Next Article in Journal
Gold Nanofilm-Coated Porous Silicon as Surface-Enhanced Raman Scattering Substrate
Next Article in Special Issue
Adsorption of Ammonium Nitrogen from Aqueous Solution on Chemically Activated Biochar Prepared from Sorghum Distillers Grain
Previous Article in Journal
Special Features on Intelligent Imaging and Analysis
Previous Article in Special Issue
Assessment of an MnCe-GAC Treatment Process for Tetramethylammonium-Contaminated Wastewater from Optoelectronic Industries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Adsorption and Desorption Mechanisms of Rare Earth Elements (REEs) by Layered Double Hydroxide (LDH) Modified with Chelating Agents

1
Jiangxi Province Key Laboratory of Polymer Micro/Nano Manufacturing and Devices, East China University of Technology, Nanchang 330013, China
2
Graduate School of Science and Technology, Niigata University, 8050 Ikarashi 2-Nocho, Nishi-ku, Niigata 950-2181, Japan
3
Department of Chemistry and Chemical Engineering, Faculty of Engineering, Niigata University, 8050 Ikarashi 2-Nocho, Nishi-ku, Niigata 950-2181, Japan
4
Department of Chemical System Engineering, Graduate School of Engineering, The University of Tokyo, Tokyo 113-8656, Japan
5
Center for Computational Sciences, Tsukuba University, Tsukuba 305-8577, Japan
6
Department of Materials Science, Graduate School of Engineering Science, Akita University, Akita 010-8502, Japan
*
Author to whom correspondence should be addressed.
Appl. Sci. 2019, 9(22), 4805; https://doi.org/10.3390/app9224805
Submission received: 10 September 2019 / Revised: 3 November 2019 / Accepted: 5 November 2019 / Published: 10 November 2019
(This article belongs to the Special Issue Advanced or Conventional Materials as Sorbent)

Abstract

:
In order to obtain the adsorption mechanism and adsorption structures of Rare Earth Elements (REEs) ions adsorbed onto layered double hydroxides (LDH), the adsorption performance of LDH and ethylenediaminetetraacetic acid (EDTA) intercalated LDH for REEs was investigated by batch experiments and regeneration studies. In addition to adsorption capacity, the partition coefficient (PC) was also evaluated to assess their true performance metrics. The adsorption capacity of LDH increases from 24.9 μg·g−1 to 145 μg·g−1 for Eu, and from 20.8 μg·g−1 to 124 μg·g−1 for La by intercalating EDTA in this work; and PC increases from 45.5 μg·g−1·uM−1 to 834 μg·g−1·uM−1 for Eu, and from 33.6 μg·g−1·μM−1 to 405 μg·g−1·μM−1 for La. Comparison of the data indicates that the adsorption affinity of EDTA-intercalated LDH is better than that of precursor LDH no matter whether the concept of adsorption capacity or that of the PC was used. The prepared adsorbent was characterized by XRD, SEM-EDS and FT-IR techniques. Moreover, quantum chemistry calculations were also performed using the GAUSSIAN09 program package. In this calculation, the molecular locally stable state structures were optimized by density functional theory (DFT). Both the quantum chemistry calculations and the experimental data showed that REEs ions adsorbed by EDTA-intercalated LDH are more stable than those adsorbed by precursor LDH. Furthermore, the calculation results of adsorption and desorption rates show that adsorption rates are larger for Eu(III) than for La(III), which agrees with the experimental result that Eu(III) has a higher adsorption ability under the same conditions. The LDHs synthesized in this work have a high affinity for removing REEs ions.

1. Introduction

Rare Earth Elements (REEs) are the collection of 17 special elements which are usually divided into heavy rare earth elements (HREE) and light rare earth elements (LREE). REEs could be widely used in many fields based on their excellent properties. For instance, the chemical, electrical, magnetic, and optical characteristics make them suitable for fields, including the nuclear industry, metallurgy, and luminescence [1]. Moreover, REEs can be combined with other element or compounds and then can be used as alloys, catalysts, ceramics, etc. Therefore, these strategic usages of REEs have been attracting much attention in recent years. However, due to the limited availability, the shortage of REEs resources and the necessity of a stable supply of these metals have been concerned over the past years. Therefore, the development of a recovery method for these trace metals is important from the viewpoint of environmental protection [2,3]. Metallic ions in aqueous solution can interact in various ways with a solid surface, including precipitation, adsorption, and ion exchange. Among these methods, adsorption has been regarded as an efficient way to remove metallic ion in aqueous solution [4,5,6]. For ecofriendly purposes, biological materials as an adsorbent for water purification have used by a number of researchers. The biological adsorbent has the advantages of recyclable, low cost, easy operation, and little possibility of secondary pollution.
From above-mentioned, the laboratory model experiments for biosorption of REEs using seaweed biomasses [7] and shell biomass [8] have been carried out in our previous work. It was found that seaweed and shell biomass show excellent sorption capacity for lanthanides even under the presence of common cations (refer to Figure S1). In addition, surface modification of activated carbon has been recognized as an attractive approach to improve and/or enhance the removal of heavy metals [9] or lanthanides [10].
On the other hand, layered double hydroxides (LDH) are a kind of lamellar ionic compound containing positively charged brucite-like layers and exchangeable anions in the interlayer. Their abundance in nature, low cost, and high specific surface area make them suitable for adsorption of metal ions. The general formula of an LDH is [ M 1 x II M x III ( OH ) 2 ] ( A n ) x n · mH 2 O , where cationic MII and MIII are divalent and trivalent metals, and occupy the center of octahedral in the brucite-like layer. An is the interlayer anion of charge n that leads to the electric neutrality of the LDH, and the coefficient x is the molar ratio of MII/(MII + MIII). The structure of LDH suggests that these materials can be intercalated easily with other anions. Various types of anions can be intercalated in its interlayer space. It has been extensively reported by both inorganic [11,12] and organic [13,14] chemistry researchers. The ability of LDH to intercalate anions makes them useful as catalysts, anion exchangers, and adsorbents. We note that several physicochemical models were proposed to rationalize the anion-exchange constants [15]. Some formalisms were derived by relating the activity of an anion to the anion-exchange constant [16], while geometric features of atoms, such as charge radii of the anions were taken into account to investigate the anion-exchange ability of halide anions on LDH [17].
Moreover, LDHs modified with hybrids have also been studied as potential adsorbents of heavy metals from aqueous solution [18,19,20]. Chelation is a type of bonding of ions or molecules to metallic ions. It involves the formation of coordinate bonds between a single central atom and ligand within multiple separate binding sites. Therefore, it can be expected that chelating agents modified LDH have higher adsorption capacity and better selectivity compared with other adsorbents for metallic ions. In our previous work [21], we synthesized and characterized LDH intercalated with the chelating agent ethylenediaminetetraacetic acid (EDTA) or N,N′-1,2-ethanediylbis-1-aspartic acid (EDDS) and studied the uptake of Cd(II), Cu(II), and Pb(II) by these hybrid compounds. It turned out that the LDHs synthesized in that work were very effective for removing heavy metal ions from aqueous solutions. In fact, in our another previous work [22], to theoretically clarify the adsorption mechanism and adsorption structures of LDH, we have performed quantum chemistry calculations of reactants, locally stable states, transition states, and products among phosphorous anion, water and hydrotalcite in a variety of pH ranges.
As mentioned above, a number of works have been devoted to researches on the adsorption phenomena of lanthanide [4,5,6,23,24,25,26]. These works were mainly focused on the potential application of materials for removal of lanthanides from wastewater. On the other hand, the objectives of our present work are to experimentally investigate the efficiency of LDHs for adsorbing REEs from aqueous solution, as well as to unveil the adsorption mechanism and adsorption structures of lanthanide adsorbed onto LDH using quantum chemistry calculations. We carried out the uptake experiments for the adsorption of REEs (lanthanum, europium) from aqueous solutions by LDH modified with EDTA. In addition, we performed quantum chemistry calculations, including locally stable states of LDH-EDTA and optimized structures of the chemical systems. Particular emphasis is put on the elucidation of the effect of EDTA on structural stability. In the calculations, we used molecular cluster models for hydrotalcite lamellae, whose chemical structure is described by Mg12Al7(OH)369+ [22,27]. Finally, to get a deeper insight into the uptake mechanism, we calculated the desorption rates of the adsorbates.

2. Materials and Methods

Chemical reagents, including Zn(NO3)2·6H2O, Al(NO3)3·9H2O, NaOH, NaNO3 and C10H14N2Na2O8 (Na2H2EDTA·2H2O) were purchased from Kanto Chemical Co., Inc. (Tokyo, Japan). REEs nitrate salts Eu(NO3)3·6H2O (99%, 446.07) and La(NO3)3·6H2O (97%, 433.01) were purchased from Kishida Chemical Co., Inc.(Osaka, Japan). All reagents used were of analytical grade. The La(III) or Eu(III) ion stock solutions were prepared and suitably diluted with ultrapure water. The ultrapure water (>18.2 MΩ) which was treated by an ultrapure water system (Suite A Dublin, California, USA, Advantec aquarius: RFU 424TA) was employed throughout the work.

2.1. Synthesis of the Adsorbent

In this section, we present the synthesis processes of L1 and L2. The synthesis of LDHs intercalated with EDTA includes two steps: The preparation of the precursor NO3-LDH (L1), and the ion-exchange reaction between NO3-LDH and Ethylenediaminetetraacetic acid disodium salt.

2.1.1. Synthesis of Precursor L1

L1 was prepared by dropping addition of 100 mL aqueous solution of 0.02 mol Zn(NO3)2·6H2O and 0.01 mol Al(NO3)3·9H2O to 100 mL NaOH/NaNO3 solution (pH 10, molar ratio 1:1). Then, the solutions were agitated at 70 °C for 8 h by maintaining the pH, separated by centrifugation, and washed until they became neutral.

2.1.2. Synthesis of L2

L2 was synthesized as follows. Under an N2 atmosphere, 0.015 mol of EDTA was added to the 150 mL of the suspended solution of L1. Then, the mixing solutions were agitated at 70 °C for 8 h under a certain pH degree, separated by centrifugation, washed until they became neutral, and then dried at 60 °C overnight.
The above synthesis processes were conducted with N2 purging to avoid CO2 uptake from the atmosphere. The preparation and purification of the adsorbent, and the characterization of the samples were reported in detail in our previous paper [21].

2.2. Characterization of the Adsorbent

Infrared spectra were obtained using the KBr disc method, with wavenumbers from 400 to 4000 cm−1 on an FT-IR (JASCO, Tokyo, Japan: FTIR-4200). The XRD pattern of LDHs samples was carried out on a RINT2500HR-PC (RIGAKU Corporation, Tokyo, Japan) using Cu Kα radiation in the scanning range of 2–80°. A Nova Nano SEM450 scanning electron microscope (FEI Co., Ltd., New York, NY, USA) was performed to examine the surface morphology and element distribution of L1 and L2 after the adsorption of lanthanide. The zeta-potentials of LDHs samples were measured by electrophoretic light scattering method (Otsuka, Tokyo, Japan: ELSZ-2000ZS).

2.3. Adsorption Experiments Using L1 and L2 as Adsorbents

Twenty milligrams of L1 or L2 were put into contact with 30 mL of an aqueous solution containing La(III) or Eu(III) ion with known initial concentration. Batch adsorption experiments were conducted in the pH range of 4–6 (optimum pH), contact time from 10 min to 8 h, temperature 25 °C, and adsorbent dosage 20 mg. The pH of each solution was adjusted using 0.1 mol·L−1 NH4OH and 0.1 mol·L−1 HNO3. Following the adsorption experiment, the suspension was filtered (0.45 μm, Mixed Cellulose Ester 47 mm, Advantec MFS, Inc., Dublin, CA, USA). In order to discuss the target ions before and after the adsorption, we have dissolved the sample and measured the concentration of Ln ions by an Inductively Coupled Plasma-Mass Spectrometry (Agilent HP4500, Santa Clara, CA, USA). The operating conditions of Inductively Coupled Plasma-Mass Spectrometry (ICP-MS) are shown in Table 1.

2.4. Desorption and Regeneration Experiments

After the adsorption experiment, the exhausted adsorbent was washed and dried overnight. Then it was shaken in a 100 mL flask which contained 30 mL HCl or HNO3 solution with different concentrations. After the system reached equilibrium, the suspension was filtered, and finally, Eu(III) and La(III) content in the filtrate was determined. Subsequently, the adsorbent was neutralized by ultrapure water for several times and reconditioned for adsorption in succeeding cycles. The second adsorption experiments were carried out by using regeneration adsorbent under the optimum condition. The adsorption-desorption processes were repeated five times in this work.

2.5. Quantum Chemistry Calculation

All of the present quantum chemistry calculations were performed using the GAUSSIAN09 program package [28]. The molecular locally stable state structures were optimized by density functional theory (DFT) at the theoretical level of B3LYP. The basis sets were SDD (Stuttgart-Dresden) for Europium atoms and 6-311G (d, p) for other atoms. To take into account the effect of a water environment on our target systems, polarizable continuum medium (PCM method) with an appropriate dielectric constant for the solvent (water) was used [29]. To model the adsorbent, molecular cluster models for hydrotalcite lamellae whose chemical structure is described by Mg12Al7(OH)369+ were used [27,30]. The interlayer distance was kept 9.3 Å to imitate the experimentally observed one [31].

2.6. Calculation of Adsorption and Desorption Rates

The following equation was derived in Reference [32]:
C L n = C L n , 0 + α ( 1 e D t ) 1 β e D t .
Here, C L n is the metal ion concentration at a given time t in the aqueous solution in mmol L−1 and C L n , 0 is its initial concentration in mmol L−1. In Equation (1), we defined
α = ( B D ) M 2 A V ,
β = B D B + D .
In Equations (2) and (3), B, D, and M/(AV) are the constants determined by adsorption constant kad, desorption constant kd, and the volume of the solution in liters V. By fitting the experimental data on the metal ion concentration, C L n , versus time, t, by Equation (1), we obtain the constants, α, β, and D because C L n , C L n , 0 , and t are determined experimentally. Using the constants, α, β, and D, obtained in this way, we can get B, D, and M/(AV) from Equations (2) and (3),
In Equation (2), we have
A = k d K M / V ,
where K is the adsorption equilibrium constant of the solute,
K = k a d / k d .
From Equation (4), we immediately find
k d = A V / ( M K ) .
Because we have obtained M/(AV) above and K is known from the experiment, we can obtain k d from Equation (6). Then, from Equation (5), k a d is expressed as
k a d = K k d .
Above, k d and k a d are the required adsorption and desorption rates, respectively.
To fit the aqueous concentration of adsorbates versus time by Equation (1) and to obtain k d and k a d , we have used a genetic algorithm [33]. The remarkable difference between our work and Ref. [32] is that k D E S , L n was determined by iterating Equation (1) until the best fit was obtained to the experimental results in the latter case, while the genetic algorithm makes it unnecessary to do such computationally demanding iterations in order to obtain k d and k a d in the former case.

3. Results and Discussion

3.1. Adsorption Experiment

Figure 1 compares the experimental results of single La(III) and Eu(III) by using L1 or L2 at pH 4. The rapid increase of the adsorption capacity in the initial stage might be associated with the abundant reactive sites. As time increases further, the accessibility of the REEs ions to unoccupied active sites on the adsorbent surface gradually decreases, and these sites ultimately become saturated when the process reaches its equilibrium. L2 adsorbs more than L1 does, which agrees with the adsorption and desorption rate calculation results, shown in Section 3.4, later. Figure 1 shows that Eu(III) has a larger adsorption ability under identical conditions. The same tendency was found under the presence of other REEs ions (i.e., in multiple solution; refer to Figure S2). This may be attributed to the fact that the stability constant of La-EDTA is smaller than that of Eu-EDTA, which are 15.5 and 17.4, respectively [34,35]. The data of adsorption isotherms, kinetic models and the thermodynamic study were shown in Supplementary Information as Figures S3–S10 and Tables S1–S6.
For many adsorption studies, adsorption performance is generally assessed and expressed by the adsorption capacity. However, the judgement based only on the adsorption capacity is not objective, because the maximum adsorption capacity is sensitively affected by the initial concentration of target ions. If sorbent is exposed to a higher concentration of target ions, it is usually exhibiting a higher adsorption capacity [36,37].
Hence, instead of using adsorption capacity alone, the concept of the partition coefficient (PC) was also used to explain the performance of the adsorbent. In general, PC evaluation permits the conclusion that higher PC refers to a good adsorption affinity. Therefore, the detailed experimental data of adsorption process under different effect factors were provided in Table 2. Both of the effects of contact time and initial concentration were compared to confirm the adsorption performance. From Table 2, as long as the initial concentration is the same (120 μg·L−1), the PC of L2 is higher than that of L1 even under different contact time. This indicates that the adsorption affinity of L2 is better than L1 no matter whether the concept of adsorption capacity or that of the PC was used. Unlike the effect of contact time, a lower PC was obtained at higher adsorption capacity. This means that the better adsorption capacity is not necessarily equivalent to an outstanding adsorption affinity. The comparison of the data obtained at different initial concentrations suggests that the adsorbent performance should be compared after adjusting these effects. Essentially, the partition coefficient for a solid–liquid adsorption system represents the ratio of analyte concentration in and on the solid adsorbent phase to its concentration in the liquid phase at equilibrium [38].
It may be presumed that the interaction strength between the absorbate and the adsorbent is dominated by the surface charge of the adsorbent. To verify this assumption, the zeta potentials of the sample (L1 and L2) dispersed in aqueous solution were measured. As shown in Table 3, the zeta potentials of L1 and L2 at pH 4 are 35.3 mV and 20.2 mV, respectively. Both the zeta potentials of L1 and L2 are positive, which implies that the host layers may be positively charged. It is important to note that the zeta potential of L1 is larger than that of L2. On the other hand, the adsorption ability is larger for L2 than for L1, as mentioned before. Moreover, with the increase of pH, the zeta potential decreases (as shown in Figure 2) and the adsorption ability of REEs ions increases. In other words, it is suggested that the sample having high adsorption capacity has a low zeta potential. Therefore, the complexation of EDTA (i.e., LDH-EDTA) leads to a lower zeta potential, which leads to the adsorption ability of REEs ions (i.e., cation, having positive charge) larger than LDH.

3.2. Characterization of L1 and L2

The Zn:Al ratio and the amount of Ln absorbed are shown in Table 4. The component analysis result, shown in Table 4, indicates that the molar ratio of Zn/Al is approximately equal to that in the preparing solution, which shows that the dosage of reagents is reasonable. L1 and L2 were dissolved by HNO3 after adsorption of Ln ions. The amounts of Ln ions in the LDH phase, which were obtained by an ICP-MS are also shown in Table 4. The adsorption ability of REEs ions can be evaluated from these results.
The SEM images of (a) L1, (b) L2, (c) L2 after the La(III) adsorption (L2-La), (d) L2 after the Eu(III) adsorption (L2-Eu), and (e) the mapping image of L2-Eu are shown in Figure 3. Panels (b), (c), and (d) indicate that the surface of the adsorbent becomes rough at the end of the adsorption processes. The mapping images indicate that REEs ions are adequately adsorbed onto the adsorbent and that they are distributed uniformly.
Figure 4 shows the (003), (006) and (110) peaks, which indicate that the layered structures are observed for all the L1, L2, and L2-Eu. After the intercalation by EDTA and the adsorption process, the (110) peak is still visible, as shown in Figure 4b,c. These figures show that the layer structure was destroyed by the intercalating process neither for EDTA nor for EDTA-Ln(III). However, the layer structure may be destroyed by repeated desorption process, which results in the decrease of the removal efficiency after the cycle adsorption. Based on the angle of peak (003) corresponding to the basal spacing, the basal spacing values of L1, L2 and L2-Eu were estimated to be 0.89 nm, 1.47 nm and 1.47 nm, respectively. These values were calculated by the Bragg equation. The interlayer distance was obtained by subtracting from the basal spacing to the layer width (0.48 nm) [39]. The interlayer distance of L2 and L2-Eu indicates that EDTA coexists and that EDTA is intercalated into L1 successfully. The adsorption process also increases the interlayer distance because it is close to the dimensions of EDTA complexes (0.9–1 nm) generally found for by single-crystal XRD of M-EDTA (M: Metallic ions) compound [40].
The very sharp peak at 1380 cm−1 in Figure 5a can be attributed to the NO3 stretching vibration, while the corresponding peak is absent in Figure 5b,c. The absorption bands around at 1624 and 1394 cm−1 are characteristics of the symmetrical and asymmetrical vibrations of COO-groups [39]. The peaks around 2950 cm−1 (-CH) and 1111 cm−1 (-CN) are typical peaks of EDTA as well. A relatively sharp band at 776 cm−1 is assigned to C-H out-of-plane bending, and a broad band at about 609 cm−1 is due either to a vibrational mode of hydroxyl groups and water or to a metal-oxygen vibration ν(M-O). Two peaks at about 562 and 430 cm−1 are attributed to ν(M-O) and δ(M-O-M) (M: Zn, Al or Eu), respectively [41,42]. No significant changes were observed in this region upon the incorporation of the Europium ions.

3.3. Regeneration Studies

From industrial and technological points of view, it is desirable to recover and reuse the adsorbed material. The effective reuse of adsorbent materials directly affects the cost factor, so its practicability in continuous batch adsorption process is important [43,44]. In this work, the regeneration of the used EDTA-LDH adsorbent was investigated. The desorption efficiency of the exhausted adsorbent is shown in Figure 6, and the removal efficiency of the regenerated L2 adsorbent is shown in Figure 7. From Figure 6, the desorption by HNO3 shows good desorption efficiency for both La(III) and Eu(III).
The optimum concentration was 200 mmol·L−1. The desorption efficiency of La(III) was found to be 95.67%, whereas that of Eu(III) was slightly lower; this may be due to the fact that the stability constant of Eu-EDTA (17.4) is higher than that of La-EDTA (15.5). Figure 7 shows the better reuse performance of L2 after three cycles. The hydrogen ions may destroy the structure during the desorption process, which results in the decrease of the removal efficiency after the fourth or fifth cycle.

3.4. Numerical Results of Adsorption and Desorption Rates

The temporal evolutions of the ion concentrations of Eu(III) and La(III) in the aqueous solution are shown in Figure 8 ((a) Eu(III), (b) La(III)). In general, the ion concentrations in aqueous solution decrease unilaterally with time, which is consistent with Equation (1). In addition, it can be found that the genetic algorithm actually fits the experimental time evolutions of the ion concentrations very well.
The adsorption and desorption rates calculated from the fittings in Figure 8 are shown in Table 5. We can see that the adsorption rate is larger for Eu(III) than for La(III), which agrees with the experimental result that Eu(III) has a higher adsorption ability under the identical conditions, as shown in Section 3.3. In addition, Table 5 suggests that Eu(III) desorbs more easily than La(III). As shown in Figure 7 experimentally, the recovery efficiency of La(III) is smaller than that of Eu(III). This means that the rate of desorption is low if the recovery efficiency is small. Therefore, our numerical results reasonably suggest that recovery efficiency is large if the rate of desorption is large.

3.5. Quantum Chemistry Calculation

The structures of the complexes themselves are shown in Supplementary Information (Figure S11 and Tables S7–S10). Figure 9 shows the energies of (a) [Eu(H2O)8]3+ adsorbed on precursor LDHs and (b) Eu3+ adsorbed on EDTA-intercalated LDH relative to the dissociation limit. It can be seen that Eu3+ adsorbed on EDTA-intercalated LDH is more stable than the dissociation limit (panel (b)), whereas [Eu(H2O)8]3+ adsorbed on precursor LDH is more unstable than the dissociation limit (panel (a)). This implies that the adsorption of Eu3+ on EDTA-intercalated LDH is energetically more probable than that on precursor LDH. Actually, the strongly intercalated structure between EDTA and LDH, shown in panel (b), indicates a more stable configuration than the weakly bonded structure between coordination H2O and LDH, shown in panel (a). This is reasonable because the Coulomb repulsion between precursor LDHs (charge = +14) and [Eu(H2O)8]3+ (charge = +3) brings about the electric instability of [Eu(H2O)8]3+ adsorbed on precursor LDH in panel (a), whereas the Coulomb attraction between precursor LDH (charge = +14) and [Eu(EDTA)(H2O)] (charge = −1) brings about the electric stability of Eu3+ adsorbed on EDTA-intercalated LDH in panel (b). This is a result consistent with the experimental finding that the complexation of EDTA improves the adsorption capacity of LDH.
We note that the experimental results were shown for different adsorption efficiencies of phosphate anions on [Mg-Al]-LDH and [Zn-Al]-LDH [15]. The work concluded that the adsorption efficiency of the latter is larger than that of the former. This is basically due to the fact that the different ionic radii lead to different crystalline structures. Because of this, [Mg-Al]-LDH does not change its crystalline structure upon the adsorption of phosphate, while [Zn-Al]-LDH dramatically changes it, in order to ease the adsorption of phosphate. If this is the case, the replacement of Mg cations by Zn cations may lead to different adsorption efficiencies, shown in the present work, which originates from their different crystalline structures.

3.6. Comparison with Other Adsorbents

In general, although the adsorption affinity is usually compared based on the adsorption capacity, it is difficult to evaluate each adsorbent on a parallel basis by using adsorption capacity alone. Adsorption experiments of each material are usually conducted under various initial concentrations or volumes of solution. From Table 2, although the adsorption capacity obtained at high initial concentration is larger than that at lower initial concentration, the PC at high initial concentration shows a reverse tendency. Hence, the concept of PC was also used here to compare the adsorption affinity with other reported adsorbents. The comparison of the adsorption properties of other previous adsorbents was shown in Table 6.
From Table 6, EDTA modified LDH which synthesized in this work has relatively high adsorption capacity (145 μg·g−1 for L2-Eu, 124 μg·g−1 for L2-La) relative to only LDH. It means the chelating agents improve the adsorption affinity of LDH. However, comparing with some adsorbents which were exposed to a higher concentration of target ions measured under their adsorption experiments, it shows a lower adsorption capacity. It is difficult to compare each other under different conditions (even thousand times difference among them). This is a result of adsorption capacity sensitively to the initial concentration of target ions. Therefore, introducing the concept of PC is also important to reduce this difference. The PC of both L2-Eu and L2-La are larger in comparison with others, which can indicate that LDH intercalated by EDTA has superior adsorption performance.

4. Conclusions

In conclusion, we performed XRD, SEM-EDS, and FT-IR experiments to characterize layered double hydroxides (LDHs) modified with EDTA and to identify the prepared LDH-NO3 (L1) and LDH-EDTA (L2). In addition, based on these experimental results, the adsorption capabilities and mechanisms of REEs ions on L1 and L2 were experimentally accessed. It was shown that both La(III) and Eu(III) could rapidly be adsorbed, and that the adsorption capacity of L2 was larger than that of L1 no matter whether the concept of adsorption capacity or that of the PC was used. In addition, comparing with other adsorbents, the PC of both L2-Eu and L2-La are relatively larger. It indicates that LDH intercalated by EDTA has superior adsorption performance. The complexation of REEs ions with EDTA (i.e., L2) brought about a lower zeta potential, which indicates that the former plays the role more dominant than the latter and that the latter is just a secondary result stemming from the former. In accord with this, quantum chemistry calculations showed that the complexation actually leads to the negatively charged species being adsorbed on LDHs, which promotes the adsorption of the complex compared with REEs without complexation. They also showed that the species in the presence of EDTA is more stable than that in its absence and that the intercalation of EDTA does indeed improve the stability of the adsorbed species. Finally, from our numerical results of adsorption and desorption rates, it was suggested that the recovery efficiency is large if the rate of desorption is large. Both the quantum chemistry calculation results and the adsorption and desorption rate results are consistent with the experimental ones. The LDH-EDTA synthesized in this work have a high affinity for removing REEs ions.

Supplementary Materials

The following are available online at https://www.mdpi.com/2076-3417/9/22/4805/s1, Figure S1: Effect of common ions (Ca2+, Mg2+, Na+ and K+) on the removal efficiency of lanthanides using ground original sample, Figure S2: Adsorption capacity of (a) La(III) and (b) Eu(III) both in single and multiple solution by using L1 or L2, Figure S3: Langmuir isotherm for La(III) adsorption onto L2, Figure S4: Freundlich isotherm for La(III) adsorption onto L2, Figure S5: Langmuir isotherm fitting for Eu(II) adsorption onto L2, Figure S6: Freundlich isotherm fitting for Eu(III) adsorption onto L2, Figure S7: The kinetic models for La(III) adsorption onto L2, Figure S8: The kinetic models for Eu(III) adsorption onto L2, Figure S9: Plots of lnKd vs. 1/T for the estimation of thermodynamic parameters obtained for the adsorption of (a) La(III) and (b) Eu(III) on L2, Figure S10: Plots of qe vs. ln (qe/ce) for (a) La(III) and (b) Eu(III) on L2, Figure S11: Optimized structures of (a) [Eu(H2O)8]3+ and (b) [Eu(EDTA)(H2O)]. Table S1: Coefficients of Langmuir and Freundlich isotherms for La(III) adsorption onto L2, Table S2: Coefficients of Langmuir and Freundlich isotherms for Eu(III) adsorption onto L2, Table S3: Kinetic parameters of Pseudo-first order and Pseudo-second order, Table S4: Kinetic parameters of Elovich model and Intraparticle diffusion, Table S5: Thermodynamic parameters for the adsorption of La(III) on L2, Table S6: Thermodynamic parameters for the adsorption of Eu(III) on L2, Table S7: Bond lengths of [Eu(H2O)8]3+ shown in Panel (a) of Figure S3, Table S8: Bond lengths of [Eu(EDTA)(H2O)] in Panel (b) of Figure S11, Table S9: Bond lengths of [Eu(H2O)8]3+ shown in Panel (a) of Figure S11, Table S10: Bond lengths of [Eu(EDTA)(H2O)] in Panel (b) of Figure S11.

Author Contributions

Experiment and writing, S.Z.; data evaluation and supervision N.K.; Calculation, K.M.; measurements of Zeta-potential, H.O.

Funding

This research was funded by Grant-in-Aid for Scientific Research Program (C), grant number 16K00599 from the Japan Society for the Promotion of Science and the project of the Jiangxi Provincial Department of Education, grant number GJJ180385. This research was also supported by a Fund for Promotion of Niigata University KAAB Projects from the Ministry of Education, Culture, Sports, Science and Technology, Japan.

Acknowledgments

The authors are grateful to M. Ohizumi of Office for Environment and Safety, M. Teraguchi of the Facility of Engineering and T. Iwafune of the Center of Instrumental Analysis in Niigata University for permitting the use of ICP-MS, FT-IR and XRD; and for giving helpful advice on measurements. Computations were performed by the Research Center for Computational Science, Okazaki, Japan.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Xu, X.; Zou, J.; Zhao, X.R.; Jiang, X.Y.; Jiao, F.P.; Yu, J.G.; Liu, Q.; Teng, J. Facile assembly of three-dimensional cylindrical egg white embedded graphene oxide composite with good reusability for aqueous adsorption of rare earth elements. Colloids Surf. A 2019, 570, 127–140. [Google Scholar] [CrossRef]
  2. Toth, M.T.; Schubert, F.; Raucsik, B.; Fintor, K. Mineralogical and geochemical constraints of the REE accumulation in the almasfuzito red mud depository in northwest Hungary. Appl. Sci. 2019, 9, 3654. [Google Scholar] [CrossRef]
  3. Laurino, J.P.; Mustacato, J.; Huba, Z.J. Rare earth element recovery from acidic extracts of Florida Phosphate mining materials using Chelating polymer 1-Octadecene, polymer with 2,5-Furandione, Sodium salt. Minerals 2019, 9, 477. [Google Scholar] [CrossRef]
  4. Chour, Z.; Laubie, B.; Morel, J.L.; Tang, Y.; Qiu, R.L.; Simonnot, M.O.; Muhr, L. Recovery of rare earth elements from Dicranopteris dichotoma by an enhanced ion exchange leaching process. Chem. Eng. Process 2018, 130, 208–213. [Google Scholar] [CrossRef]
  5. Iftekhar, S.; Srivastava, V.; Ramasamy, D.L.; Naseer, W.A.; Sillanpaa, M. A novel approach for synthesis of exfoliated biopolymeric-LDH hybrid nanocomposites via in-stiu coprecipitation with gum Arabic: Application towards REEs recovery. Chem. Eng. J. 2018, 347, 398–406. [Google Scholar] [CrossRef]
  6. Yang, M.J.; Liang, X.L.; Ma, L.Y.; Huang, J.; He, H.P.; Zhu, J.X. Adsorption of REEs on kaolinite and halloysite: A link to the REE distribution on clays in the weathering crust of granite. Chem. Geol. 2019, 525, 210–217. [Google Scholar] [CrossRef]
  7. Sakamoto, N.; Kano, N.; Wang, Y.; Gao, L.D.; Imaizumi, H. Biosorption of Lanthanides using three kinds of seaweed biomasses. Radioisotopes 2010, 59, 623–636. [Google Scholar] [CrossRef]
  8. Koto, Y.; Kano, N.; Wang, Y.; Sakamoto, N.; Imaizumi, H. Biosorption of lanthanides from aqueous solutions using pretreated Buccinum tenuissimum shell biomass. Bioinorg. Chem. Appl. 2010. [Google Scholar] [CrossRef] [PubMed]
  9. Wang, J.; Chen, C. Biosorbents for heavy metals removal and their future. Biotechnol. Adv. 2009, 27, 195–226. [Google Scholar] [CrossRef]
  10. Iannicelli-Zubiani, E.M.; Gallo Stampino, P.; Cristiani, C.; Dotelli, G. Enhanced lanthanum adsorption by amine modified activated carbon. Chem. Eng. J. 2018, 341, 75–82. [Google Scholar] [CrossRef]
  11. Sotiles, A.R.; Baika, L.M.; Grassi, M.T.; Wypych, F. Cation exchange reactions in layered double hydroxides intercalated with sulfate and alkaline cations (A(H2O)6)[M2+6Al3(OH)18(SO4)2]·6H2O (M2+ = Mn, Mg, Zn; A+ = Li, Na, K). J. Am. Chem. Soc. 2019, 141, 531–540. [Google Scholar] [CrossRef] [PubMed]
  12. Sotiles, A.R.; Wypych, F. Converting Mn/Al layered double hydroxide anion exchangers into cation exchangers by topotactic reactions using alkali metal sulfate solutions. Chem. Commun. 2019, 55, 7824–7827. [Google Scholar] [CrossRef] [PubMed]
  13. Du, J.Z.; Jin, L.; Zeng, H.Y.; Shi, X.K.; Zhou, E.G.; Feng, B.; Sheng, X. Flame retardancy of organic-anion-intercalated layered double hydroxides in ethylene vinyl acetate copolymer. Appl. Clay Sci. 2019, 180, 105193. [Google Scholar] [CrossRef]
  14. Zhao, P.W.; Liu, X.H.; Tian, W.L.; Yan, D.P.; Sun, X.M.; Lei, X.D. Adsolubilization of 2,4,6-trichlorophenol from aqueous solution by surfactant intercalated ZnAl layered double hydroxides. Chem. Eng. J. 2015, 279, 597–640. [Google Scholar] [CrossRef]
  15. Bernardo, M.P.; Ribeiro, C. [Mg-Al]-LDH and [Zn-Al]-LDH as matrices for removal of high loadings of phosphate. Mater. Res. 2018, 21, e20171001. [Google Scholar] [CrossRef]
  16. Oestreicher, V.; Jobbágy, M.; Regazzoni, A.E. Halide exchange on Mg(II)-Al(III) layered double hydroxides: Exploring affinities and electrostatic predictive models. Langmuir 2014, 30, 8408–8415. [Google Scholar] [CrossRef]
  17. Costantino, U.; Vivani, R.; Bastianini, M.; Costantino, F.; Nocchetti, M. Ion exchange and intercalation properties of layered double hydroxides towards halide anions. Dalton Trans. 2004, 30, 11587–11596. [Google Scholar] [CrossRef]
  18. Zubair, M.; Daud, M.; McKay, G.; Shehzad, F.; Al-Harthi, M.A. Recent progress in layered double hydroxides (LDH)-containing hybrids as adsorbents for water remediation. Appl. Clay Sci. 2017, 143, 279–292. [Google Scholar] [CrossRef]
  19. Chen, H.; Lin, J.H.; Zhang, N.; Chen, L.Z.; Zhong, S.P.; Wang, Y.; Zhang, W.Q.; Ling, Q.D. Preparation of MgAl-EDTA-LDH based electrospun nanofiber membrane and its adsorption properties of copper(II) from wastewater. J. Hazard. Mater. 2018, 345, 1–9. [Google Scholar] [CrossRef]
  20. Zhang, Z.Q.; Zeng, H.Y.; Liu, X.J.; Xu, S.; Chen, C.R.; Du, J.Z. Modification of MgAl hydrotalcite by ammonium sulfate for enhancement of lead adsorption. J. Taiwan Inst. Chem. Eng. 2016, 60, 361–368. [Google Scholar] [CrossRef]
  21. Zhang, S.; Kano, N.; Imaizumi, H. Adsorption of Cu (II), Pb (II) by Mg-Al Layered Double Hydroxides (LDHs): Intercalated with the chelating agents EDTA and EDDS. J. Chem. Eng. Jpn. 2014, 47, 324–328. [Google Scholar] [CrossRef]
  22. Mishima, K.; Zhang, S.; Minagawa, S.; Kano, N. Combined experimental and quantum chemical study on the adsorption mechanism of phosphorous anions on the hydrotalcite surfaces. Funct. Mater. Lett. 2016, 9, 1650061. [Google Scholar] [CrossRef]
  23. Smith, Y.R.; Bhattacharyya, D.; Willhard, T.; Misra, M. Adsorption of aqueous rare earth elements using carbon black derived from recycled tires. Chem. Eng. J. 2016, 296, 102–111. [Google Scholar] [CrossRef] [Green Version]
  24. Gandhi, M.R.; Meenakshi, S. Preparation and characterization of La(III) encapsulated silica gel/chitosan composite and its metal uptake studies. J. Hazard. Mater. 2012, 203–204, 29–37. [Google Scholar] [CrossRef]
  25. Ren, X.M.; Chen, C.L.; Nagatsu, M.; Wang, X.K. Carbon nanotubes as adsorbents in environmental pollution management: A review. Chem. Eng. J. 2011, 170, 395–410. [Google Scholar] [CrossRef]
  26. Yao, T.; Wu, X.W.; Chen, X.; Xiao, Y.P.; Zhang, Y.G.; Zhao, Y.L.; Li, F.B. Biosorption of Eu(III) and U(VI) on Bacillus subtilis: Macroscopic and modeling investigation. J. Mol. Liq. 2016, 219, 32–38. [Google Scholar] [CrossRef]
  27. Pu, M.; Zhang, B.F. Theoretical study on the microstructures of hydrotalcite lamellae with Mg/Al ratio of two. Mater. Lett. 2005, 59, 3343–3347. [Google Scholar] [CrossRef]
  28. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09; Gaussian, Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  29. Tomasi, J.; Mennucci, B.; Cammi, R. Quantum mechanical continuum solvation models. Chem. Rev. 2005, 105, 2999–3094. [Google Scholar] [CrossRef]
  30. Farouq, R.; Yousef, N.S. Equilibrium and kinetics studies of adsorption of copper (II) ions on natural biosorbent. Int. J. Chem. Eng. Appl. 2015, 6, 319–324. [Google Scholar] [CrossRef]
  31. Narita, E.; Yamagishi, T.; Tazawa, K.; Ichijo, O.; Umetsu, Y. Uptake behavior of chelating agents by magnesium-aluminum oxide precursor with reconstruction of hydrotalcite-like layer structure. Clay Sci. 1995, 9, 187–197. [Google Scholar]
  32. Machida, M.; Aikawa, M.; Tatsumoto, H. Prediction of simultaneous adsorption of Cu(II) and Pb(II) onto activated carbon by conventional Langmuir type equations. J. Hazard. Mater. 2005, 120, 271–275. [Google Scholar] [CrossRef] [PubMed]
  33. FORTRAN GA Version 1.7.1. Available online: http://www.cuaerospace.com/Technology/GeneticAlgorithm/PurchaseGeneticAlgorithm.aspx (accessed on 24 February 2019).
  34. Kameda, T.; Saito, S.; Umetsu, Y. Mg-Al layered double hydroxide intercalated with Ethylene diaminetetraacetate anion: Synthesis and application to the uptake of heavy metal ions from an aqueous solution. Sep. Purif. Technol. 2005, 47, 20–26. [Google Scholar] [CrossRef]
  35. Zhao, F.; Repo, E.; Meng, Y.; Wang, X.; Yin, D.; Sillanpaa, M. An EDTA-β-cyclodextrin material for the adsorption of rare earth elements and its application in preconcentration of rare earth elements in seawater. J. Colloid Interface Sci. 2016, 465, 215–224. [Google Scholar] [CrossRef] [PubMed]
  36. Vikrant, K.; Kim, K.H. Nanomaterials for the adsorptive treatment of Hg(II) ions from water. Chem. Eng. J. 2019, 358, 264–282. [Google Scholar] [CrossRef]
  37. O’Connor, D.J.; Connolly, J.P. The effect of concentration of adsorbing solids on the partition coefficient. Water Res. 1980, 14, 1517–1523. [Google Scholar] [CrossRef]
  38. Na, C.J.; Yoo, M.J.; Tsang, D.C.W.; Kim, H.W.; Kim, K.H. High-performance materials for effective sorptive removal of formaldehyde in air. J. Hazard. Mater. 2019, 366, 452–465. [Google Scholar] [CrossRef]
  39. Barnabas, M.J.; Parambadath, S.; Mathew, A. Highly efficient and selective adsorption of ln3+ on pristine Zn/Al layered double hydroxide (Zn/Al-LDH) from aqueous solutions. J. Solid State Chem. 2016, 233, 133–142. [Google Scholar] [CrossRef]
  40. Zheng, Y.Q.; Cheng, B.; You, W.; Yu, J.G.; Ho, W.K. 3D hierarchical graphene oxide-NiFe LDH composite with enhanced adsorption affinity to Congo red, methyl orange and Cr(VI) ions. J. Hazard. Mater. 2019, 369, 214–225. [Google Scholar] [CrossRef]
  41. Kloprogge, J.T.; Hickey, L.; Frost, R.L. FT-Raman and FT-IR spectroscopic study of synthetic Mg/Zn/Al-hydrotalcites. J. Raman Spectrosc. 2004, 35, 967–974. [Google Scholar] [CrossRef] [Green Version]
  42. Dolatyari, L.; Yaftian, M.R.; Rostamnia, S. Adsorption characteristics of Eu(III) and Th(IV) ions onto modified mesoporous silica SBA-15 materials. J. Taiwan Inst. Chem. Eng. 2016, 60, 174–184. [Google Scholar] [CrossRef]
  43. Shen, Y.; Zhao, X.; Zhang, X.; Li, S.; Liu, D.; Fan, L. Removal of Cu2+ from the aqueous solution by tartrate-intercalated layered double hydroxide. Desalin. Water Treat. 2016, 57, 2064–2072. [Google Scholar] [CrossRef]
  44. Iftekhar, S.; Srivastava, V.; Sillanpää, M. Synthesis and application of LDH intercalated cellulose nanocomposite for separation of rare earth elements (REEs). Chem. Eng. J. 2017, 309, 130–139. [Google Scholar] [CrossRef]
  45. Kano, N.; Pang, M.L.; Deng, Y.L.; Imaizumi, H. Adsorption of Rare Earth Elements (REEs) onto activated carbon modified with Potassium Permanganate (KMnO4). J. Appl. Solut. Chem. Model. 2017, 6, 51–61. [Google Scholar] [CrossRef]
  46. Saito, G.P.; Romero, J.H.S.; Cebim, M.A.; Davolos, M.R. Eu (III) doped LDH intercalated with cinnamate anion as multifunctional sunscreens. J. Lumin. 2018, 203, 160–164. [Google Scholar] [CrossRef]
  47. Shao, D.D.; Fan, Q.H.; Li, J.X.; Niu, Z.W.; Wu, W.S.; Chen, Y.X.; Wang, X.K. Removal of Eu(III) from aqueous solution using ZSM-5 zeolite. Micropor. Mesopor. Mat. 2009, 123, 1–9. [Google Scholar] [CrossRef]
  48. Chevinli, A.S.; Najafi, M.; Sillanpaa, M. Removal of La (III) ions from aqueous solution by Lanthanide MOF; characterization, synthesizing and process conditions study. Environ. Nanotechnol. Monit. Manag. 2019, 12, 100216. [Google Scholar]
Figure 1. The adsorption capacity of La(III) and Eu(III) by using L1 or L2.
Figure 1. The adsorption capacity of La(III) and Eu(III) by using L1 or L2.
Applsci 09 04805 g001
Figure 2. Zeta potential of the L2 dispersed in aqueous solution with different pH.
Figure 2. Zeta potential of the L2 dispersed in aqueous solution with different pH.
Applsci 09 04805 g002
Figure 3. SEM images of (a) L1, (b) L2, (c) L2-La, (d) L2-Eu and (e) mapping image of L2-Eu.
Figure 3. SEM images of (a) L1, (b) L2, (c) L2-La, (d) L2-Eu and (e) mapping image of L2-Eu.
Applsci 09 04805 g003
Figure 4. XRD patterns of (a) L1, (b) L2, (c) L2-Eu.
Figure 4. XRD patterns of (a) L1, (b) L2, (c) L2-Eu.
Applsci 09 04805 g004
Figure 5. FT-IR spectra of (a) L1, (b) L2, (c) L2-Eu.
Figure 5. FT-IR spectra of (a) L1, (b) L2, (c) L2-Eu.
Applsci 09 04805 g005
Figure 6. Desorption efficiency of Eu(III) and La(III) ions using HCl or HNO3.
Figure 6. Desorption efficiency of Eu(III) and La(III) ions using HCl or HNO3.
Applsci 09 04805 g006
Figure 7. The recovery efficiency of La(III) and Eu(III) ions using regenerated L2.
Figure 7. The recovery efficiency of La(III) and Eu(III) ions using regenerated L2.
Applsci 09 04805 g007
Figure 8. Temporal evolutions of the ion concentrations of (a) Eu(III) and (b) La(III) in the aqueous solution.
Figure 8. Temporal evolutions of the ion concentrations of (a) Eu(III) and (b) La(III) in the aqueous solution.
Applsci 09 04805 g008
Figure 9. Relative energies of (a) [Eu(H2O)8]3+ adsorbed on precursor LDH and (b) Eu3+ adsorbed on EDTA-intercalated LDH.
Figure 9. Relative energies of (a) [Eu(H2O)8]3+ adsorbed on precursor LDH and (b) Eu3+ adsorbed on EDTA-intercalated LDH.
Applsci 09 04805 g009
Table 1. Operating conditions of Inductively Coupled Plasma-Mass Spectrometry (ICP-MS).
Table 1. Operating conditions of Inductively Coupled Plasma-Mass Spectrometry (ICP-MS).
Radio Frequency Power1400 W
Plasma gas flow15 L·min−1
Carrier gas flow1.2 L·min−1
Sampling depth6.5 mm
Sample uptake rate0.5 mL·min−1
Measurement point3 points/peak
Integration time1.0 sec/point
Measured Isotope139La, 153Eu
Table 2. Detailed experimental data of adsorption process under different effect factor.
Table 2. Detailed experimental data of adsorption process under different effect factor.
Target IonsAdsorbentEffect FactorFinal Concentration (μg·L−1)Adsorption Capacity (μg·g−1)Partition Coefficient (μg·g−1·μM−1)
EuL1Contact time (min)1096.25.678.95
3093.59.8316.0
6091.313.121.9
12083.424.945.5
L2Contact time (min)1061.293230
3038.2127506
6031.0138676
12026.4145834
Initial concentration (μg·L−1)207.824.0468
5018.245.0376
20054.6165275
300153222220
LaL1Contact time (min)1098.81.82.53
3088.71726.6
6086.919.731.5
12086.120.833.6
L2Contact time (min)1023.9114341
3046.3115346
6046.3119345
12040.8124405
Initial concentration (μg·L−1)2040.717.4422
5020.939.6263
200101148203
300163205174
Table 3. Zeta potentials of the sample (L1 and L2) dispersed in aqueous solution at pH 4.
Table 3. Zeta potentials of the sample (L1 and L2) dispersed in aqueous solution at pH 4.
SampleL1L2
Zeta potentials35.3 mV20.2 mV
Table 4. Chemical analysis results of L1 and L2.
Table 4. Chemical analysis results of L1 and L2.
Sample%wtAtomic RatiosLn Absorbed (μg·g−1)Proposed Formula
CNHZn/AlC/NH/NLaEu
L10.064.262.382.100.007.8215.918.4[Zn2Al (OH)6]NO3
L213.73.353.601.670.3115.187.6103[Zn2Al(OH)6]2[C10H14N2O8]
Table 5. Adsorption rate ( k a d ) and desorption rate ( k d ) of Eu(III) and La(III).
Table 5. Adsorption rate ( k a d ) and desorption rate ( k d ) of Eu(III) and La(III).
Target Ion k a d / L · mmol 1 · h 1 k d / h 1
Eu(III)0.0015350.5686
La(III)0.0010010.2176
Table 6. The comparison of adsorption properties of various of adsorbent.
Table 6. The comparison of adsorption properties of various of adsorbent.
Target IonsMaterialsFinal Concentration (μg·L−1)Adsorption Capacity (μg·g−1)Partition Coefficient (μg·g−1·μM−1)Reference
EuActivated carbon61.292.7230[45]
Buccinum tenuissimum34.798429[8]
Sargassum hemiphyllum20.7119746[7]
Fe3O4@MnOx28.1 × 103138 × 103745[46]
ZSM-5 zeolite30.424.2121[47]
L1 (ZnAl-NO3)83.424.945.5This study
L2 (ZnAl-EDTA)26.4145834This study
LaActivated carbon52.771.0205[45]
Buccinum tenuissimum34.099.0401[8]
Sargassum hemiphyllum35.197.3421[7]
Schizymenia dubyi39.790.5392[7]
Lanthanide MOF62.0 × 10373.0 × 103179[48]
L1 (ZnAl-NO3)86.120.833.6This study
L2 (ZnAl-EDTA)17.7124422This study

Share and Cite

MDPI and ACS Style

Zhang, S.; Kano, N.; Mishima, K.; Okawa, H. Adsorption and Desorption Mechanisms of Rare Earth Elements (REEs) by Layered Double Hydroxide (LDH) Modified with Chelating Agents. Appl. Sci. 2019, 9, 4805. https://doi.org/10.3390/app9224805

AMA Style

Zhang S, Kano N, Mishima K, Okawa H. Adsorption and Desorption Mechanisms of Rare Earth Elements (REEs) by Layered Double Hydroxide (LDH) Modified with Chelating Agents. Applied Sciences. 2019; 9(22):4805. https://doi.org/10.3390/app9224805

Chicago/Turabian Style

Zhang, Shuang, Naoki Kano, Kenji Mishima, and Hirokazu Okawa. 2019. "Adsorption and Desorption Mechanisms of Rare Earth Elements (REEs) by Layered Double Hydroxide (LDH) Modified with Chelating Agents" Applied Sciences 9, no. 22: 4805. https://doi.org/10.3390/app9224805

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