy REEs, namely, Pr3+ and Er3+. Variables such as pH, contact time, initial concentration, and temperature, were investigated to optimize the adsorption process. Additionally, adsorption kinetics and thermodynamics were analyzed to understand the adsorption mechanism. The effect of competing ions on adsorption and the reusability of the hydrogel absorbents were also evaluated.

2. Materials and Methods

2.1 Materials

Acrylic acid (AA), CA, N, N'-methylenebisacrylamide (MBA), xanthan, scandium nitrate hydrate (Sc(NO3)3·xH2O, 99.99%), thulium nitrate pentahydrate (Tm(NO3)3·5H2O, 99.9%), lutetium nitrate hexahydrate (Lu(NO3)3·6H2O), gadolinium nitrate hexahydrate (Gd(NO3)3·6H2O), copper nitrate trihydrate (Cu(NO3)2·3H2O, 99.99%), zinc chloride (ZnCl2, 99.95%), magnesium chloride hexahydrate (MgCl2·6H2O, 99.99%), and 3Å molecular sieve were purchased from Aladdin Biochemical Technology Co., Ltd. (China). Gelatin (gel strength ~100 g Bloom), p-xylene (PX), ammonium persulfate (APS), tetramethylethylenediamine (TEMED), yttrium nitrate hexahydrate (Y(NO3)3·6H2O, 99.5%), lanthanum nitrate hexahydrate praseodymium (La(NO3)3·6H2O, 99%), cerium nitrate hexahydrate (Ce(NO3)3·6H2O, 99.99%), nitrate hexahydrate (Pr(NO3)3·6H2O, 99%), neodymium nitrate hexahydrate (Nd(NO3)3·6H2O, 99%), samarium nitrate hexahydrate (Sm(NO3)3·6H2O, 99.9%), europium nitrate hexahydrate (Eu(NO3)3·6H2O, 99.99%), dysprosium nitrate hexahydrate (Dy(NO3)3·6H2O, 99.9%), holmium nitrate pentahydrate (Ho(NO3)3·5H2O, 99.9%), erbium nitrate hexahydrate (Er(NO3)3·6H2O, 99.99%), ytterbium nitrate pentahydrate (Yb(NO3)3·5H2O, 99.9%), aluminum nitrate nonahydrate (Al(NO3)3·9H2O, 99%), and iron nitrate nonahydrate (Fe(NO3)3·9H2O, 99.9%) was purchased from Macklin Biochemical Co., Ltd. (China). Sodium hydroxide (NaOH, 96%), hydrogen chloride (HCl, 36–38%), potassium chloride (KCl, 99.5%), and anhydrous calcium chloride (CaCl2, 96%) was purchased from Sinopharm Chemical Reagent Co., Ltd. (China). All chemicals and reagents used in this study were analytical reagent grade. Ultrapure water was used throughout the experiments.

2.2 Preparation of PAA-CA/PAA hydrogels

The synthetic route of the PAA-CA hydrogel adsorbent is shown in Fig. 1. As seen in Fig. 1, an emulsion-templating method was used to achieve controllability for a porous absorbent. AA and CA were crosslinked by APS and MBA, and added to the continuous phase with the presence of gelatin and xanthan. Stable emulsions were obtained after homogenization. A porous absorbent, PAA-CA was finally achieved by curing. The reference absorbent, PAA was prepared in the same methods without the addition of CA. Gelatin (0.1 g) and xanthan (0.05 g) were added to 10 mL of water and heated at 40°C while stirring until fully dissolved. MBA (0.15 g), AA (1.5 mL), and CA (0.5 g) were then added and stirred for 1 h at room temperature. The mixture was homogenized (homogenizer LC-SFJ-10) at 10,000 rpm while adding PX (30 mL) dropwise until the formation of emulsions. APS and TEMED were added and mixed for 3 min. The emulsion was transferred to a sealed container and cured at 60°C for 24 h in an oven. PX and residual monomers were removed by Soxhlet extraction with ethanol for 24 h. The material was sliced, immersed in a 5 wt% NaOH ethanol-water solution (Vethanol/Vwater = 7/3) for 24 h, dehydrated with industrial alcohol, embedded in 3Å zeolite molecular sieves, soaked in anhydrous ethanol for 24 h, and dried under vacuum at 60°C for 12 h. The reference sample of PAA hydrogels were prepared in similar methods at the same MBA and AA dosage.

https://cdn.apub.kr/journalsite/sites/ktappi/2025-057-02/N0460570203/images/ktappi_2025_572_29_F1.jpg
Fig. 1.

Synthetic route of the PAA-CA hydrogel adsorbent.

2.3 Characterization

The morphologies of samples were characterized using scanning electron microscopy (SEM, ZEISS Gemini SEM 300). The content of metal ions was determined by an inductively coupled plasma emission spectrometer (ICP-OES, PerkinElmer Optima 5300 DV). Fourier transform infrared (FTIR) spectroscopy spectra were collected on a Bruker Tensor 27 FT-IR spectrometer using the KBr tableting method. The X-ray photoelectron spectroscopy (XPS) was recorded by an AXIS SUPRA+ instrument (Kratos Analytical, UK) with an Al X-ray source.

2.4 Adsorption

The stock solution with an initial concentration of 1,000 mg/L were prepared for the adsorption of Pr3+ (praseodymium) and Er3+(erbium) REEs. The stock solutions were then directly diluted to obtain the adsorption working solution. Additionally, a binary mixed solution containing Er3+/K+, Mg2+, Ca2+, Zn2+, Al3+, and Fe3+ was prepared and the ion initial concentration was 50 mg/L. All experiments were conducted in 20 mL glass bottles by contacting 10 mL of the metal ion solution with 10 mg of hydrogel absorbents in a thermostatic shaker set at a constant speed of 150 rpm and a temperature of 298 K for a specified duration. The reaction mixtures were then filtered using 0.45 µm hydrophilic microporous filter membranes (polyethersulfone) and analyzed for metal ion concentrations using ICP-OES. The adsorption performance of PAA/PAA-CA for both single rare earth ions and binary component mixtures was evaluated. The optimal pH, initial concentration, and contact time for the maximum removal of Pr3+ and Er3+ were determined. The initial pH was adjusted by the dropwise addition of 0.1 mol/L HCl or NaOH solution, and pH values ranging from 1 to 7 were tested to prevent metal ion precipitation during adsorption. The water absorption swelling rate of 10 mg PAA/PAA-CA hydrogel was tested at different pH levels (1–9) to identify the optimal pH for the working solution. For the kinetic, isotherm, and thermodynamic experiments, the procedure remained the same, except that the time (1 to 480 min), concentration (100 to 1,000 mg/L), and temperature (298–318 K) were varied.

The amount of Pr3+, Er3+ or metal ions adsorbed by PAA/PAA-CA hydrogels was calculated from the following Eq. [1].

[1]
Q=Vm(C0-Ce)

where Q is the adsorption capacity for ions at any time (mg/g), C0 and Ce is the initial and equilibrium concentrations of rare earth ions or metal ions in solution (mg/L), V is the volume of solution (mL), m is the mass of hydrogels adsorbent (mg).

The competitive adsorption of K+, Mg2+, Ca2+, Cu2+, Zn2+, Al3+, and Fe3+ for Er3+ was examined. The distribution ratio was used to assess the selectivity of the PAA and PAA-CA (Kd), and the selectivity factor (kEr/M) of Er3+ in comparison to other metal ions was determined using the following Eqs. [2] and [3].

[2]
Kd=VCem(C0-Ce)
[3]
kEr/M=Kd(Er)Kd(M)

where C0, Ce represent the initial and final metal ion concentrations (mg/L), m represents the polymer mass (g), and V represents the solution volume (L). Kd is the ratio of Er3+ to other metal ions in milliliters per kilogram (mL/g). PAA and PAA-CA both have a selectivity coefficient of k.

The swelling ratio was calculated according to the following Eq. [4].

[4]
S=Ms-MdMd

where S is the swelling ratio (g/g), Ms is the mass of the absorbent after swelling (g), Md is the mass of the absorbent after drying (g).

2.5 Reusability

The reusability of the porous PAA/PAA-CA hydrogels adsorbent was investigated by ten consecutive adsorption-desorption cycles. When the adsorption equilibrium was reached, the adsorbent was separated and desorbed with 0.1 mol/L HCl solution. Then the adsorbent was regenerated in 0.1 mol/L NaOH solution for 30 min, washed with distilled water several times, dried at 60°C in a vacuum oven, and reused in the next adsorption experiments. There is no difference in gel content between PAA-CA and PAA hydrogels during preparation and circulation experiments.

3. Results and Discussion

3.1 Characterization and comparison of PAA and PAA-CA hydrogels

To confirm the formation of PAA-CA from the emulsion template, FTIR spectra of PAA-CA, PAA, and CA were recorded. As shown in Fig. 2, the characteristic absorption bands of CA at 3,496 cm-1 (hydroxyl group), 1,748 and 1,705 cm-1 (symmetric carboxyl groups), and 1,128 cm-1 (stretching vibration of the tertiary alcohol hydroxyl group) were significantly weakened in PAA-CA. While, the infrared absorption peaks at 3,400 cm-1 (hydroxyl group), 2,941 cm-1 (asymmetric stretching vibration of methylene), and 1,574 and 1,421 cm-1 (stretching vibrations of carbon-oxygen double and single bonds) were noticeably enhanced in PAA-CA, indicating the grafting of CA onto PAA.23) Additionally, PAA-CA exhibited a new broad absorption band at 1,228–1,007 cm-1, assigning to the asymmetric stretching vibration of ester groups, further confirming the grafting between PAA and CA. The absence of a characteristic absorption band between 1,710 and 1,730 cm-1 attributed to the C=O stretching of carboxyl groups, suggests that the carboxyl group of AA has been completely neutralized.

https://cdn.apub.kr/journalsite/sites/ktappi/2025-057-02/N0460570203/images/ktappi_2025_572_29_F2.jpg
Fig. 2.

FTIR spectra of PAA-CA, PAA and CA.

Fig. 3 show the morphology and pore size distribution of PAA and PAA-CA hydrogels. PAA-CA has an increased number of smaller pores (average pore size of 8.87 ± 0.47 µm) than PAA (average pore size of 11.89 ± 0.53 µm). The increased number of smaller pores allows larger specific surface areas for adsorption, which may increase the adsorption capacity of REEs.24) Apart from the increased number of smaller pores, PAA-CA exhibits more uniform pore structure than PAA, which facilitates the structural integrity of absorbents both in dry (Fig. 3c) and wet (Fig. 3d, inset) conditions. The swelling ratios of absorbents were characterized at the pH range of 1–7. According to the results in Fig. 3d, the swelling ratios of PAA and PAA-CA increase with rising pH, which can be ascribed to the higher affinity of absorbents to H2O molecules in less protonated conditions. In general, PAA has a bit higher swelling ratio than PAA-CA, possibly due to the relatively structural nonuniformity of PAA.

https://cdn.apub.kr/journalsite/sites/ktappi/2025-057-02/N0460570203/images/ktappi_2025_572_29_F3.jpg
Fig. 3.

SEM images of the cross-section of PAA (a1, a2) and PAA-CA (b1, b2). (a2) and (b2): SEM images at a higher magnification, and corresponding pore size distributions. (c): Photo of PAA-CA absorbents. (d): Swelling ratio of PAA and PAA-CA at the pH values range of 1–7. (d): Insert a fully swelled PAA-CA hydrogel at pH = 7.

The absorbents shown in Fig. 3 were prepared at an oil phase volume fraction of 75%, which is HIPEs. In order to confirm the effect of surface area on adsorption capacity, PAA-CA absorbents with less specific surface area were also prepared by the same method at lower oil phase volume fractions for comparison. As shown in Fig. 4, the adsorption capacities of PAA-CA at the oil phase volume fractions of 50, 67, and 75% of Er3+ at room temperature are 243.4, 314.3, and 362 mg/g, respectively. Thus, the absorbents used in subsequent experiments were prepared with HIPEs (i.e. oil phase volume fraction of 75%).

https://cdn.apub.kr/journalsite/sites/ktappi/2025-057-02/N0460570203/images/ktappi_2025_572_29_F4.jpg
Fig. 4.

Er3+ adsorption capacity of PAA-CA prepared at different oil phase volume fractions. Adsorption conditions: C0 of 400 mg/L, time of 10 h, adsorption dosage of 10 mg/10 mL, pH = 5, and room temperature.

3.2 REEs enrichment by absorbents

3.2.1 Effect of pH value

To evaluate the adsorption capacity of REEs, PAA and PAA-CA were subjected to the adsorption tests of two REEs, including Pr3+ and Er3+, which belong to the light and heavy REEs, respectively. The effect of pH on the adsorption capacity of Pr3+and Er3+ by PAA and PAA-CA is presented in Fig. 5. The adsorption capacities of both hydrogels are similar at different pHs. Specifically, the adsorption capacity of Pr3+and Er3+ increases significantly as the pH increases from 1 to 3 and then stabilizes within the pH range of 3 to 7. This adsorption behavior is closely related to the interaction between the carboxyl groups (-COOH) of PAA and PAA-CA and the metal ions. In highly acidic conditions (pH from 1 to 3), most of the carboxylate ions (-COO) are protonated into carboxyl groups (-COOH), which decreases the number of available active sites to bind metal ions.25) Additionally, the excess H+ ions in the solution compete with Pr3+ and Er3+ for the ion-exchange sites on the adsorbent surfaces, further reducing the adsorption capacity.21) As the pH increases from 3 to 7, the carboxyl groups deprotonate, rendering more carboxylate ions (-COO) that can bind with REEs through electrostatic interactions and coordination. In addition, Er3+ consistently shows a higher adsorption capacity than Pr3+ in the pH range from 2 to 7, which can be ascribed to the influence of ion radii and coordination preferences. Typically, heavy REE ions exhibit a slightly smaller size than light REE ions which render their diffusion into the absorbents’ networks. The reduced atomic radius of heavy REEs tends to have a decreased coordination number, facilitating the bulk diffusion of heavy REEs onto chelating sites of absorbents. Although PAA-CA has larger surface area and more carboxyl groups than PAA, no apparent increase in the adsorption capacity can be observed in PAA-CA, which is possibly a result of the restricted diffusion of REEs into the hydrogel networks.

https://cdn.apub.kr/journalsite/sites/ktappi/2025-057-02/N0460570203/images/ktappi_2025_572_29_F5.jpg
Fig. 5.

Adsorption capacity of Pr3+ and Er3+ by PAA (a) and PAA-CA (b) at pH = 1–7. Adsorption conditions: C0 of 400 mg/L, adsorbent dosage of 10 mg/10 mL, time of 0–8 h, and temperature of 25°C.

3.2.2 Effect of initial REE concentration

Due to their strong complexing abilities, hydrogel-based adsorbents exhibit a strong ability to adsorb metal ions.26) To accurately determine the maximum adsorption capacity and facilitate comparisons with other reported adsorbents, a relatively wide concentration range (100 to 1,100 mg/L) was chosen to investigate the effect of the initial concentration (Fig. 6). Both hydrogel absorbents exhibit similar maximum adsorption capacities for Pr3+ and Er3+. At concentrations below 400 mg/L, the adsorption capacity increases rapidly due to the abundant adsorption sites. From 400 to 600 mg/L, the adsorption capacity rises slowly until equilibrium at concentrations above 600 mg/L, which is ascribed to the limited adsorption sites of hydrogel absorbents at higher concentrations. Comparing Fig. 6a and b, one can see that the hydrogel absorbents reach adsorption capacity for Pr3+ at a bit lower concentration (ca. 500 mg/L) than that of Pr3+ (600 mg/L). As a result of the larger ion radii of Pr3+, absorbents for Pr3+ saturate at lower initial concentrations and adsorption capacities.

https://cdn.apub.kr/journalsite/sites/ktappi/2025-057-02/N0460570203/images/ktappi_2025_572_29_F6.jpg
Fig. 6.

Adsorption capacity of Pr3+(a) and Er3+(b) at different initial concentrations. The inset displays the Langmuir isotherm model fitting results. Adsorption conditions: adsorbent dosage of 10 mg/10 mL, pH = 5, time of 0–8 h, and temperature of 25°C.

In adsorption studies, the Langmuir and Freundlich isotherm models are commonly used to analyze the adsorption mechanism. The Langmuir model assumes monolayer adsorption on localized sites without any interaction between adsorbed molecules, while the Freundlich model is used to describe adsorption on heterogeneous surfaces with varying energy distributions.27) The formula of these models are expressed the following Eqs. [5] and [6].

Langmuir isotherm model:

[5]
CeQe=1QmKL+CeQm

Freundlich isotherm model:

[6]
lnQe=lnKF+1nlnCe

where Ce (mg/L) and Qe (mg/g) represent the equilibrium concentration and adsorption capacity at equilibrium for a given initial concentration. Qm (mg/g) is the maximum adsorption capacity, while KL (L/mg) and KF (mg1−1/n L1/n/g) are the constants determined by linear regression of the experimental data for the Langmuir and Freundlich models, respectively. The constant n (dimensionless) describes the adsorption intensity. The estimated model parameters along with their correlation coefficients (R2) are presented in Table S1.

The adsorption capacity derived from the Langmuir isotherm model closely matched the experimental data, and the higher correlation coefficient (RL2 > 0.998) suggests that the Langmuir model is well-suited for describing the adsorption behavior. Note that the correlation coefficients of the Freundlich model are RF2 < 0.795. The key characteristic of the Langmuir isotherm can be represented by the dimensionless separation factor RL (RL = 1/(1 + KLC0), C0 as the initial concentration of Pr3+ and Er3+ in mg/L, and KL as the Langmuir constant. These factors are used to evaluate the preference of the adsorption process. Since the calculated RL values are all in the range of 0 to 1 (Table S2), the adsorption of Pr3+ and Er3+ onto the hydrogel absorbents is preferred. Additionally, the decrease of RL values with increasing initial concentration suggests that higher concentrations of Pr3+ and Er3+ enhance the adsorption process.25,28) This indicates that Pr3+ and Er3+ are adsorbed as a monolayer on the adsorbent surface.

3.2.3 Adsorption kinetics and thermodynamics

Contact time is a crucial parameter for evaluating the adsorption kinetics of an adsorbent. Fig. 7 shows the variation in the adsorption capacity of Pr3+ and Er3+ on PAA and PAA-CA as a function of contact time at three different temperatures, including 298, 308, and 318 K. The results show that the adsorption capacity of PAA reaches 90% within 240 min, while the adsorption capacity of PAA-CA reaches 90% in just 120 min. The adsorption capacities for Pr3+ and Er3+ increase with rising temperature, indicating that the adsorption processes are endothermic. This suggests that higher temperatures enhance the interaction between the metal ions and the polymeric adsorption sites, possibly by providing more kinetic energy to overcome activation energy barriers associated with ion binding.29) The influence of temperature is more evident on the adsorption capacity of PAA than that of PAA-CA, possibly due to the higher flexibility of hydrogel structure under increased temperature. To analyze the controlling factors of the adsorption process, the pseudo-first-order and pseudo-second-order kinetic models were applied,30) the Eqs. [7] and [8] are given below:

https://cdn.apub.kr/journalsite/sites/ktappi/2025-057-02/N0460570203/images/ktappi_2025_572_29_F7.jpg
Fig. 7.

Adsorption capacity of Pr3+and Er3+ by PAA (a, b) and PAA-CA (c, d) as a function of contact time at 298, 308, and 318 K. Adsorption conditions: C0 of 400 mg/L, adsorbent dosage of 10 mg/10 mL, pH = 5, and time of 0–8 h.

[7]
ln(Qe-Qt)=lnQe-K1t
[8]
tQt=1K2Qe2+tQe

where Qe and Qt are the adsorbed amount of REEs (mg/g) at equilibrium and the time t, respectively. K1 and K2 are the constants of the pseudo-first-order equation (min-1) and pseudo-second-order equation (g/mg·min), respectively. The related parameters and correlation coefficients R2 can be obtained by regressing the experimental data (Table S3). Both PAA and PAA-CA hydrogels have higher correlation coefficients of the pseudo-second-order model, indicating chemisorption was involved in the process, wherein ion exchange or complexation may play a dominant role.

To evaluate the thermodynamic parameters for the adsorption of Pr3+ and Er3+ by PAA and PAA-CA, following Eqs. [9] and [10] were utilized:

[9]
ΔG°=-RTlnKT
[10]
lnKT=ΔS°R-ΔH°RT

where KT represents the thermodynamic equilibrium constant (L/g), R is the universal gas constant (8.314 J/mol/K), T is the temperature (K), and ΔG°, ΔS°, ΔH° denote the Gibbs free energy (kJ/mol), entropy (J/mol/K), and enthalpy (kJ/mol), respectively. These thermodynamic parameters can be calculated using Eqs. [9] and [10] and are listed in Table 1. The negative values of ΔG° for Pr3+ and Er3+ indicate that the adsorption process is spontaneous. The positive ΔH° values suggest that the process is endothermic, and the magnitude of ΔH° (ΔH° > 0 kJ/mol) implies that chemisorption occurs. Additionally, the positive ΔS° values indicate an increase in disorder, likely due to the exchange of metal ions with more mobile ions in the exchanger during adsorption.

Table 1.

Thermodynamic parameters of Pr3+ and Er3+ adsorption on PAA and PAA-CA

Sample ΔS°
(kJ/mol)
ΔH°
(kJ/mol)
ΔG° (kJ/mol)
298 K 308 K 318 K
Pr3+ PAA 0.087 23.813 -2.280 -2.702 -4.040
PAA-CA 0.119 33.468 -1.549 -3.914 -3.871
Er3+ PAA 0.108 28.474 -3.530 -5.234 -5.662
PAA-CA 0.188 52.352 -3.139 -6.543 -6.826

3.2.4 Selectivity

REEs are often present alongside other common metal ions in environmental and mining wastes, the high selectivity of adsorbents toward rare earth ions is critical for practical applications. As both PAA and PAA-CA have higher adsorption capacity for Er3+, the selectivity for Er3+ of absorbents was studied only in the binary mixtures of Mn+/Er3+. Mn+ refers to common metal ions including K+, Mg2+, Ca2+, Cu2+, Zn2+, Al3+, and Fe3+. As illustrated in Fig. 8, PAA and PAA-CA exhibit a consistently higher adsorption capacity for Er3+ than other metal ions. According to Eqs. [9] and [10], one can obtain the selectivity factors in Table 2. Both PAA and PAA-CA have the highest selectivity in the binary mixtures of K+/Er3+, followed by Fe3+/Er3+ and M2+/Er3+, but exhibit little selectivity for Al3+/Er3+. This selectivity follows the general principle of selective absorbance for metal ions - the higher the valence of metal ions, the easier the absorbance. For comparison, PAA-CA exhibits evidently higher selectivity for all binary mixtures than PAA, suggesting its potential as a more effective adsorbent for selective separation of REEs.

https://cdn.apub.kr/journalsite/sites/ktappi/2025-057-02/N0460570203/images/ktappi_2025_572_29_F8.jpg
Fig. 8.

Adsorption capacity of PAA (a) and PAA-CA (b) in Mn+/Er3+ binary metal ion mixtures. Adsorption conditions: C0 of 50 mg/L, adsorbent dosage of 10 mg/10 mL, pH = 5, time of 240 min, and room temperature.

Table 2.

Selectivity of PAA and PAA-CA for Er3+ compared to other metal ions

Mn+/Er3+ PAA PAA-CA
Kd (Er3+) Kd (Mn+) kEr/MKd (Er3+) Kd (Mn+) kEr/M
K+ 51.9 1.8 28.8 388.6 1.6 242.88
Mg2+ 13.9 7.9 1.8 302.2 29.6 10.21
Ca2+ 57.1 36.3 1.6 312.4 21.4 14.60
Cu2+ 32.7 34.5 0.95 1,057.5 375.9 2.81
Zn2+ 94.2 73.3 1.29 207.6 35.1 5.91
Fe3+ 21.1 1.4 15.07 138.7 3.2 42.71
Al3+ 2.4 2.4 1.00 2.4 2.4 1.00

3.2.5 Reusability

The reusability of an adsorbent is a crucial factor in determining its practical application. To assess the reusability of PAA and PAA-CA, ten consecutive adsorption-desorption cycles were conducted. The relationship between adsorption capacity and cycle number is illustrated in Fig. 9. It was observed that both adsorbents maintained high and stable adsorption capacities across all 10 cycles, which is not commonly seen in other biobased absorbents. This suggests that the adsorption of REEs onto PAA and PAA-CA is reversible in the presence of H+, facilitated by ion exchange. Besides, the maintained recyclability of absorbents can be ascribed to their hydrogel state which endures the shape recovery of absorbents in wet conditions. These hydrogel absorbents consistently demonstrated a higher adsorption capacity of Er3+ than Pr3+. These findings indicate that both PAA and PAA-CA exhibit excellent reusability for the adsorption of REEs.

https://cdn.apub.kr/journalsite/sites/ktappi/2025-057-02/N0460570203/images/ktappi_2025_572_29_F9.jpg
Fig. 9.

Adsorbed amount of Pr3+ and Er3+ on PAA (a) and PAA-CA (b) as a function of adsorption-desorption cycle. Adsorption conditions: C0 of 400 mg/L, adsorbent dosage of 10 mg/10 mL, pH = 5, time of 240 min, and room temperature.

3.3 Adsorption mechanism

To analyze the adsorption mechanism, the FTIR (Fig. 10) and XPS (Figs. 11 and 12) spectra of PAA and PAA-CA before and after the adsorption of Pr3+ and Er3+ were recorded. The FTIR spectra underline the changes in the functional groups before and after adsorption. It is obvious that the absorption band of the hybrid hydrogel adsorbent at 3,424 cm-1 shifts to lower wavenumber (3,385 cm-1) after adsorption of Pr3+ and Er3+, which is attributed to the interaction -OH groups with Pr3+ or Er3+. The asymmetric stretching vibration of -COO at 1,575 cm-1 shifts to 1,545 cm-1 after the adsorption of Pr3+ and Er3+. The absorption band at 1,419 cm-1 derived from the symmetric stretching vibration of carboxylate reduces and the bending vibration C-H at 1,457 cm-1 mergences with the neighouring adsorption peak.26,31) The results above listed indicate that the carboxylate groups chelated with Pr3+ and Er3+ during the adsorption process. PAA-CA exhibits a similar C=O stretching band, but at different in intensity and position compared to PAA. Notably, the absorption bands around 1,417 and 1,451 cm-1 show significant shifts following ion adsorption, indicating enhanced interaction between the modified polymer and the metal ions. These shifts in the characteristic absorption bands demonstrate that metal ion coordination primarily occurs through the carboxyl groups.32)

https://cdn.apub.kr/journalsite/sites/ktappi/2025-057-02/N0460570203/images/ktappi_2025_572_29_F10.jpg
Fig. 10.

FTIR spectra of PAA (a) and PAA-CA (b) before and after adsorption of Pr3+ and Er3+.

The XPS survey spectra of PAA and PAA-CA after the adsorption of Pr3+ and Er3+ are presented in Figs. 11 and 12. It is apparent in Fig. 11 that after adsorption, new peaks related to REEs at 217 and 169 eV appear, respectively, which confirms the successful adsorption by PAA and PAA-CA. The binding energy shifts of C1s and O1s spectra (Fig. 12) indicate that the oxygen atoms in the carboxyl groups serve as the primary adsorption sites for the REEs. After the adsorption of Pr3+ and Er3+ by PAA and PAA-CA materials, the binding energy of the C=O bond shifts to a higher energy level, indicating a decrease in electron cloud density. The larger shift of C=O in PAA-CA than in PAA can be correlated with the higher selectivity of PAA-CA than PAA. Besides, this shift can be attributed to the high charge density of Pr3+ and Er3+, which strongly coordinate with the carboxyl groups of polyacrylic acid, forming stable complexes. This coordination enhances the polarity of the C=O bond, redistributing its electron density and increasing its binding energy. Additionally, the interaction between rare earth ions and carboxyl groups withdraws electrons from the carboxyl oxygen, leading to a reduction in electron density, a shortened C=O bond length, and a higher binding energy. Furthermore, prior to adsorption, some carboxyl groups may have formed hydrogen bonds, which could have lowered the initial C=O binding energy. Upon adsorption of Pr3+ and Er3+, the deprotonation of carboxyl groups (COOH → COO) further enhances the electron, resulting in a stronger C=O bond and a further increase in binding energy. The shifts in the N1s spectra of C-NH2 groups indicate that C-NH2 groups also participated in REEs adsorption.33)

https://cdn.apub.kr/journalsite/sites/ktappi/2025-057-02/N0460570203/images/ktappi_2025_572_29_F11.jpg
Fig. 11.

XPS spectra of PAA (a) and PAA-CA (b) before and after adsorption of Pr3+ and Er3+.

https://cdn.apub.kr/journalsite/sites/ktappi/2025-057-02/N0460570203/images/ktappi_2025_572_29_F12.jpg
Fig. 12.

XPS analysis of PAA and PAA-CA before and after the adsorption of Pr3+ and Er3+.

4. Conclusions

Porous PAA and PAA-CA hydrogels were successfully synthesized using a gelatin-based HIPE-templating method. Their adsorption properties were tested both for heavy and light REEs. The high adsorption capacity of 400 mg/g is achieved both in PAA and PAA-CA in the pH value range of 3–7. The high adsorption capacity of PAA and PAA-CA hydrogel can be ascribed to the large surface area resulting from the HIPE template, and the carboxyl groups on material surfaces. The initial concentration of REEs Pr3+ and Er3+ have an important influence on the adsorption capacity of both PAA and PAA-CA hydrogels. Adsorption kinetics followed a pseudo-second-order model, and the process is spontaneous and endothermic. PAA-CA hydrogel reaches the adsorption capacity faster than PAA, indicating the efficiency of PAA-CA. PAA-CA hydrogel shows a higher affinity of rare earth ions than common metal ions, particularly for Er3+. Additionally, the prepared hydrogel absorbents maintained high adsorption capacity over 10 adsorption-desorption cycles, demonstrating excellent reusability. The selective adsorption performance of PAA-CA hydrogel is better than that of PAA hydrogel. The porous PAA-CA hydrogel is a promising candidate for the selective recovery of REEs from aqueous solutions with stable reproducibility.

Acknowledgements

This work is supported by the Ministry of Science and Technology of China (No. 2021YFC2901500), the National Natural Science Foundation of China (22108280), the Jinan Science and Technology Bureau project (Grant No. 20233046), Special funds for Taishan Industrial Project, Shandong Provincial Natural Science Foundation (No. ZR2024MC189) and Shandong Province Higher Education Youth Innovation and Technology Support Program (2024KJH047).

References

1

Sakthivel, P., Asaithambi, S., Karuppaiah, M., Sheikfareed, S., Yuvakkumar, R., and Ravi, G., Different rare earth (Sm, La, Nd) doped magnetron sputtered CdO thin films for optoelectronic applications, Journal of Materials Science: Materials in Electronics 30(10):9999-10012 (2019).

10.1007/s10854-019-01342-9
2

Liang, S., Wang, H., Li, Y., Qin, H., Luo, Z., Huang, B., Zhao, X., Zhao, C., and Chen, L., Rare-earth based nanomaterials and their composites as electrode materials for high performance supercapacitors: A review, Sustainable Energy & Fuels 4(8):3825-3847 (2020).

10.1039/D0SE00669F
3

Hono, K. and Sepehri-Amin, H., Prospect for HRE-free high coercivity Nd-Fe-B permanent magnets, Scripta Materialia 151:6-13 (2018).

10.1016/j.scriptamat.2018.03.012
4

Deng, X. and Ge, J., Global wind power development leads to high demand for neodymium praseodymium (NdPr): A scenario analysis based on market and technology development from 2019 to 2040, Journal of Cleaner Production 277, 123299 (2020),

10.1016/j.jclepro.2020.123299
5

Fabrizio, O., Rare earth starting materials and methodologies for synthetic chemistry, Chemical Reviews 122(6):6040-6116 (2022).

10.1021/acs.chemrev.1c0084235099940PMC9007467
6

Cao, Y., Shao, P., Chen, Y., Zhou, X., Yang, L., Shi, H., Yu, K., Luo, X., and Luo, X., A critical review of the recovery of rare earth elements from wastewater by algae for resources recycling technologies, Resources, Conservation & Recycling 169, 105519 (2021).

10.1016/j.resconrec.2021.105519
7

Meng, X., Zhao, H., Zhao, Y., Shen, L., Gu, G., and Qiu, G., Heap leaching of ion adsorption rare earth ores and REEs recovery from leachate with lixiviant regeneration, Science of The Total Environment 898:165417-165417 (2023).

10.1016/j.scitotenv.2023.16541737429479
8

Hu, Y., Florek, J., Larivière, D., Fontaine, F., and Kleitz, F., Recent advances in the separation of rare earth elements using mesoporous hybrid materials, Chemical Record (New York, NY), 18(7-8):1261-1276 (2018).

10.1002/tcr.20180001229806123PMC6147058
9

Traore, M., Gong, A., Wang, Y., Qiu, L., Bai, Y., Zhao, W., Liu, Y., Chen, Y., Liu, Y., Wu, H., Li, S., and You, Y., Research progress of rare earth separation methods and technologies, Journal of Rare Earths 41(2):182-189 (2023).

10.1016/j.jre.2022.04.009
10

He, Q., Qiu, J., Chen, J., Zan, M., and Xiao, Y., Progress in green and efficient enrichment of rare earth from leaching liquor of ion adsorption type rare earth ores, Journal of Rare Earths 40(03):353-364 (2022).

10.1016/j.jre.2021.09.011
11

Jeong, I. H., Kyu, Y. L., and Myoung, J. W. Characteristics of aerogels manufacturing using wastepaper, Journal of Korea Technical Association of the Pulp and Paper Industry 51(1):100-106 (2019).

10.7584/JKTAPPI.2019.02.51.1.100
12

Liu, L., Bai, L., Tripathi, A., Yu, J., Wang, Z., Borghei, M., Fan, Y., and Rojas, O. J., High axial ratio nanochitins for ultrastrong and shape-recoverable hydrogels and cryogels via ice templating, ACS Nano 13(3):2927-2935 (2019).

10.1021/acsnano.8b0723530689367PMC6439435
13

Yu, H., Zhu, Y., Xu, J., and Wang, A., Fabrication porous adsorbents templated from modified sepiolite-stabilized aqueous foams for high-efficient removal of cationic dyes, Chemosphere 259, 126949 (2020).

10.1016/j.chemosphere.2020.12694932634719
14

Bao, L., Zhu, X., Dai, H., Tao, Y., Zhou, X., Liu, W., and Kong, Y., Synthesis of porous starch xerogels modified with mercaptosuccinic acid to remove hazardous gardenia yellow, International Journal of Biological Macromolecules 89:389-395 (2016).

10.1016/j.ijbiomac.2016.05.00327151673
15

Peng, X., Mo, S., Li, R., Li, J., Tian, C., Liu, W., and Wang, Y., Effective removal of the rare earth element dysprosium from wastewater with polyurethane sponge-supported graphene oxide-titanium phosphate Environmental Chemistry Letters 19:719-728 (2021).

10.1007/s10311-020-01073-y
16

Chen, G., Yu, S., Ni, J. P., Yang, X. B., Liu, Y. Z., and Chen, T., Investigation of porous structure of aerogel prepared from nanofibrillated cellulose, Journal of Korea Technical Association of the Pulp and Paper Industry 48(6):17-24 (2016).

10.7584/JKTAPPI.2016.12.48.6.17
17

Wu, B., Yang, C., Xin, Q., Kong, L., Eggersdorfer, M., Ruan, J., Zhao, P., Shan, J., Liu, K., Chen, D., Weitz, D. A., and Gao, X., Attractive pickering emulsion gels, Advanced Materials 33(33):2102362 (2021).

10.1002/adma.20210236234242431
18

Lee, M. H., Kim, K. J., and Eom, T. J., Preparation of emulsion from biodegradable polymer(II) - Characteristics of paper treated as PLA and PBS emulsion, Journal of Korea Technical Association of the Pulp and Paper Industry 45(2):13-20 (2013).

10.7584/ktappi.2013.45.2.013
19

Wang, F., Zhu, Y. F., and Wang, A. Q., Preparation of Carboxymethyl cellulose-g-poly(acrylamide)/attapulgite porous monolith with an eco-friendly pickering-MIPE template for Ce(III) and Gd(III) adsorption, Frontiers in Chemistry 8:398 (2020).

10.3389/fchem.2020.0039832528928PMC7262556
20

Rinez, T., Tuomo, N., Petri, T., Juha, M., Jouko, V., Vesa-Pekka, L., and Joakim, R., Bisphosphonate modified mesoporous silicon for scandium adsorption, Microporous and Mesoporous Materials 296:109980 (2020).

10.1016/j.micromeso.2019.109980
21

Hwang, K. J., Kwon, G. J., Yang, J. W., Hwang, W. J., Hwang, J. H., and Kim, D. Y., Cadmium adsorption characteristic of cellulose-gel manufacture using alkali solvent, Journal of Korea Technical Association of the Pulp and Paper Industry 47(6):113-122 (2015).

10.7584/ktappi.2015.47.6.113
22

Ayesha, K., Poly(acrylic acid) nanocomposites: Design of advanced materials, Journal of Plastic Film & Sheeting 37(4):409-428 (2021).

10.1177/8756087920981615
23

Li, C., Ma, H. Y., Venkateswaran, S., and Hsiao, B. S., Sustainable carboxylated cellulose filters for efficient removal and recovery of lanthanum, Environmental Research 188:109685 (2020).

10.1016/j.envres.2020.10968532512372
24

Olena, A., Freire, R. S. D., and Volodymyr, Z., Recent advances in functional materials for rare earth recovery: A review, Sustainable Materials and Technologies 37, e00681 (2023).

10.1016/j.susmat.2023.e00681
25

Zhu, Y. F., Wang, W. B., Zheng, Y. A., Wang, F., and Wang, A. Q., Rapid enrichment of rare-earth metals by carboxymethyl cellulose-based open-cellular hydrogel adsorbent from HIPEs template, Carbohydrate Polymers 140:51-58 (2016).

10.1016/j.carbpol.2015.12.00326876827
26

Yan, H., Dai, J., Yang, Z., Yang, H., and Cheng, R. S., Enhanced and selective adsorption of copper(II) ions on surface carboxymethylated chitosan hydrogel beads, Chemical Engineering Journal 174(2):586-594 (2011).

10.1016/j.cej.2011.09.064
27

Wang, X., Sun, R., and Wang, C., pH dependence and thermodynamics of Hg(II) adsorption onto chitosan-poly(vinyl alcohol) hydrogel adsorbent, Colloids and Surfaces A: Physicochemical and Engineering Aspects 441:51-58 (2014).

10.1016/j.colsurfa.2013.08.068
28

Jing, H. J., Huang, X., Du, X. J., Mo, L., Ma, C. Y., and Wang, H. X., Facile synthesis of pH-responsive sodium alginate/carboxymethyl chitosan hydrogel beads promoted by hydrogen bond, Carbohydrate Polymers 278:118993 (2022).

10.1016/j.carbpol.2021.11899334973796
29

Ni, C. Q., Liu, Q. M., Ren, Z., Hu, H. Q., Sun, B. H., Liu, C., Shao, P. H., Yang, L. M., Pavlostathis, S. G., and Luo X. B., Selective removal and recovery of La(III) using a phosphonic-based ion imprinted polymer: Adsorption performance, regeneration, and mechanism, Journal of Environmental Chemical Engineering 9(6):106701 (2021).

10.1016/j.jece.2021.106701
30

Wu, Y. H., Pang, H. W., Yao, W., Wang, X. X., Yu, S. J., Yu, Z. M., and Wang, X. K., Synthesis of rod-like metal-organic framework (MOF-5) nanomaterial for efficient removal of U(VI): Batch experiments and spectroscopy study, Science Bulletin 63(13):831-839 (2018).

10.1016/j.scib.2018.05.02136658962
31

Wamea, P., Pitcher, M. L., Muthami, J., and Sheikhi, A., Nanoengineering cellulose for the selective removal of neodymium: Towards sustainable rare earth element recovery, Chemical Engineering Journal 428:131086 (2022).

10.1016/j.cej.2021.131086
32

Wang, F., Wang, W. B., Zhu, Y. F., and Wang, A. Q., Evaluation of Ce(III) and Gd(III) adsorption from aqueous solution using CTS-g-(AA-co-SS)/ISC hybrid hydrogel adsorbent, Journal of Rare Earths 35(7):697-708 (2017).

10.1016/S1002-0721(17)60966-9
33

Zhong, J., Yuan, X., Xiong, J., Wu, X. L., and Lou, W. Y., Solvent-dependent strategy to construct mesoporous Zr-based metal-organic frameworks for high-efficient adsorption of tetracycline, Environmental Research 226:115633 (2023).

10.1016/j.envres.2023.11563336931373

[Appendix] Supplementary Information

Table S1.

The constant parameters and correlation coefficients of Langmuir model and Freundlich model for Pr3+ and Er3+ adsorption onto the adsorbent

RE3+ Sample qe, expa
(mg/g)
Langmuir Freundlich
Pr3+ PAA 368.45 Qmb(mg/g) 347.22 KF 120.05
KL(L/mg) 0.13 n 5.45
RL2 0.9981 RF2 0.7387
PAA-CA 353 Qmb(mg/g) 357.14 KF 93.75
KL(L/mg) 0.11 n 4.48
RL2 0.9998 RF2 0.7947
Er3+ PAA 424.6 Qmb(mg/g) 448.43 KF 139.93
KL(L/mg) 0.10 n 5.28
RL2 0.9993 RF2 0.7198
PAA-CA 435 Qmb(mg/g) 452.49 KF 78.35
KL(L/mg) 0.04 n 3.44
RL2 0.9984 RF2 0.6557

a is the equilibrium adsorption capacity obtained from the adsorption experiment.

b is the equilibrium adsorption capacity calculated from the isotherm model.

Table S2.

The separation factor RL after adsorption of Pr3+ and Er3+ was calculated by Langmuir model fitting data

RE3+ Sample RL
Pr3+ PAA 0.0177
PAA-CA 0.0208
Er3+ PAA 0.0218
PAA-CA 0.0528
Table S3.

Estimated adsorption kinetic parameters for Pr3+ and Er3+ by pseudo-first-order and pseudo-second-order model

RE3+ Sample T (K) qe, expa
(mg/g)
Pseudo-first-order Pseudo-second-order
qe, calb
(mg/g)
k1
(min-1)
R12qe, calb
(mg/g)
k2
(g/mg·min)
R22
Pr3+ PAA 298.15 305.3 292.95 1.18 × 10-2 0.9499 347.22 4.10 × 10-5 0.9876
308.15 316.7 301.38 1.76 × 10-2 0.8861 347.22 6.53 × 10-5 0.9857
318.15 350.84 320.83 2.02 × 10-2 0.8543 378.79 6.26 × 10-5 0.9891
PAA-CA 298.15 330.4 308.73 2.13 × 10-2 0.9683 358.42 6.50 × 10-5 0.9953
308.15 350.85 300.12 2.86 × 10-2 0.9188 366.30 7.07 × 10-5 0.9876
318.15 346.74 308.85 2.73 × 10-2 0.9030 367.65 7.84 × 10-5 0.9955
Er3+ PAA 298.15 361.79 332.05 1.40 × 10-2 0.9543 403.23 3.75 × 10-5 0.9923
308.15 397.39 375.36 1.37 × 10-2 0.9186 442.48 3.81 × 10-5 0.9838
318.15 401.67 391.31 1.95 × 10-2 0.9915 438.60 5.81 × 10-5 0.9987
PAA-CA 298.15 350.18 338.78 1.93 × 10-2 0.9821 383.14 6.33 × 10-5 0.9972
308.15 416.51 380.28 1.91 × 10-2 0.9579 448.43 4.74 × 10-5 0.9939
318.15 417.3 360.42 2.63 × 10-2 0.9008 442.48 5.52 × 10-5 0.9897

a is the equilibrium adsorption capacity obtained from the adsorption experiment.

b is the equilibrium adsorption capacity calculated from the pseudo-first-order model or pseudo-second-order model.

페이지 상단으로 이동하기