Next Article in Journal
Geographical Origin Authentication of Edible Chrysanthemum morifolium Ramat. (Hangbaiju) Using Stable Isotopes
Previous Article in Journal
Effect of Na+ on the Adsorption Behavior of Polystyrene Nanoparticles onto Coal and Quartz Surfaces
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Improvement of Gd(III) Solvent Extraction by 4-Benzoyl-3-methyl-1-phenyl-2-pyrazolin-5-one: Non-Aqueous Systems

1
Department of General and Inorganic Chemistry, University of Chemical Technology and Metallurgy, 8 Kliment Okhridski Blvd, 1756 Sofia, Bulgaria
2
Institute of General and Inorganic Chemistry, Bulgarian Academy of Science, Acad. G. Bonchev Street, Block 11, 1113 Sofia, Bulgaria
*
Author to whom correspondence should be addressed.
Separations 2023, 10(5), 286; https://doi.org/10.3390/separations10050286
Submission received: 2 April 2023 / Revised: 20 April 2023 / Accepted: 24 April 2023 / Published: 3 May 2023

Abstract

:
The study of the liquid–liquid extraction of gadolinium (Gd(III) ion) with a chelating compound, 4-benzoyl-3-methyl-1-phenyl-2-pyrazolin-5-one (HP), and the determination of the process parameters are presented by employing two ionic liquids, namely, ([C1Cnim+][Tf2N], n = 4, 10) and CHCl3, as diluents. Compared to CHCl3, the ionic liquid offers increased distribution ratios in an aqueous medium. A step forward, enhanced solvent extraction, and improved separation upon the addition of ethylene glycol are demonstrated, i.e., a boost of two immiscible organic phases, compared to traditional aqueous solutions. However, this is noticeable when using CHCl3, but unfortunately not with ionic liquid combination, [C1C10im+][Tf2N]. Several conclusions are given, highlighting the role of the ionic diluent in complexation processes and selectivity with an employment of the chelating agent HP for various metal s-, p-, d-, and f-cations, i.e., nearly 25 metals. A detailed evaluation of the selectivity between these metals was made when changing both the aqueous phase completely with ethylene glycol or partially with glycerol (1:1). Electron paramagnetic resonance (EPR) spectroscopy has been used to study the established chemical species in the obtained organic extracts, such as Gd3+, Fe3+, Cu2+, and Cr3+, with unpaired electrons.

1. Introduction

In reality, the intriguing alchemists’ research on a “Menstruum universal” (universal diluent) somehow indicates the importance given to the diluent and the process of solubility in a larger scheme, with the development of the popular pronouncement nowadays in chemistry, i.e., “Similia Similibus Solvuntur” (similar solubilizes similar). The diluents (a solvent is a liquid phase containing more than one substance, i.e., solvent = diluent + extractant + modifier…IUPAC Gold Book) [1] usually used in all fields of chemistry are often classical ones and the choice can affect all solvent extraction processes as well as the separation effort. Thus, when planning the solvent system several factors can be taken into account choosing one, regarding its viscosity, polarity, solubility in water, a high flash point, a low freezing point, preferably a high boiling point, price, etc. [2]. Further, the main criteria for the choice include toxicity, chemical reaction hazard, etc. Another consideration when selecting an extraction diluent is, of course, its density. In other words, diluents that are denser than water will form the lower layer of the pair when mixing together, while diluents that are less dense than water will form the upper layer or “float” on water. However, it should contain preferably only carbon, hydrogen, nitrogen, and oxygen in order to be totally incinerable (the “CHON principle”). In fact, various diluents exist, both aromatic and aliphatic, and they range from straightforward molecules to complex mixtures. As a whole, classical diluents are low molecular weight organic substances that belong to two different broad chemical families: hydrocarbons (aliphatic, cyclic, and aromatic hydrocarbons) and oxygenates (esters and ethers). Despite the fact that it is difficult to trace them far back in history, an enormous quantity of liquid organic compounds were synthesized since the early nineteenth century [3]. However, high water solubility in the organic phase is definitely a decreasing factor for the solvent extraction ability of the system in use [4]. Thus, the liquid–liquid extraction method has a significant drawback: the requirement of water-immiscible organic diluents that are often toxic, flammable, or volatile [5]. The growing awareness of safety and environmental impact related to the use of molecular diluents renders their replacement with less noxious alternatives desirable [6]. As a whole, the development of new types of diluents is likely an ongoing process in chemistry. One important idea under intensive research studies nowadays is, of course, the replacement of traditional organic diluents with ionic liquid (ILs) compounds [7,8,9]. In order to highlight the ionic character of these similar reagents, they are usually denoted as [Cat+][Ani] considering the ILs’ bulky organic cations with a basic cyclic structure and smaller inorganic or organic anions. In particular, in the field of metal extraction, separation, and recycling by liquid–liquid processes, hydrophobic ILs of the imidazolium type [CnCmim+][Tf2N] have already proved to be tremendously efficient media for a large variety of metals in the periodic table including lanthanoids (cations are denoted as [CnCmim+] for n-alkyl-m-alkylimidazolium, whereas they are denoted as the anion [(CF3SO2)2N] (bis(trifluoromethylsulfonyl) for imide, [Tf2N]) [10,11,12,13,14,15]. However, in one way or another, the prediction of diluent effects is a challenging task in solvent extraction chemistry [16,17,18] or separation methods such as adsorption, biosorption, precipitation, reverse osmoses, ion exchange, [19,20,21,22,23,24,25,26], etc.
Furthermore, the innovative non-aqueous solvent extraction process, i.e., solvometallurgy with only several papers written by Prof. Binneman’s group, is available at the moment [27,28,29,30,31,32]. Needless to say, an obvious advantage is the reduced use of water and the avoidance of wastewater treatment problems posterior. As no-aqueous solvent extraction systems often have two liquid organic phases, a distinction needs to be made between the more polar (MP) phase containing the target metal(s) to be extracted, and the less polar (LP) phase, comprising the extractant(s) as well as a diluent, but also in some cases a modifier. Therefore, the substitution of pure water by polar organic compounds used as diluents generally modifies the solvation of metals in this liquid phase, which in some way may alter the metal extraction sequence. Nevertheless, extractant molecules in these liquid−liquid extraction systems still reside in the typical LP phase, ex. CHCl3 or another organic diluent. In general, extractant molecules are rarely used in pure form and are often mixed with an organic diluent. However, diluents may have a significant influence on the extracting behavior due to various possible interactions that may exist between them: dipole–dipole, π-electron, hydrogen bonding, cavity formation, polymerization of the ligand, etc. The formation of a third phase should be avoided; thus, diluents should not form third phases during experimental loading conditions [5].
In fact, the 4f-ions are outlined because they are rare and precious due to their unique mixture of chemical characteristics in combination with their physical properties and specific high-value applications, not to mention their recent high, volatile prices and demands. Thus, Gd(III) ion was selected for this investigation as a typical representative, i.e., situated in the middle of the 4f series. On the other hand, acylpyrazolone ligands have often been used in the field of solvent extraction and separation science of almost all metal ions as they can easily form stable chelate complexes in fairly acidic conditions due to their pKa in the order of four or less [33,34,35,36,37]. However, the complex performance of this class of ligands towards metallic species has not been evaluated in two immiscible organic surroundings.
In this study, the use of switchable hydrophilicity phases is demonstrated mainly with the ethylene glycol compound in the role of a more polar (MP) phase between the two immiscible organic phases of the liquid−liquid extraction system incorporating a chelating agent, 4-benzoyl-3-methyl-1-phenyl-2-pyrazolin-5-one (HP), in order to enhance or tune 4f-ions separations. Scientific results obtained in the comparative study, chloroform was chosen as a representative of widely used organic diluents in solvent extraction chemistry, in order to make a comparison with the data already acquired during years in the field. In addition, a few non-polar molecular organic diluents such as C6H12, n-hexane and n-heptane are also tested for comparison purposes. The results obtained from these scientific tests and the challenges encountered are further discussed in detail, including electron paramagnetic resonance analysis (EPR).

2. Materials and Methods

2.1. Reagents

All reagents were purchased from Aldrich, Merck, and Fluka and were used without further purification. The diluents were 1-butyl-, 1-decyl-3-methylimidazolium-bis(trifluoromethanesulfonyl)imides, (purity, 99.5%, average water content is ca. 200 ppm) purchased from Solvionic (Toulouse, France) and CHCl3 (Merck, p.a.), C6H12 (Merck, p.a.), CH3(CH2)4CH3, i.e., n-hexane (Merck, ≥95%), CH3(CH2)5CH3, i.e., n-heptane (Merck, 99%), ethylene glycol (Merck, 99.5%), glycerol (Thermoscientific, 99+%). The commercial products 4-benzoyl-3-methyl-1-phenyl-2-pyrazolin-5-one (HP, purity > 99%, Fluka) and 2-thenoyltrifluoroacetone (HTTA, Fluka, ≈ 99%) were used as received. All other commercially available analytical-grade reagents were used without any further purification.
Stock solution of gadolinium ion (3 × 10−3 mol/dm3) was prepared from Gd(NO3)3∙6H2O (Fluka, puriss) by dissolving and diluting with distilled water or ethylene glycol to the required volume. Nitric acid, 65% was used (Merck, p.a.) to adjust the pH of the aqueous solutions or MP phase added to 0.1 mol/dm3 2-morpholinoethanesulfonic acid (MES) buffer (Alfa Aesar, 98%).

2.2. Solvent Extraction Studies

Extraction experiments were carried out at room temperature by mixing the two immiscible phases in a 1:1 v/v ratio (1.5 mL) for 2 h (1500 rpm), which was sufficient for attaining equilibrium. After the separation of the liquid phases, the Gd(III) ion concentration in the aqueous or more polar organic phase was determined by using ICP-OES spectroscopy (“Prodigy” High dispersion ICP-OES, Teledyne Leeman Labs, Hudson, NH, USA). The concentration of the metal ion in the organic phase (less polar phase) was obtained by material balance. Extractant solutions in an IL were prepared by precisely weighted samples. The acidity of the aqueous phase at equilibrium was measured by a pH meter (pH 211 HANNA, Woonsocket, RI, USA) with an accuracy of 0.01 pH unit. The Gd(III) ion initial concentration was 6 × 10−4 mol/dm3 in all experiments.
For competitive extraction tests, a volume of 2 mL of the prepared aqueous solution (or no-aqueous compound) containing various Mn+ metal ions (the corresponding nitrate salts were used Mn+(NO3)n/Mn+(NO3)n·xH2O) was equilibrated for 3 h (1500 rpm) with 2 mL organic phase (LP phase), which includes the studied ligand molecule, HP. After phase separation, the metal ion concentrations in the aqueous solution were determined by ICP-OES (“Prodigy” High dispersion ICP-OES, Teledyne Leeman Labs, USA).
The distribution ratio (D) at equilibrium was calculated as follows:
D = [ M n + ] a q , i n [ M n + ] a q , f [ M n + ] a q , f × V a q V I L
where [Mn+]aq,in is the concentration of Mn+ ion in the aqueous phase (MP phase) before liquid–liquid extraction tests, and [Mn+]aq,f is the concentration of the same metal ion in the aqueous phase after extraction. In general, Vaq and VIL are the volumes of aqueous and organic phases as well as the volumes of two immiscible phases used to perform experiments, herein a 1:1 v/v extraction. For instance, duplicate experiments showed that the reproducibility of D measurements was generally within 95%.
Extractability (%E) was evaluated as follows:
e x t r a c t a b i l i t y = [ M n + ] a q , i n [ M n + ] a q , f [ M n + ] a q , i n × 100 .
The metal separation between elements in the periodic table can be estimated using separation factors (SF) determined as a ratio of distribution ratios of two metal ions, namely, the lighter and heavier ones:
SF = D(Z)/D(Z+n).

2.3. EPR Measurements

A Bruker EMX Premium X EPR spectrometer operating in the X-band at 9.4 GHz was used to perform the EPR analysis in the present study. The analysis of the extract solutions at room temperature was carried out with aqueous solution cell ER-160FC-Q, while for the measurements at 100 K a thermoregulatory system ER 4141 VTM was used. The concentration of Gd3+ ion was evaluated using a Bruker licensed software Spin count. As it is known, the Gd3+ ion adopts a 4f7 electronic configuration. In general, the EPR spectrum of Gd3+ displays many components due to transitions between Kramers doublets (ms = ±1/2, ms = ±3/2, ms of = ±5/2 and ms = ±7/2). For Gd3+ located in low-symmetry crystal field, the EPR spectrum is additionally complicated by zero-field splitting parameters (D, E).
All samples provided were recorded in the frozen state at a temperature of 120 K and in the liquid state at 295 K. The measurement conditions of the samples containing Fe and Cr at both temperatures were Modulation Amplitude, MA = 2 mT, and microwave power 1.796 mW (Att = 18). The measurement conditions of the samples containing Cu, at both temperatures, were Modulation Amplitude, MA = 1 mT, and microwave power 1.796 mW (Att = 20).

3. Results and Discussion

3.1. Investigation of Gd(III) Solvent Extraction: Comparison of Molecular Diluent and IL as Well as Aqueous and Non-Aqueous Systems

The main goal of this study is to develop a non-aqueous process for switchable separation of 4f-ions, which is to be competitive over conventional aqueous solvent extraction systems, and to test this process further with other switchable hydrophilicity diluents. It is worth noting that preliminary tests with ethanol amine and ethyl lactate showed the impossibility of being used for these research purposes. Unlike conventional solvent extraction chemistry, the new approach uses two immiscible organic phases: more polar (MP) as well as less polar (LP). The influence of the ligand HP concentration on the extraction of Gd(III) ions was studied with CHCl3/EtG and [C1C10im+][Tf2N]/EtG solutions in order to assess the advantages of non-aqueous chemical process and IL compounds as well as to check the difference in 4f-ion extraction behavior. The results are shown in Figure 1. DP is the distribution ratio due to the gadolinium extraction with HP alone under the experimental conditions, applying various concentrations of the chelating ligand. It is evident that when EtG solutions are used, the organic diluent CHCl3 somehow clearly excels the IL compound [C1C10im+][Tf2N]. In fact, the opposite picture is usually seen, i.e., ionic liquids outperforming typical organic diluents during metal solvent extraction. It should be noted that all physical and chemical diluent properties, including gaining more insight into the solvation of metal ions in a non-aqueous environment, contribute to the extraction mechanism, resulting in higher efficiency for a similar molecular system. The obtained two plots for HP concentration (when two different, less polar organic phases were applied to CHCl3 and the hydrophobic IL) give a good linear log–log confidence with a slope of 0.48 but strongly deviate from the value of one. This means that no more than one molecule of the acidic ligand is involved in the extraction process under the applied experimental conditions and the rest of the charge of Ln3+ ion could be compensated probably by [Tf2N] anions present in the organic phase [38,39,40,41]. It is very likely that the reaction mechanism in both cases is the same and it can be said that there is no transfer of the IL cation ([C1C10im+]) in the second organic phase, i.e., ethylene glycol:
Gd3+(MP) + ½ HP(LP) + n[Tf2N](LP) ⇌ Gd(P)½z+(Tf2N)n(LP) + H+(MP)
where the subscripts “MP” and “LP” denote the two organic phases, more and less polar, respectively.
In order to investigate the role of ionic liquid on the traditional solvent extraction process, a set of experiments was performed under conditions of the aqueous phase comparable with those carried out in CHCl3. The distinct extraction behavior of the corresponding two systems towards Gd(III) ions was described in Figure 2. As a whole, the higher extraction efficiency of various extractants diluted in ILs compared to traditional molecular diluents is a common feature and different stoichiometries of the extracted metal species are usually, but not always, found [10,34,41,42]. Therefore, as expected, the solvent extraction of Gd(III) ion increased rapidly when CHCl3 was changed with the IL compound. Furthermore, as shown in Figure 2, linear relationships between logD and log[HP] with a slope close to one (1.05 (IL) and 0.98 (CHCl3)) were obtained for investigated two systems. Consequently, the possible extraction mechanism can be elucidated by applying the following equations:
Gd3+(aq) + HP(o) + n[C1C10im+][Tf2N](o) ⇌ Gd(P)2+(Tf2N)n(o) + H+(aq) + n[C1C10im+](aq)
Gd3+(aq) + HP(o) + ⇌ Gd(P)2+(o) + H+(aq)
where the subscripts “aq” and “o” denote aqueous and organic phases, respectively.
It is certain that under other experimental conditions, i.e., changing the acidity of the aqueous phase would produce a different result for the system involving chloroform, usually with the participation of three or four molecules of the chelating ligand in the reaction mechanism [43]. However, several general conclusions can be drawn from the information obtained herein. First, when using an ionic liquid, water is preferable to be used as the second phase containing the target metal for solvent extraction, rather than ethylene glycol. Thus, in other words, unsatisfactory results were obtained precisely with this second organic phase compared to water. Second, as a whole, the exact opposite trend can be seen with the use of CHCl3 as a less polar phase; i.e., in an ethylene glycol medium, the metal extraction is much more efficient than when water is applied as a medium (MP).
Another noteworthy point is that the process is not so affected by pH (0.1 mol/dm3 MES and HNO3/NaOH):
(a) The solvent system EtG/[C1C10im+][Tf2N] has a pH in the range of 1.60 to 4.45 at [HP] = 2.5 × 10−2 mol/dm3 produce logD~0.52;
(b) The H2O/[C1C10im+][Tf2N] system has a pH in the range of 1.90 to 3.04 at [HP] = 4 × 10−2 mol/dm3, leading to logD > 1.4; i.e., 100% extraction was achieved.
What happens when the diluent is chloroform in combination with water (1.855 D) and ethylene glycol (EtG has a zero-dipole moment) can be seen in Figure 3. The solvent system in used EtG/CHCl3 is superior to H2O/CHCl3.
Another interesting fact, however, is that a similar trend is not observed when non-polar diluents are used instead of investigated CHCl3 (dipole moment, 1.15 D) such as C6H12 (0.0 D), n-hexane (0.08 D), or n-heptane (0.0 D). From the experiments carried out with various concentrations of the chelating ligand (HP), it can be seen that the aqueous phase gives extremely good results and as a consequence is preferable (MP) than ethylene glycol in these cases, Figure 4. Thus, the superiority of the non-aqueous solvent system such as EtG/n-heptane is emphasized compared to H2O/n-heptane, changing the pH values, as shown in Figure 5.

3.2. Solvent Extraction and Selectivity across the Periodic Table and 4f Series

It is a known fact in coordination chemistry that different metal ions show a different affinity towards a given specific extractant molecule, so they will probably distribute differently over both immiscible liquid phases and metal separation can be easily achieved. The competitive solvent extraction test of almost 25 metal ions by the chelating ligand HP diluted in two ILs and CHCl3 has also been conducted in order to evaluate the switchable hydrophilic diluent, EtG as a more polar liquid phase, Figure 6.
In summary, the studied 25 metal ions could be classified into the two groups listed below, according to the degree of HP extraction performance in an IL media applying EtG:
(1)
Unextractable metal ions, %E < 1: Li+, Na+, K+, Ca2+, Mg2+, Ba2+, Sr2+, Pb2+, Zn2+, Cd2+, Cr3+, Co2+, La3+, Ce3+;
(2)
Overall extraction %E ≤ 50: Al3+ as well as Ni2+ and Hg2+ with only [C1C4im+][Tf2N] as a diluent.
Overall extraction %E > 50: Fe3+, Cu2+, Tl+, Bi3+, Lu3+, while Ag+, Eu3+, Gd3+ only with the participation of the [C1C4im+][Tf2N] compound in the system.
Therefore, metal cations such as Hg2+ and Ni2+ could be removed from the ion mixture by applying only 1 butyl homolog of the imidazolium IL, and, at the same time, it is not at all possible to extract them with the more hydrophobic IL [C1C10im+][Tf2N]. What is noticeable about the f-elements is the obtained zero extraction for lanthanum and cerium, which can be explained by the relatively low acidity of the environment. If we compare the two ionic compounds, a negligible difference is observed for Ag+, Eu3+, and Gd3+ ions, about two times lower solvent extraction with [C1C10im+][Tf2N] compared to its short-chain analog. When using chloroform as a diluent, the picture is the same in parts above or equal to 50% extraction, but only for d-elements, i.e., Fe3+, Cu2+, Tl+, and Bi3+. Unfortunately, lanthanoids are not extracted under these experimental conditions by applying EtG/CHCl3. A lack of affinity towards the chelating agent HP in this organic medium was observed for all metal cations examined, with one exception: silver ions. For the first time, the advantage of ionic liquids as an extractant diluent leading to better extraction, which can be used successfully in the separation chemistry for various metal cations, has been observed herein, of course in combination with EtG, i.e., non-aqueous processes.
When using the immiscible system H2O/ILs, the grouping of metal ions is as follows according to the data presented in Figure 6:
(1)
Unextractable metal ions, %E < 1: Li+, Na+, K+, Ca2+, Mg2+, Ba2+, Sr2+, Zn2+, Cd2+, Co2+, Ni2+, Ce3+;
(2)
Overall extraction %E ≤ 50: Cr3+, Pb2+, Al3+ but Hg2+ and La3+ only when the [C1C4im+][Tf2N] compound is a diluent;
(3)
Overall extraction %E > 50: Fe3+, Cu2+, Tl+, Bi3+, Eu3+, Gd3+, Lu3+, but Ag+ only reacts with the [C1C4im+][Tf2N] compound.
Of interest is the chemical element lead, which, although below 5%, it is extracted, while in an ethylene glycol medium, it is not. The same is true for the lighter member of the 4f series, La3+. Under certain experimental conditions, however, this simplification is perhaps unjustified. Understanding could be achieved by analyzing the exact parameters of the process. The water environment turns out to be favorable for another metal cation, such as chromium, Cr3+, but not with the IL having n = 4 ([C1Cnim+][Tf2N]), so the hydrophobic diluents are favorable in this case, i.e., n = 10 and CHCl3. Again, the advantage of the [C1C4im+][Tf2N] compound for both solvent extraction and separation processes of metal cations is worth noting, and this is particularly evident for ions such as Al3+, Ag+, Eu3+, and Lu3+. On the other hand, the advantage of ILs has been demonstrated for the umpteenth time in this water system as well, in comparison to the organic molecular diluent CHCl3. Among the lanthanoids, only Gd3+ is extracted at ~3%, while the s-ion, Na+ is about 24%. More than 50% extraction efficiency is observed as usual in similar studies for Fe3+, Cu2+, and the heavy Bi3+ [44,45], while less than 50% solvent extraction from water to chloroform medium could be seen for Al3+, Pb2+, and Cr3+.
It is a known fact that the good selectivity of metal ions during liquid−liquid extraction process largely relies on the affinity of metals to the ligands, which normally reside in the organic (less polar) phase because of their high hydrophobicity. We follow a different route herein to apply a typical and well-known chelating extractant in coordination chemistry such as 4-acylpyrazolones in order to see the influence of the organic (more polar) phase on separation chemistry. The first thing that is noticeable from the obtained selectivity data is the absence of the solvent extraction process for nickel in aqueous environments (see Tables S1 and S2 in the Supplementary Material). However, the calculated values for the separation coefficient of nickel with all heavier ions are extremely high when using ethylene glycol. For example, the SFs for Ni/Cu pair is 324, while for the Ni/Gd pair, this value is ca. 211. This is also a characteristic feature of Al3+, whose selectivity is more pronounced in EtG than in water. However, the opposite is observed for Fe3+, Hg2+, as well as Cu2+ in a non-aqueous environment, wherein their separation from other metal ions is extremely weak. The calculated SF values for Fe/Hg and Cu/Hg pairs are 1 × 10−2 and 7 × 10−3, while in a water environment, these values reach up to 525 and 88, respectively. It should be emphasized that lanthanoids, in general, have a poorer separation with Hg2+ in EtG (approximately over 1000 times). For instance, it is very difficult to draw a similar conclusion for the ion Tl+ and Bi3+ as well, but water seems to be the preferred medium, like a polar phase, in this case. What can be said about the 4f-ions is that they are better separated into EtG. In other words, a conclusion that deserves further research investigations of different non-aqueous phases. However, the 4f separation from heavier ions such as Hg2+, Bi3+ and Tl+ is likely much better in the system H2O//[C1C4im+][Tf2N], in contrast to EtG//[C1C4im+][Tf2N]. However, the results indicate that 4f-ions were recovered efficiently from the aqueous phase. Thus, the employment of the EtG compound, i.e., a non-aqueous solvent extraction system, appears to be a versatile scientific approach when the intragroup 4f selectivity is searched for.
Moreover, since ethanol amine, ethyl lactate, and glycerol did not prove to be suitable to replace water 100%, and the last compound was not able to dissolve a sufficient amount of MES to achieve the desired concentration, a test was carried out to investigate the influence of glycerol in a ratio of 1 to 1 with water (v/v). The obtained results are presented in Figure 7. It can be concluded on the basis of obtained results that the influence is particularly pronounced in a very positive way, especially referring to the solvent extraction of s-elements, something that was not noticeable in the previous solvent system combinations, as well as for d-elements such as Co, Ni, Zn, and Cd. In addition, the typical good extraction of metals such as Fe, Hg, Al, Tl, Bi, and, of course, the 4f series is achieved again. If we look at a Table S3 (Table S4) in the Supplementary material, where the separation coefficients for the studied metal ions from the periodic table, whole number 25, and the various ion pairs are placed, it can be seen that the selectivity in the 4f series is reasonably good above ca. 0.3. The high values of SFs calculated for Zn and Sr with other metals ranging from Li to Cu must be noted. The high values obtained for SFs between cobalt and lanthanoids could be useful and could be further used in recycling processes, while it is low for nickel and Lns [46]. We cannot skip the high selectivity obtained also for K and Al ions as forming pairs with other chemical elements with an atomic number after 29, i.e., Cu. In general, the obtained data herein show that the addition of this compound, glycerol, to the aqueous phase can be cleverly exploited in the future in separation chemistry.
Further, from the obtained results, general rules mainly within the IL phase [C1Cnim+][Tf2N] could be derived: first, whatever the extracting compound, its ability in the IL phase is always largely above that in the traditional organic diluent. However, as a whole, the chelating compounds do not provide improved SFs between two lanthanoids in an IL media as compared to molecular diluents, i.e., CHCl3. On that account, the results from solvent extraction are clear advantages of the IL media compared to molecular, while the selectivity might be considered likely as rather disappointing at first glance. Nevertheless, it is worth explaining that the extraction ability and selectivity are almost always a pair of contradictory elements (videlicet, frequently high extraction efficiency was not accompanied by marked selectivity with any change in the solvent system, i.e., the nature of the two liquid phases or extracting moiety). This statement is demonstrated once again herein based on the data obtained from various investigated solvent systems, as shown in Figure 8.
Hence, the type of the used diluent usually has only a pronounced qualitative effect on the solvent extraction efficiency of 4f-ions without a noticeable influence to such a degree on the reaction mechanism, Figure S1. For a given metal, the values of the equilibrium constant increase in the following order CHCl3 < C6H6 < C2H4Cl2 < CCl4 < C6H12 [51,52,53]. In many cases, spectrophotometry shows that the coordination of the metal chelate is invariant from one traditional organic diluent to another. As a starting assumption, the molecular diluent is usually considered inert, since it does not appear further in the extraction equilibrium expression, unlike ILs. It may be concluded that, as a whole, the increase in the diluent solvating ability hinders the extraction process [54]. Thus, it seems that the diluent effect on solvent extraction follows the general schematic behavior described in the solvation process of a solute in a liquid. Nevertheless, as a whole, based on our previous investigations and other research studies, while the selectivity decreases, the lanthanoid solvent extraction increases in the following order: β-diketone ˂ 4-acylpyrazolone ˂ 4-acyl-3-phenyl-5-isoxazolone. Another appropriate way to evaluate the solvent extraction power of the chelating ligands is to compare the corresponding values of pH50 [44]. Evidently, the values of pH50 for the liquid–liquid extraction of 4f-elements obtained when the chelating compound is the typical β-diketone, thenoyltrifluoroacetone (HTTA) are more than one pH unit larger compared to when 4-acylpyrazolone is in use, as well as approximately three pH units larger compared with 4-benzoyl-3-phenyl-5-isoxazolone reagents (see Figure S1). In addition, the difference of calculated ∆pH50 values decreases with increasing the atomic number of the investigated lanthanoids. Moreover, the sequence of the pH50 is in the following order: Lu < Ho < Eu < Nd < La. Consequently, the value of pH50 rises as the ionic radius of the lanthanoid element increases, as shown in Figure S2.

3.3. EPR Investigation of the Extracting LP Phases

In this investigation concerning the complexation of Gd(III) ion in the organic phase (LP), two chelating ligands have been chosen for parallel EPR analysis 4-benzoyl-3-methyl-1-phenyl-2-pyrazolin-5-one (HP) as well as thenoyltrifluoroacetone (HTTA). The data in Figure 9 indicates that the EPR spectrum of solutions c (ligand—HTTA in CHCl3) consists of a slightly asymmetric signal with the following EPR parameters: g = 1.99 and ∆Hpp = 3.0 mT. In samples a and b, in which the ligand HTTA was used for complexation, respectively, in [C1C4im+][Tf2N] and [C1C10im+][Tf2N], signals with g ≈ 1.99 were also detected; however, they were extremely weak (Figure 9, narrow range). The spectra in a wide range of the magnetic field (Figure 9, left) do not show additional signals of interest. The signal with g ≈ 1.99 corresponds to a transition between energy states with ms −1/2 ↔ + 1/2 and is attributed to Gd3+ located in a ligand field with high symmetry. In order to clear up the origin of all registered signals in the spectrum, we have to mention that the asterisk-marked signals are probably due to the well-resolved hyperfine structure of Cu2+ ions found as impurities in ILs.
The wide magnetic field EPR spectra of solutions one, two, and three (HP as a ligand, respectively, in [C1C4im+][Tf2N], [C1C10im+][Tf2N], and CHCl3) consist of multiple signals located in the range from 0 to 600 mT and at least nine lines are clearly visible. In Figure 9, the effective values of the g-factors of the observed signals are indicated within a wide range. These signals could be assigned to the Gd3+ fine structure, as the existence of some forbidden transition could not be excluded. The presence of multiple signals with g >> 2.0 and g < 2.0 is associated with Gd3+ ions in a low-symmetry crystal field with large zero-field splitting parameters (D~hv) [55]. In the spectra registered in the narrow magnetic field range (300–360 mT), the signal with g = 1.98 could be seen as a constituent of the fine structure. A signal with g = 2.01 and ∆Hpp = 0.9 mT was also registered, on both sides of which two symmetrically located less intense signals were observed. Most likely, the narrow signal with g = 2.01 and the two satellite lines originate from the presence of a free radical in the ligand. Using the SpinCount program, part of the Bruker Xenon software, concentrations of Gd3+ ions in the investigated frozen solutions at 120 K were found (see Table S5). Results were obtained directly from the analysis without further experiments and quantitative EPR sample measurements such as those described by Yordanov in 1994 [56]. Thus, despite that the same analyses are performed with a higher concentration of the chelating extractants, In general, there is no particular change in the obtained spectra, although the amount of extracted gadolinium is different, as shown in Figure S4 and Table S6 (see as well Figure S3 for information).
On the other hand, from the conducted tests of simultaneous extraction of different metal ions, the established relatively high extraction of copper and iron is really striking. So, this led to the thought of studying the corresponding organic phases obtained after their solvent extraction using the EPR technique. As the reader can see, Figure 10 shows the EPR spectra of Cu2+ complexes extracted from EtG solution with HP dissolved in two iLs (n = 4 and 10) as well as their simulations. The EPR spectra were recorded both in a liquid state at room temperature and in frozen solutions at 100 K. As seen in Figure 10, all recorded EPR spectra show similar signals composed of parallel and perpendicular components with g > g. Such spectra are usually attributed to Cu2+ ions in tetragonally distorted octahedral symmetry, with an unpaired electron occupying the dz2 orbital (dx2−y2 ground state of Cu2+). The frozen solutions spectra are characterized with clearly resolved hyperfine structure (hfs) lines in parallel and perpendicular components, while liquid spectra show only resolved hfs lines in their parallel part. The performed simulation reveals the existence of hyperfine structure lines that belong to two kinds of Cu2+ ions in slightly different environments (Table 1). The described spectra with well-separated hfs lines are characteristic of the presence of well-isolated Cu2+ ions. In addition, no substantial differences in parameters for liquid and frozen samples were established. Taking into account the EPR results, we could summarize that Cu2+ ions possess similar coordination with HP ligand, regardless of the used ionic diluent ([C1C4im+][Tf2N] or [C1C10im+][Tf2N]). The quantitative analysis performed on the frozen solutions showed that the concentration of Cu2+ ions in the investigated quantity of LP phases [C1C4im+][Tf2N] and [C1C10im+][Tf2N] are 2.00 × 10−3 mol/dm3 and 1.17 × 10−3 mol/dm3, respectively.
Further, the EPR spectra of the corresponding frozen samples containing Fe3+ are presented in Figure 11. As seen, the two spectra consist of a signal with g~4.23 and line width ∆Hpp = 15 mT. Such a signal is characteristic of isolated Fe3+ ions located in highly distorted symmetry. The Fe3+ concentrations calculated from the EPR analysis of frozen solutions for Fe3+ applying both ILs (n = 4 and n = 10) are 3.7 × 10−4 mol/dm3 and 4.8 × 10−4 mol/dm3, respectively.
Since chromium (III) compounds (d3) are paramagnetic (S = 3/2), EPR spectroscopy provides an additional analytical method. However, the attempt to obtain an EPR spectrum of Cr3+ complexes with HP ligand extracted in both ionic liquids brought about difficulties. Both registered spectra are shown in Figures S5 and S6 (Supplementary material). In fact, the EPR spectrum of Cr3+ in EtG consists of several lines that could be set in two groups according to their line width: one set of narrower lines (g = 5.7, 2.08, 1.92, 1.98) (∆Hpp = 5 mT) and one set of broader signals (g = 4.7, 2.4, 2.2, 1.5) (Figure S7). The narrow line set is ascribed to well-isolated Cr3+ ions, while the broad line set is ascribed to Cr3+ ions that experience some exchange interaction. In the central part of the spectrum (g~1.98) is visible a broad signal of exchanged coupled Cr3+ ions, on which are superimposed the narrow signals with g = 2.08, 1.92, 1.98. Taking into account the standard spectrum (Figure S7), it seems that the signals with g = 5.7, 2.4, and 2.2 in samples that have undergone extractions into IL could be related to Cr3+ ions in ligand fields with distorted symmetry and large zero-field splitting constants. However, their extremely low intensities should be outlined.

4. Conclusions

An environmentally sound solvent extraction process of Gd(III) ions was developed in the present study, changing the chemical nature of the two liquid phases. In order to improve the process efficiency, the substitution of typical organic diluents with ILs as well as pure water with ethylene glycol or partly with glycerol (1:1) was considered. Overall, the results obtained from the investigation are as follows:
1. As a general rule, our results indicate that “green media” such as ILs have fantastic solvent extraction potential, which usually results in less selectivity. This phenomenon can be changed by changing the nature of the opposite liquid phase: replacing the water.
2. All 4f-ions could be recovered in higher yield when HP is used as an extractant with ILs combined with both MP organic phases: aqueous or non-aqueous. However, the process is not so efficient with the CHCl3/H2O solvent system. When EtG solutions are used, the organic diluent CHCl3 somehow clearly expels the IL compound [C1C10im+][Tf2N].
3. The metals with the highest competitive impact on rare earth recovery from mixed e-waste extracted by chelating ligand HP are almost always Fe, Al, Cu, and Ag, while Pb could only be extracted from the aqueous phase.
4. Improvement of the solvent extraction process was observed with a small addition of glycerol (1:1, v/v) in the aqueous phase. After this ploy, the elements whose competitive effect increased the most are those in two s-blocks, Co and Ni. Therefore, this switchable separation is relevant to the recycling demands nowadays. In other words, the leveling of separation factors as obtained usually in IL mediums, at values around or below 1, together with the very high distribution ratios normally seen, would be an interesting property of such diluents to be further explored in non-aqueous solvent systems.
5. Unfortunately, EPR analysis could give good informative results in solvent extraction chemistry when the process is efficient, i.e., only above 50–60% metal extraction.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/separations10050286/s1, Figure S1: LogK vs. Z.; Figure S2: pH50 in the solvent extraction of lanthanoids with HL in CHCl3 as a function of the ionic radius. Figure S3: EPR spectra of Gd3+ ion in ethylene glycol at room temperature and in the frozen state at 120 K: [Gd3+] = 3 × 10−3 mol/dm3. Figure S4: EPR spectra of frozen solutions of complexes of Gd3+ ([Gd3+]in = 6 × 10−4 mol/dm3) with ligands [HP] = 2.5 × 10−2 mol/dm3 (A, B, and C) and [HTTA] = 3 × 10−2 mol/dm3 (1, 2 and 3) obtained after solvent extraction from EtG, respectively, in the following diluents [C1C4im+][Tf2N] (A: pHeq = 1.98 and 1: pHeq = 2.60), ([C1C10im+][Tf2N] (B: pHeq = 2.02 and 2: pHeq = 2.57), CHCl3, (C: pHeq = 1.98 and 3: pHeq = 2.21). On the left: the range of the magnetic field (0−700 mT). On the right: range of the magnetic field (300−360 mT). Figure S5: EPR spectra of Cr3+ obtained after solvent extraction with HP = 2 × 10−2 mol/dm3 in [C1C4im+][Tf2N] and [C1C10im+][Tf2N] recorded in frozen solutions: MP phase—[Cr3+]in = 6 × 10−4 mol/dm3 and 0.1 MES in EtG. Calculated %E is 71.23% and 54.60%—ICP-OES based on an average of three parallel determinations. Figure S6: EPR spectra of Cr3+ obtained after solvent extraction with HP = 2 × 10−2 mol/dm3 in [C1C10im+][Tf2N]//EtG (1), [C1C4im+][Tf2N]//EtG (2) and [C1C4im+][Tf2N]//H2O (3) recorded in frozen solutions: MP phase—[Cr3+]in = 1 × 10−3 mol/dm3 and 0.1 MES in EtG. Calculated %E is 99.04% (1) and 99.57% (2) and below 0.5% (3)—ICP-OES based on an average of three parallel determinations. Figure S7: EPR spectra of Cr3+ ion in ethylene glycol at room temperature and in the frozen state at 120 K: [Cr3+] = 3 × 10−3 mol/dm3. The Cr3+ concentration in the so-called “standard sample” was found to be about 4 × 10−4 mol/dm3 (the spectra were integrated from 0 to 500 mT.). Table S1: Separation factors of Mn+ metal ions extracted with HP = 2 × 10−2 mol/dm3 ligand in [C1C4im+][Tf2N] applying EtG and H2O. Table S2: Initial concentrations of metal ions used in competitive extraction after 7 and 6.3-fold dilution. Table S3: Separation factors of Mn+ metal ions extracted with HP = 2 × 10−2 mol/dm3 ligand in [C1C10im+][Tf2N] applying H2O and glycerol, 1:1. Table S4: Initial concentrations of metal ions used in competitive extraction test applying H2O and glycerol, 1:1. Table S5: Concentrations of the investigated Gd3+ ions using the SpinCount program, part of the Bruker Xenon software. Table S6: Concentrations of the investigated Gd3+ ions using the SpinCount program, part of the Bruker Xenon software.

Author Contributions

Conceptualization, M.A.; methodology, M.A.; validation, M.A. and R.K.; investigation, M.A. and R.K.; writing—original draft preparation, M.A.; writing—review and editing, M.A.; visualization, M.A. and R.K.; project administration, M.A.; funding acquisition, M.A. All authors have read and agreed to the published version of the manuscript.

Funding

The research leading to those results received funding from the Bulgarian National Science Fund: Grant Agreement NO. KП-06-H69/5(2022), “Green twist to synergistic solvent extraction and separation of rare earth metals”.

Data Availability Statement

Data sharing is not applicable to this article.

Acknowledgments

M. Atanassova personally thanks the European Union-NextGenerationEU, through the National Recovery and Resilience Plan of the Republic of Bulgaria, project NO. BG-RRP-2.004-0002-C01, “BiOrgaMCT”. The authors thank eng. Nina Todorova, as well as the BSc students Katrin Chavdarova, Ilina Todorova, and Aleyna Chamdere for their contribution to some experiments that triggered this study, and the authors would also like to acknowledge eng. Stoyko Petrin, for performing the ICP-OES measurements.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. IUPAC. Gold Book, Compendium of Chemical Terminology, 2nd ed.; IUPAC: Zurich, Switzerland, 2019; Available online: https://goldbook.iupac.org (accessed on 1 April 2023).
  2. Löfström-Engdahl, E.; Ekberg, C.; Foreman, M.; Skarnemark, G. Diluents effect. In Proceedings of the First ACSEPT International Workshop, Lisbon, Portugal, 31 March–2 April 2010. [Google Scholar]
  3. Shimojo, K. Solvent extraction in analytical separation techniques. Anal. Sci. 2018, 34, 1345. [Google Scholar] [CrossRef]
  4. Healy, T. Synergism in the solvent extraction of di-, tri- and tetravalent metal ions—II: Synergic effects in so-called inert diluents. J. Inorg. Nucl. Chem. 1961, 19, 328. [Google Scholar] [CrossRef]
  5. Petrova, M. The Crucial Performance of Mutual Solubility among Water and Ionic Liquids in the Time of Liquid-Liquid Extraction of Metallic Species; Academic Solutions: Sofia, Bulgaria, 2020. [Google Scholar]
  6. El-Nadi, Y.A. Solvent extraction and its applications on ore processing and recovery of metals: Classical approach. Sep. Purif. Rev. 2017, 46, 195. [Google Scholar] [CrossRef]
  7. Jose Ruiz-Angel, M.; Carda-Broch, S. Recent advances on ionic liquid uses in separations techniques. Separations 2022, 9, 96. [Google Scholar] [CrossRef]
  8. He, J.; Tao, W.; Dong, G. Study on extraction performance of vanadium(V) from aqueous solutions by octyl-imidazole ionic liquids extractants. Metals 2022, 12, 854. [Google Scholar] [CrossRef]
  9. Feng, X.; Song, H.; Zhang, T.; Yao, S.; Wang, Y. Magnetic technologies and green solvents in extraction and separation of bioactive molecules together with biochemical objects: Current opportunities and challenges. Separations 2022, 9, 346. [Google Scholar] [CrossRef]
  10. Atanassova, M. Solvent extraction of metallic species in ionic liquids: An overview of s-, p- and d-elements. J. Chem. Technol. Met. 2021, 56, 443. [Google Scholar]
  11. Mohapatra, P.K. Actinide ion extraction using room temperature ionic liquids: Opportunities and challenges for nuclear fuel cycle applications. Dalton Trans. 2017, 46, 1730. [Google Scholar] [CrossRef]
  12. Dietz, M. Ionic liquids as extraction solvents: Where do we stand? Sep. Sci. Technol. 2006, 41, 2047. [Google Scholar] [CrossRef]
  13. Okamura, H.; Hirayama, N. Recent Progress in Ionic Liquid Extraction for the Separation of Rare Earth Elements. Anal. Sci. 2021, 37, 119. [Google Scholar] [CrossRef]
  14. Turanov, A.; Karandashev, V.; Yarkevich, A.; Khsostikov, V. Extraction of REE(III), U(VI), and Th(IV) from nitric acid solutions with diphenyl(dibutylcarbamoylmethyl)phosphine oxide in the presence of quaternary ammonium bis[(trifluoromethyl)sulfonyl]imides. Radiochemistry 2019, 61, 694. [Google Scholar] [CrossRef]
  15. Lebedeva, O.; Kultin, D.; Zakharov, A.; Kustov, L. Advantages of electrochemical polishing of metals and alloys in ionic liquids. Metals 2021, 11, 959. [Google Scholar] [CrossRef]
  16. Turanov, A.N.; Matveeva, A.G.; Kudryavtsev, I.Y.; Pasechnik, M.P.; Matveev, S.V.; Godovikova, M.I.; Baulina, T.V.; Karandashev, V.K.; Brel, V.K. Tripodal organophosphorus ligands as synergistic agents in the solvent extraction of lanthanides(III). Structure of mixed complexes and effect of diluents. Polyhedron 2019, 161, 276. [Google Scholar] [CrossRef]
  17. Sun, Z.; Zheng, L.; Zhang, Z.-Y.; Cong, Y.; Wang, M.; Wang, X.; Yang, J.; Liu, L.; Huai, Z. Molecular modelling of ionic liquids: Situations when charge scaling seems insufficient. Molecules 2023, 28, 800. [Google Scholar] [CrossRef]
  18. Dukov, I.; Atanassova, M. Effect of the diluents on the synergistic solvent extraction of some lanthanides with thenoyltrifluoroacetone and quaternary ammonium salt. Hydrometallurgy 2003, 68, 89. [Google Scholar] [CrossRef]
  19. Khamseh, A.; Ghorbanian, S. Experimental and modeling investigation of thorium biosorption by orange peel in a continuous fixed-bed column. J. Radioanal. Nucl. Chem. 2018, 317, 871. [Google Scholar] [CrossRef]
  20. Merino-Garcia, I.; Velizarov, S. New insights into the definition of membrane cleaning strategies to diminish the fouling impact in ion exchange membrane separation processes. Sep. Purif. Technol. 2021, 277, 119445. [Google Scholar] [CrossRef]
  21. Chen, F.; Liu, F.; Wang, L.; Wang, J. Comparison of the preparation process of rare earth oxides from the water leaching solution of waste Nd-Fe-B magnets’ sulfate roasting products. Processes 2022, 10, 2310. [Google Scholar] [CrossRef]
  22. Zhang, J.; Xue, N.; Zhang, Y.; Zheng, Q. High-efficiency stepped separation and recoveries of vanadium and molybdenum via low-temperature carbonation conversion of high-chromium vanadium residue. Processes 2023, 11, 470. [Google Scholar] [CrossRef]
  23. Murtaza, G.; Ahmed, Z.; Eldin, S.; Ali, I.; Usman, M.; Iqbal, R.; Rizwan, M.; Abdel-Hameed, U.; Haider, A.; Tariq, A. Biochar as a green sorbent for remediation of polluted soils and associated toxicity risks: A critical review. Separations 2023, 10, 197. [Google Scholar] [CrossRef]
  24. Pyrzynska, K. Recent applications of carbon nanotubes for separation and enrichment of lead ions. Separations 2023, 10, 152. [Google Scholar] [CrossRef]
  25. Tatsumi, T.; Tahara, Y.; Matsumoto, M. Adsorption of metallic ions on amidoxime-chitosan/cellulose hydrogels. Separations 2021, 8, 202. [Google Scholar] [CrossRef]
  26. Admawi, H.; Mohammed, A. A comprehensive review of emulsion liquid membrane for toxic contaminants removal: An overview on emulsion stability and extraction efficiency. J. Environ. Chem. Eng. 2023, 11, 109936. [Google Scholar] [CrossRef]
  27. Binnemans, K.; Jones, P. Solvometallurgy: An emerging branch of extractive metallurgy. J Sustain. Metall. 2017, 3, 570. [Google Scholar] [CrossRef]
  28. Wellens, S.; Thijs, B.; Möller, C.; Binnemans, K. Separation of cobalt and nickel by solvent extraction with two mutually immiscible ionic liquids. Phys. Chem. Chem. Phys. 2013, 15, 9663–9669. [Google Scholar] [CrossRef] [PubMed]
  29. Li, Z.; Dewulf, B.; Binnemans, K. Nonaqueous solvent extraction for enhanced metal separations: Concept, systems, and mechanisms. Ind. Eng. Chem. Res. 2021, 60, 17285. [Google Scholar] [CrossRef] [PubMed]
  30. Batchu, N.K.; Dewulf, B.; Riano, S.; Binnemans, K. Development of a solvometallurgical process for the separation of yttrium and europium by Cyanex 923 from ethylene glycol solutions. Sep. Purif. Technol. 2020, 235, 116193. [Google Scholar] [CrossRef]
  31. Dewulf, B.; Riano, S.; Binnemans, K. Separation of heavy rare-earth elements by non-aqueous solvent extraction: Flowsheet development and mixer-settler tests. Sep. Purif. Technol. 2022, 290, 1220882. [Google Scholar] [CrossRef]
  32. Rizk, S.; Gamal, R.; El-Hefny, N. Insights into non-aqueous solvent extraction of gadolinium and neodymium from ethylene glycol solution using Cyanex 572. Sep. Purif. Technol. 2021, 275, 119160. [Google Scholar] [CrossRef]
  33. Marchetii, F.; Rettinari, C.; Pettinari, R. Acylpyrazolone ligands: Synthesis, structures, metal coordination chemistry and applications. Coor. Chem. Rev. 2005, 249, 2909. [Google Scholar] [CrossRef]
  34. Turanov, A.N.; Karandashev, V.K.; Baulin, V.E.; Baulin, D.V.; Tsivadze, A.Y. Synergistic solvent extraction of lanthanides(III) with mixtures of 4-benzoyl-3-methyl-1-phenyl-5-pyrazolone and phosphoryl-containing podands. Acta Chim. Slovenica 2020, 67, 246. [Google Scholar] [CrossRef]
  35. Turanov, A.N.; Karandashev, V.K.; Baulin, D.V.; Baulin, V.E.; Tsivadze, A.Y. Solvent extraction of lanthanides(III) with mixtures of 1,3,5-tris(2-diphenylphosphoryl-4-ethylphenoxymethyl)benzene and 4-benzoyl-5-methyl-2-phenyl-3-pyrazolone. Russ. Chem. Bull. 2021, 70, 2416. [Google Scholar] [CrossRef]
  36. Turanov, A.N.; Karandashev, V.K.; Baulin, D.V.; Baulin, V.E.; Tsivadze, A.Y. Extraction of rare earth elements(III) with mixtures of 1-phenyl-3-methyl-4-benzoyl-5-pyrazolone and 2-phosphorylphenoxyacetamides. Russ. J. Inorg. Chem. 2019, 64, 407. [Google Scholar] [CrossRef]
  37. Turanov, A.N.; Karandashev, V.K.; Khvostikov, V.A.; Artyushin, O.I.; Bondarenko, N.A. Extraction of rare earth elements(III) with mixtures of some new tridentate carbamoylmethylphosphine oxides and 4-benzoyl-3-methyl-1-phenyl-1H-pyrazol-5(4H)-one from hydrochloric acid solutions. Russ. J. of Gen. Chem. 2021, 91, 383. [Google Scholar] [CrossRef]
  38. Jensen, M.; Borkowski, M.; Laszak, I.; Beitz, J.; Rickert, P.; Dietz, M. Anion effects in the extraction of lanthanide 2-thenoyltrifluoroacetone complexes into an ionic liquid. Sep. Sci. Technol. 2012, 47, 233. [Google Scholar] [CrossRef]
  39. Okamura, H.; Takagi, H.; Isomura, T.; Morita, K.; Nagatani, H.; Imura, H. Highly selective synergism for the extraction of lanthanoid(III) ions with β-diketones and trioctylphosphine oxide in an ionic liquid. Anal. Sci. 2014, 30, 323–325. [Google Scholar] [CrossRef]
  40. Rama, R.; Rout, A.; Venkatesan, K.; Antony, M.; Vasudeva Rao, P. Extraction behavior of americium(III) in benzoylpyrazolone dissolved in pyrrolidinium based ionic liquid. Sep. Sci. Technol. 2015, 50, 2164. [Google Scholar] [CrossRef]
  41. Atanassova, M. Worthy extraction and uncommon selectivity of 4f-ions in ionic liquid medium: 4-acylpyrazolones and CMPO. ACS Sustain. Chem. Eng. 2016, 4, 2366. [Google Scholar] [CrossRef]
  42. Atanassova, M. Solvent extraction chemistry in ionic liquids: An overview of f-ions. J. Mol. Liq. 2021, 343, 117530. [Google Scholar] [CrossRef]
  43. Atanassova, M.; Okamura, H.; Eguchi, A.; Ueda, Y.; Sugita, T.; Shimojo, K. Extraction ability of 4-benzoyl-3-phenyl-5-isoxazolone towards 4f-ions into ionic and molecular media. Anal. Sci. 2018, 34, 973. [Google Scholar] [CrossRef]
  44. Atanassova, M.; Kukeva, R.; Stoyanova, R.; Todorova, N.; Kurteva, V. Synergistic and antagonistic effects during solvent extraction of Gd(III) ion in ionic liquids. J. Mol. Liqs. 2022, 353, 118818. [Google Scholar] [CrossRef]
  45. Atanassova, M.; Todorova, S.; Kurteva, V.; Todorova, N. Insights into the synergistic selectivity of 4f-ions implementing 4-acyl-5-pyrazolone and two new unsymmetrical NH-urea containing ring molecules in an ionic liquid. Sep. Purif. Technol. 2018, 204, 328–335. [Google Scholar] [CrossRef]
  46. Soukeur, A.; Szymczyk, A.; Berbar, Y.; Amara, M. Extraction of rare earth elements from waste products of phosphate industry. Sep. Purif. Technol. 2021, 256, 117857. [Google Scholar] [CrossRef]
  47. Atanassova, M.; Dukov, I. Synergistic solvent extraction and separation of trivalent lanthanide metals with mixtures of 4-benzoyl-3-methyl-1-phenyl-2-pyrazolin-5-one and Aliquat 336. Sep. Purif. Technol. 2004, 40, 171. [Google Scholar] [CrossRef]
  48. Atanasova, M.; Dukov, I. Crown ethers as synergistic agents in the solvent extraction of trivalent lanthanoids with thenoyltrifluoroacetone. Sep. Sci. Technol. 2005, 40, 1104. [Google Scholar] [CrossRef]
  49. Atanassova, M. Synergictic solvent extraction of lanthanide(III) ions from perchlorate medium with 4-benzoyl-3-phenyl-5-isoxazolone in the presence of quaternary ammonium salt. Solvent Extr. Ion Exch. 2009, 27, 159. [Google Scholar] [CrossRef]
  50. Atanassova, M.; Kurteva, V. Synergism as a phenomenon in solvent extraction of 4f-elements with calixarenes. RSC Adv. 2016, 6, 11303. [Google Scholar] [CrossRef]
  51. Atanassova, M.; Dukov, I. Synergistic Solvent extraction of trivalent lanthanoids with mixtures of 1-phenyl-3-methyl-4-benzoyl-5-pyrazolone and crown ethers. Acta Chim. Slovenika 2006, 53, 457. [Google Scholar]
  52. Healy, T.V. Synergism with thenoyltrifluoracetone in the solvent extraction of metallic species. Nucl. Sci. Eng. 1963, 16, 413. [Google Scholar] [CrossRef]
  53. Nakamura, E.; Hiruta, Y.; Watanabe, T.; Iwasawa, N.; Citterio, D.; Suzuki, K. A fluorous biphasic solvent extraction system for lanthanides with a fluorophilic β-diketone type extractant. Anal. Sci. 2015, 31, 923. [Google Scholar] [CrossRef]
  54. Schmit, V. Ekstraktziya Aminami; Atomizdat: Moscow, Russia, 1980. (In Russian) [Google Scholar]
  55. Hong, S.-W.; Chang, Y.; Hwang, M.-J.; Rhee, I.-S.; Kang, D.-S. Determination of Electron Spin Relaxation Time of the Gadolinium-Chealted MRI Contrast Agents by Using an X-band EPR Technique. J. Korean Soc. Magn. Reson. Med. 2000, 4, 27. [Google Scholar]
  56. Yordanov, N. Quantitative EPR spectrometry—“State of the Art”. Appl. Magn. Reson. 1994, 6, 241. [Google Scholar] [CrossRef]
Figure 1. LogDP vs. log[HP] for Gd(III) liquid–liquid extraction at pHin = 3.30 from EtG solutions.
Figure 1. LogDP vs. log[HP] for Gd(III) liquid–liquid extraction at pHin = 3.30 from EtG solutions.
Separations 10 00286 g001
Figure 2. LogDP vs. log[HP] for Gd(III) liquid–liquid extraction from aqueous solutions.
Figure 2. LogDP vs. log[HP] for Gd(III) liquid–liquid extraction from aqueous solutions.
Separations 10 00286 g002
Figure 3. LogDP vs. pH of the more polar phase at [HP] = 4 × 10−2 mol/dm3 (slopes 0.67 and 1.15 for EtG and H2O, respectively).
Figure 3. LogDP vs. pH of the more polar phase at [HP] = 4 × 10−2 mol/dm3 (slopes 0.67 and 1.15 for EtG and H2O, respectively).
Separations 10 00286 g003
Figure 4. LogDP vs. log[HP] at constant pH (C6H12, pH = 2.16 and 0.43; n-hexane, pH = 2.70 and 1.26; n-heptane, pH = 2.19 and 0.51, respectively, H2O and EtG).
Figure 4. LogDP vs. log[HP] at constant pH (C6H12, pH = 2.16 and 0.43; n-hexane, pH = 2.70 and 1.26; n-heptane, pH = 2.19 and 0.51, respectively, H2O and EtG).
Separations 10 00286 g004
Figure 5. LogDP vs. pH of the more polar phase at [HP] = 4.2 × 10−2 mol/dm3 using n-heptane as a diluent.
Figure 5. LogDP vs. pH of the more polar phase at [HP] = 4.2 × 10−2 mol/dm3 using n-heptane as a diluent.
Separations 10 00286 g005
Figure 6. Solvent extraction performance of HP ligand (2 × 10−2 mol/dm3) diluted in two ILs, i.e., [C1C4im+]/[C1C10im+][Tf2N] and CHCl3 for 25 metal ions. Extraction data with the more polar organic phase consisting of EtG or water were also involved, in the first three and the second three positions, respectively, in the cells (see the explanation under NO.11 cell). The reported extractability values (%) represent the average of three measurements with a deviation of less than 5%.
Figure 6. Solvent extraction performance of HP ligand (2 × 10−2 mol/dm3) diluted in two ILs, i.e., [C1C4im+]/[C1C10im+][Tf2N] and CHCl3 for 25 metal ions. Extraction data with the more polar organic phase consisting of EtG or water were also involved, in the first three and the second three positions, respectively, in the cells (see the explanation under NO.11 cell). The reported extractability values (%) represent the average of three measurements with a deviation of less than 5%.
Separations 10 00286 g006
Figure 7. Extraction performance of HP ligand (2 × 10−2 mol/dm3) diluted C1C10im+][Tf2N] for 25 metal ions. Extraction data with the more polar organic phase consisting of glycerol and water (1:1, v/v). The reported extractability values (%) represent the average of three measurements with deviation of less than 5%.
Figure 7. Extraction performance of HP ligand (2 × 10−2 mol/dm3) diluted C1C10im+][Tf2N] for 25 metal ions. Extraction data with the more polar organic phase consisting of glycerol and water (1:1, v/v). The reported extractability values (%) represent the average of three measurements with deviation of less than 5%.
Separations 10 00286 g007
Figure 8. SFs for Lu/Eu pair using various solvent systems including chelating ligands: 4,4,4-trifluoro-1-(2-thienyl)-1,3-butanedione (HTTA), 3-methyl-1-phenyl-4-(4-trifluoromethylbenzoyl)-pyrazol-5-one (HPMTFBP), 3-methyl-1-phenyl-4-(4-phenylbenzoyl)-pyrazol-5-one (HPPMBP), 3-methyl-4-(4-methylbenzoyl)-1-phenyl-pyrazol-5-one (HPMMBP), 4-(4-fluorobenzoyl)-3-methyl-1-phenyl-pyrazol-5-one (HPMFBP), and 4-benzoyl-3-phenyl-5-isoxazolone (HPBI) [47,48,49,50].
Figure 8. SFs for Lu/Eu pair using various solvent systems including chelating ligands: 4,4,4-trifluoro-1-(2-thienyl)-1,3-butanedione (HTTA), 3-methyl-1-phenyl-4-(4-trifluoromethylbenzoyl)-pyrazol-5-one (HPMTFBP), 3-methyl-1-phenyl-4-(4-phenylbenzoyl)-pyrazol-5-one (HPPMBP), 3-methyl-4-(4-methylbenzoyl)-1-phenyl-pyrazol-5-one (HPMMBP), 4-(4-fluorobenzoyl)-3-methyl-1-phenyl-pyrazol-5-one (HPMFBP), and 4-benzoyl-3-phenyl-5-isoxazolone (HPBI) [47,48,49,50].
Separations 10 00286 g008
Figure 9. EPR spectra of frozen solutions of complexes of Gd3+ ([Gd3+]in = 6 × 10−4 mol/dm3) with ligands [HP] = 8 × 10−3 mol/dm3 (1, 2 and 3) and [HTTA] = 8 × 10−3 mol/dm3 (a, b, and c) obtained after solvent extraction from EtG, respectively, in the following diluents [C1C4im+][Tf2N] (1: pHeq = 1.90 and a: pHeq = 3.18), ([C1C10im+][Tf2N] (2: pHeq = 1.80 and b: pHeq = 2.76), CHCl3, (3: pHeq = 2.32 and c: pHeq = 5.81). On the left: range of the magnetic field 0−700 mT, on the right: range of the magnetic field 300−360 mT. Calculated logDL are 0.27, 0.12, and −0.37 for samples 1-2-3 as well as −0.59, −0.06, and −0.92 for samples a-b-c.
Figure 9. EPR spectra of frozen solutions of complexes of Gd3+ ([Gd3+]in = 6 × 10−4 mol/dm3) with ligands [HP] = 8 × 10−3 mol/dm3 (1, 2 and 3) and [HTTA] = 8 × 10−3 mol/dm3 (a, b, and c) obtained after solvent extraction from EtG, respectively, in the following diluents [C1C4im+][Tf2N] (1: pHeq = 1.90 and a: pHeq = 3.18), ([C1C10im+][Tf2N] (2: pHeq = 1.80 and b: pHeq = 2.76), CHCl3, (3: pHeq = 2.32 and c: pHeq = 5.81). On the left: range of the magnetic field 0−700 mT, on the right: range of the magnetic field 300−360 mT. Calculated logDL are 0.27, 0.12, and −0.37 for samples 1-2-3 as well as −0.59, −0.06, and −0.92 for samples a-b-c.
Separations 10 00286 g009
Figure 10. EPR spectra of Cu2+ obtained after solvent extraction with HP = 2 × 10−2 mol/dm3 in [C1C4im+][Tf2N] (1) and [C1C10im+][Tf2N] (2) recorded in frozen solutions (left) and at room temperature (right): MP phase—[Cu2+]in = 6 × 10−4 mol/dm3 and 0.1 mol/dm3 MES in EtG. Calculated %E is 98.66% and 99.53%—ICP-OES based on an average of three parallel determinations.
Figure 10. EPR spectra of Cu2+ obtained after solvent extraction with HP = 2 × 10−2 mol/dm3 in [C1C4im+][Tf2N] (1) and [C1C10im+][Tf2N] (2) recorded in frozen solutions (left) and at room temperature (right): MP phase—[Cu2+]in = 6 × 10−4 mol/dm3 and 0.1 mol/dm3 MES in EtG. Calculated %E is 98.66% and 99.53%—ICP-OES based on an average of three parallel determinations.
Separations 10 00286 g010
Figure 11. EPR spectra of Fe3+ obtained after solvent extraction with HP = 2 × 10−2 mol/dm3 in [C1C4im+][Tf2N] and [C1C10im+][Tf2N] recorded in frozen solutions: MP phase—[Fe3+]in = 6 × 10−4 mol/dm3 and 0.1 mol/dm3 MES in EtG. Calculated %E is 95.82% and 96.12%—ICP-OES based on an average of three parallel determinations.
Figure 11. EPR spectra of Fe3+ obtained after solvent extraction with HP = 2 × 10−2 mol/dm3 in [C1C4im+][Tf2N] and [C1C10im+][Tf2N] recorded in frozen solutions: MP phase—[Fe3+]in = 6 × 10−4 mol/dm3 and 0.1 mol/dm3 MES in EtG. Calculated %E is 95.82% and 96.12%—ICP-OES based on an average of three parallel determinations.
Separations 10 00286 g011
Table 1. Values of EPR parameters of Cu2+ ions established after HP complexation in ILs.
Table 1. Values of EPR parameters of Cu2+ ions established after HP complexation in ILs.
Cu2+−[C1C4im+][Tf2N] 100KCu2+−[C1C10im+][Tf2N] 100KCu2+−[C1C4im+][Tf2N] 295KCu2+−[C1C10im+][Tf2N] 295K
g1 = 2.0575, g2 = 2.0636
g3 = 2.3034
A = 17.62 mT
 
g1 = 2.0526, g2 = 2.0662
g3 = 2.31057
A = 15.53 mT
g1 = 2.0583, g2 = 2.0645
g3 = 2.2980
A = 17.66 mT
 
g1 = 2.0625, g2 = 2.0663
g3 = 2.3155
A = 15.16 mT
g1 = 2.0746, g2 = 2.0596
g3 = 2.2975
A = 14.45 mT
g1 = 2.0692
g2 = 2.0560
g3 = 2.2952
A = 15.00 mT
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Atanassova, M.; Kukeva, R. Improvement of Gd(III) Solvent Extraction by 4-Benzoyl-3-methyl-1-phenyl-2-pyrazolin-5-one: Non-Aqueous Systems. Separations 2023, 10, 286. https://doi.org/10.3390/separations10050286

AMA Style

Atanassova M, Kukeva R. Improvement of Gd(III) Solvent Extraction by 4-Benzoyl-3-methyl-1-phenyl-2-pyrazolin-5-one: Non-Aqueous Systems. Separations. 2023; 10(5):286. https://doi.org/10.3390/separations10050286

Chicago/Turabian Style

Atanassova, Maria, and Rositsa Kukeva. 2023. "Improvement of Gd(III) Solvent Extraction by 4-Benzoyl-3-methyl-1-phenyl-2-pyrazolin-5-one: Non-Aqueous Systems" Separations 10, no. 5: 286. https://doi.org/10.3390/separations10050286

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