Next Article in Journal
Effects of α-Melanocortin Enantiomers on Acetaminophen-Induced Hepatotoxicity in CBA Mice
Next Article in Special Issue
The Solubility of Hen Lysozyme in Ethylammonium Nitrate/H2O Mixtures and a Novel Approach to Protein Crystallization
Previous Article in Journal
Hepatoprotective Potential of Extracts from Seeds of Areca catechu and Nutgalls of Quercus infectoria
Previous Article in Special Issue
Cellulose/Gold Nanocrystal Hybrids via an Ionic Liquid/Aqueous Precipitation Route
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

18-Crown-6 and Dibenzo-18-crown-6 Assisted Extraction of Cesium from Water into Room Temperature Ionic Liquids and Its Correlation with Stability Constants for Cesium Complexes

1
State Research Institute of Reagents and High Purity Substances (IREA), Bogorodskii val 3, 107258, Moscow, Russia
2
Department of Chemistry, M.V. Lomonosov Moscow State University, 1-3 Leninskiye gory, 119992, Moscow, Russia
3
Department of Physical and Colloid Chemistry, Moscow State University of Food Technologies, Volokolamskoye Sh. 11, 125080, Moscow, Russia
4
Zelinsky Institute of Organic Chemistry RAS, Leninsky pr. 47, 119991, Moscow, Russia
*
Author to whom correspondence should be addressed.
Molecules 2009, 14(12), 5001-5016; https://doi.org/10.3390/molecules14125001
Submission received: 1 September 2009 / Revised: 24 November 2009 / Accepted: 27 November 2009 / Published: 2 December 2009
(This article belongs to the Collection Ionic Liquids)

Abstract

:
The pH-profiles of the extraction of Cs+ into four conventional (1-butyl-3- methylimidazolium hexafluorophosphate and bis[trifluoromethyl)sulphonyl]imides of 1- butyl-3-methylimidazolium, 1-hexyl-3-methylimidazolium, and 1-(2-ethylhexyl)-3-methylimidazolium) and two novel (trioctylmethylammonium salicylate and tetrahexyl-ammonium dihexylsulfosuccinate) room temperature ionic liquids have been determined both in the absence and in the presence of crown ether (18-crown-6 or dibenzo-18-crown-6). The pH-profiles of distribution ratio of crown ethers have been established in the same conditions. The relationship of cesium extraction efficiency both with the stability of its complexes with crown ethers and crown ethers’ distribution ratio has been clarified.

Graphical Abstract

1. Introduction

Hydrophobic crown ethers are applied as extractants in the separation and recovery of 137Cs and 90Sr radioactive isotopes from high-level nuclear waste solutions, containing high concentrations of acids and salts, into molecular solvents [1,2,3,4,5]. Room-temperature ionic liquids (RTILs) are attracting increasing attention in solvent extraction processes due to important advantages over conventional organic diluents such as negligible vapor pressure, low flammability, moisture stability, unusual extraction properties and possibility to eliminate aqueous phase acidification [6,7,8,9,10,11,12,13]. It was demonstrated that extraction efficiency of RTIL can be modulated by using complexing agents, for example, crown ethers. Dai et al. [8] discovered that highly efficient extraction of strontium ions can be achieved when dicyclohexane-18-crown-6 is combined with RTILs. Rogers et al. [9] and Bartsch et al. [12] reported the extraction of various alkali metal ions with crown ethers in RTILs. Visser and Rogers demonstrated that octyl(phenyl)-N,N-diisobutylcarbamoylmethylphosphine oxide dissolved in RTILs enhances the extractability of lanthanides and actinides in comparison to conventional organic solvents [13]. The extraction of silver ions was found to be greatly enhanced by a combined application of RTIL and calix[4]arene compared to that of chloroform [14]. In addition, the task-specific RTILs with complexing functionality built in the RTIL cation have been reported [15,16]. Recently the efficiency of chelate extraction of 3d-cations with 8-sulfonamidoquinoline [17], Pu(IV) with carbamoylmethylphosphine oxide [18] and uranyl ion with tributylphosphate (TBP) [19] from aqueous phase into RTILs was reported. The higher selectivity of dibenzo-18-crown-6 to K+ over Na+ in N-octadecylisoquinolinium tetrakis[3,5-bis(trifluoromethyl)phenyl]borate compared with that in molecular solvents suggests that RTIL provides a unique solvation environment for the complexation of crown ethers with the ions [20].
Besides the issues of cation, ligand and complex solubility in water and in RTIL, the relative stabilities of complex formation in both phases are of significant importance for extraction selectivity. In our recent communication [21] we have reported on the thermodynamics of complex formation of cesium ions with 18-crown-6 (18C6, L) in six hydrophobic RTILs (see Scheme 1): trioctylmethylammonium salicylate ([TOMA][Sal]), tetrahexylammonium dihexylsulfosuccinate ([THA][DHSS]), 1-butyl-3-methylimidazolium hexafluorophosphate ([BMIM][PF6]), 1-butyl-3- methylimidazolium bis[trifluoromethyl)sulphonyl]imide ([BMIM][N(Tf)2]), 1-hexyl-3-methyl- imidazolium bis[trifluoromethyl)sulphonyl]imide ([HMIM][N(Tf)2]), and 1-(2-ethylhexyl)-3-methyl- imidazolium bis[trifluoromethyl)sulphonyl]imide ([EtHMIM][N(Tf)2]) by a 133Cs-NMR technique. The present work aims to study the extraction capability of cesium by the above mentioned RTILs in the presence of 18C6 and dibenzo-18-crown-6 (DB18C6, L) and to clarify its relationship with the stability of complexes formed in aqueous and organic phases. Another objective is to study the impact of crown ether’s nature on the extraction efficiency. For this reason the two crown ethers with similar cavity, but different hydrophobicity (18C6 and DB18C6) have been chosen.

2. Results and Discussion

The results of the extraction experiments are presented in Table 1, Table 2 and Table 3, where DCs refers to cesium distribution ratio for the extraction from aqueous phase into RTIL in the absence of crown ethers, DCs18C6 as well as DCsDB18C6 indicate the same in the presence of 18C6 and DB18C6, respectively, while D18C6Cs and DDB18C6Cs are distribution ratios for 18C6 and DB18C6 in the presence of cesium ions.
The RTILs studied demonstrate different extraction behavior towards cesium. [TOMA][Sal] and [THA][DHSS] appeared to be better solvents for this cation then water and provide positive logDCs values within the whole pH range, Table 1. This fact can be explained by the ability of salicylate and dihexylsulfosuccinate anions to interact with metal ion. Though such an interaction is not expected to be strong, worthy is mentioning that the bulk concentration of RTILs complexing anion is rather high: [Sal] = 1.68 and [DHSS] = 1.29 mol·dm-3. In contrast, conventional-anion RTILs [BMIM][PF6], [BMIM][N(Tf)2], [HMIM][N(Tf)2], and [EtHMIM][N(Tf)2] reveal negative logDCs values, which are in a good agreement with those found for [BMIM][PF6] earlier [12], Table 4. The cesium extraction ability of all RTILs, with the exception of [EtHMIM][N(Tf)2], slightly diminishes as pH decreases.
The situation becomes completely different when crown-ether molecules are introduced into RTIL/water biphasic systems. 18C6 significantly decreases cesium’s uptake by [THA][DHSS] but increases it for all other RTILs, Table 2. As far as we know, [THA][DHSS] is the first example of a negative influence of crown ether on a cation extraction into organic phase. Probably, the main reason is that crown-ether competes with a RTIL’s anion for cesium and, being better soluble in water, provides a partial transfer of the metal ion into an aqueous phase. This suggestion is indirectly supported by the low value of [Cs(18C6)]+ stability constant in [THA][DHSS] (logK1), Table 4, which appears to be even lower than in water.
For DB18C6 an enhancement of cesium extraction is much more pronounced then for 18C6, Table 3. Notably, increasing HNO3 concentration in an aqueous phase decreases cesium’s recovery for all RTILs, but at very low pH both logDCs18C6 and logDCsDB18C6 start to increase. Thus, there appears a minimum at the curves logDCs = f(pH), in between pH 0 and pH 7 (Figure 1).
A similar effect has been observed in [9] for cesium distribution in [BMIM][PF6] in the presence of 4,4’-(5’)-di-(tert-butylcyclohexano)-18-crown-6. Notably, cesium recovery in strongly acidic media is usually lower than in neutral solutions (with the exception of systems 18C6/[THA][DHSS], 18C6/[TOMA][Sal], and DB18C6/[TOMA][Sal]). Thus, unlike crown ether solutions in conventional molecular solvents, such solutions in the studied RTILs do not need acidification to increase a recovery. Moreover, in most cases the addition of acid leads to the decrease of cesium recovery.
Worthy of mention is the fact that the extraction efficiency of solutions of CE in conventional solvents is often unsatisfactory even in strongly acidic media. According to Yakshin et al. [22], in the case of Cs+ (1 × 10-3 M) extraction from 2 M HNO3 aqueous solution into 0.1 M 18C6 in CHCl3 the value of DCs is extremely low (3 × 10-3; logDCs = -2.52). This value is lower than even the minimal values of DCs obtained in the case of our RTILs-based systems. As a rule, the extraction efficiency of the systems based on conventional solvents may be raised to the level of CE/RTIL-systems only by addition of a bulk hydrophobic counter-ion (such as picrate) to the aqueous phase. So, the logDCs18C6 values for 1,2-dichloroethane in the presence of picrate in aqueous phase are of the same level that for some RTILs (e.g., [HMIM][N(Tf)2] and [TOMA][Sal], Table 4), while the hazardous properties of the former are incomparably higher.
The crown ether influence on cesium extraction by RTILs is evidently explained by a formation of rather stable complexes. Previously, we found [21] that Cs+ forms stable complexes [Cs(18C6)]+ in [TOMA][Sal], [THA][DHSS], [BMIM][PF6], [BMIM][N(Tf)2], [HMIM][N(Tf)2] and [EtHMIM][N(Tf)2], and the stability constants have been measured. Besides, in [BMIM][N(Tf)2], [HMIM][N(Tf)2], and [EtHMIM][N(Tf)2] the species [Cs(18C6)2]+ are also formed [21]. In an aqueous phase only [Cs(18C6)]+ species are registered [24].
Generally, when an aqueous phase of cesium nitrate is in equilibrium with a RTIL organic phase, containing crown ether L, the cesium distribution ratio DCsL (for such RTILs as [TOMA][Sal], [THA][DHSS] and [BMIM][PF6]), where only CsL species are formed is represented by a simple equation (1):
DCsL = ([Cs]RTIL + [CsL]RTIL)/([Cs]w + [CsL]w)
where the superscripts “RTIL” and “w” denote organic and aqueous phase, respectively. When the total concentration [L]t >> [Cs]t and the stabilities of complexes in both phases are high enough (logK1RTIL ≥ 2; logK1w ≥ 2), then [Cs]RTIL << [CsL]RTIL and [Cs]w << [CsL]w. In this case, the equilibrium concentrations of [Cs]RTIL and [Cs]w in (1) can be neglected. The equation (1) is therefore transformed into (2):
DCsL = [CsL]RTIL/[CsL]w
or:
DCsL = (K1RTIL[Cs]RTIL[L]RTIL)/(K1w[Cs]w[L]w).
As far as [Cs]RTIL/[Cs]w = DCs and [L]RTIL/[L]w = DL:
DCsL = K1RTIL DCsDL/K1w
or:
logDCsL = logK1RTIL + logDCs + logDL − logK1w.
Thus, the plot of (logDCsL − logDCs − logDL) versus logK1RTIL was expected to be linear with a slope of 1. In our case the initially stated requirement logK1RTIL ≥ 2; logK1w ≥ 2 does not take place for some RTILs and for water. Indeed, logK1w = 0.98 [24] and for the RTILs the following values of logK1RTIL at 25 °C have been determined [21]: 1.45 (0.05) ([TOMA][Sal]); 0.77 (0.04) ([THA][DHSS]); 2.4 (0.2) ([BMIM][PF6]); 3.4 (0.5) ([BMIM][N(Tf)2]); 4.4 (0.1) ([HMIM][N(Tf)2]), and 3.4 (0.4) ([EtHMIM][N(Tf)2]). Nevertheless, the plot of (logDCs18C6 − logDCs − logD18C6Cs) versus logK1RTIL for [TOMA][Sal], [THA][DHSS], and [BMIM][PF6] represents a perfectly strait line with a slope 0.73 and R2 = 0.999, Figure 2, Table 4. In order to confirm the legitimacy of the derivation of the equation 4 we have determined the concentration profiles of DCs for all RTILs in the range Cw0(Cs) = 2 × 10−4 − 5 × 10−2 M. In all cases DCs remains nearly constant at 2·10−4 ≤ Cw0(Cs) ≤ 1.5·10−3 M (≤ 5·10−3 M − for [TOMA][Sal] and [THA][DHSS]). At the same time the maximal cesium recovery (in the case of 18C6) is observed for 18C6/[BMIM][N(Tf)2] at pH = 6.88 (Table 2), which corresponds to the residual free cesium concentration being equal to 2.9 × 10−4 M (i.e., even in this case the ratio [free Cs in the organic phase]/[free Cs in the aqueous phase] is the same that in the case of Cs+-extraction from 1.5 × 10−3 M aqueous solution in the absence of crown ether in RTIL-phase).
In the case of [BMIM][N(Tf)2], [HMIM][N(Tf)2], [EtHMIM][N(Tf)2] the equations (1), (2), (4) and (5) become more complicated due to formation of CsL2 complexes and the lack of such complexes in water. However, all six RTILs fit equation (5) with a slope 0.89 and a correlation coefficient R2 = 0.98, Figure 3. This can be explained by a domination of logK1RTIL values over logK2RTIL ones, Table 4. Another reason of low contribution of K2RTIL to the extraction is in the absence of enough sufficient excess of crown re. cesium concentration. Anyway, the linear relationship of (logDCs18C6 − logDCs − logD18C6Cs) and logK1RTIL indicates clearly an importance of complex stability contribution for the extraction of cesium into hydrophobic RTILs.
Besides logK1RTIL, also a contribution of logDL should be examined. Thus, along with cesium extraction we have studied the water/RTIL distribution of crown ethers, Table 2 and Table 3. 18C6 demonstrates similar affinity towards as the RTILs, so the conventional molecular solvents: 1,2-dichloroethane, chloroform and dichloromethane [5]. The detaining of DB18C6 in extracts is sufficiently higher for all RTILs then that of 18C6, Table 2 and Table 3. At the same time logDDB18C6Cs values are one to two log unites lower then those found for benzene, 1,2-dichloroethane, chloroform and nitrobenzene [5].
It should be noted that the variation of logDCs18C6 and logDCsDB18C6 with pH shows the same trend as the crown ether distribution between aqueous phase and RTIL, logD18C6Cs and logDDB18C6Cs, Figure 1. A linear relationship between logD18C6Cs and logDCs18C6 (slope 0.91; R2 = 0.96) as well as between logDDB18C6Cs and logDCsDB18C6 (slope 1.24; R2 = 0.91) was found for TOMAS, Figure 4, indicating the dominating role of relative crown ether solubility in water and RTIL.
Analysis of the data presented in Table 4 for [BMIM][N(Tf)2], [HMIM][N(Tf)2], and [EtHMIM][N(Tf)2] indicates that the contribution of logD18C6Cs into logDCs18C6 prevails over logKRTIL values. Indeed, the variation of logDCs18C6 within this set of similar RTILs with the same anion definitely follows the changes in logD18C6Cs values but not the stability constants variations.
A comparison of 18C6 and DB18C6 indicates a much higher extraction efficiency of the latter. The stability constants of cesium complexes with DB18C6 in the studied RTILs have not been measured yet, but for other than RTIL solvents they reveal 0.2 to 0.5 log units lower values than those for 18C6 [25]. However, bearing in mind that this is also valid for an aqueous solution, the difference in (logK1RTIL − logK1w) in Equation (5) becomes negligible. Moreover, this unfavorable for extraction loss is excessively compensated by the fact that the values of logDDB18C6Cs are 2 to 3 log units higher than logD18C6Cs values.

3. Experimental

3.1. General

[HMIM][N(Tf)2] (Solvent Innovation; puriss. 99%), Bromothymol Blue (Baum-Lux, Russia; analytical reagent grade), KNO3 (analytical reagent grade), chloroform (Ekros, Russia), and absolute ethanol (analytical reagent grade) were used as received. Cesium nitrate (Aldrich, 99.99%) was dried for 6 hours at 70 °C, and then kept 24 h in a desiccator before use. Dibenzo-18-crown-6 (DB18C6, 99% purity) was synthesized by Dr. V.M. Polosin and analyzed by 1H-NMR. NMR revealed no other peaks than those corresponding to DB18C6. 18C6 (Aldrich, 99%) and DB18C6 have been melted at 70 and 180 °C, respectively, in a drying box for 3 h and then kept at 25 °C in a desiccator for 24 h to remove water before use. NMR spectra were recorded on a 500 MHz Bruker DRX500 instrument is the indicated solvents using TMS as reference.

3.2. Synthesis and characterization of the RTILs

Trioctylmethylammonium salicylate. Aliquat® 336 (~0.2 mol) was mixed with 30% excess of sodium salicylate in chloroform (200 mL). The mixture was shaken for 4 h and then rinsed 20 times with a large amount of distilled water, then the solvent was evaporated and the liquid residue was heated up to 100 °C under reduced pressure for 5 h. After cooling to room temperature a white solid matter was obtained with a density 0.943 g·cm−3; Tm = (32.8 ± 0.4) °C, Tf = (14 ± 2) °C. Yield: 90%. 1H-NMR (CDCl3) 0.88 (9 H), 1.24 (30 H), 1.59 (6 H), 3.19 (3 H), 3.27 (6 H), 7.20 (1 H), 7.92 ppm (1 H); 13C-NMR (DMSO-d6) 13.80 (C*CH3), 21.93 (C*H2CH3), 21.25, 25.68, 28.27, 28.32, 31.04 (various CH2CH2 fragments); 47.41 (CH3N); 60.50 (CH2N); 115.35, 115.62, 120.73, 129.73, 130,88 (aromatic C); 163.17 (COH); 171.00 ppm (COO). Analysis: found: C, 76.06; H, 11.64; N: 2.62; calc.: C, 76.49; H, 11.89; N: 2.62 (taking into consideration 2:1 mol/mol relation between methyltrioctyl- and methyltridecylammonium cations in Aliquat® 336). After equilibration with water at ambient temperature the solid product transforms into a slightly yellowish viscous liquid with 0.942 g·cm−3 density and a freezing point below −18 °C. Water content measured by Karl Fisher titration constituted 0.18% wt. (0.09 mol·dm−3) for a solid product and 4.83% wt. (2.52 mol·dm−3) for that equilibrated with water RTIL.
Tetrahexylammonium dihexylsulfosuccinate was synthesized according to [26] as a transparent viscous liquid (yield: 85%), analyzed by NMR, and then used without further purification. 1H-NMR (CDCl3) 0.88 (18 H), 1.33 (36 H), 1.55 (12 H), 3.2 (12 H), 4.1 (3 H) ppm; 13C-NMR (DMSO-d6) 13.64, 13.69, 13.71 (CH3); 20.91, 21.76 (C*H2CH3), 21.87, 24.82, 24.85, 25.34, 27,95, 27.99, 30.47, 30.75, 30.82 (CH2CH2 fragments); 34.01 (OOC*CH2CH-); 57.63 (CH2N); 63.89 (CH2O); 168.31, 170.97 COOR); Analysis: found: C, 66.67; H, 11.31; N: 1.97, S: 4.49; calc.: C, 66.71; H, 11.34; N: 1.94, S: 4.45. An absence of halogen-ions was proven by a AgNO3 test.
1-Butyl-3-methylimidazolium hexafluorophosphate. A 1-L, one-necked, round-bottomed flask was equipped with a magnetic stirrer and charged with 1-butyl-3-methylimidazolium chloride (367.2 g, 2.10 mol, 1 equiv.) and potassium hexafluorophosphate (387.3 g, 2.10 mol, 1 equiv.) in distilled water (700 mL). The reaction mixture was stirred at room temperature for 2 h, affording a two-phase system. The organic phase was separated and washed with water (5 × 40 mL) until the aqueous fraction observed to be free of chloride (AgNO3). Then dichloromethane (400 mL) was added. The dichloromethane solution was mixed with activated charcoal, stirred for 2 h, filtered and dried over anhydrous magnesium sulfate. After 1 h, the suspension was filtered and the volatile material was removed by rotary evaporation. The resulting colourless or light-yellow viscous liquid was dried under reduced pressure (0.5 mm Hg) at 70 °C for 12 h. Yield 462 g (77.3%). 1H-NMR (300 MHz, DMSO-d6): 0.91 (3H, t, NHCH2CH2CH2CH3), 1.27 (2H, m, NCH2CH2CH2CH3), 1.77 (2H, p, NCH2CH2CH2CH3), 3.85 (3H, s, NCH3), 4.16 (2H, t, NCH2CH2CH2CH3), 7.72 (2H, m, C(5)H, C(4)H), 9.09 (1H, s, C(2)H); Analysis: found: C, 33.97; H, 5.37; N, 9.95; F, 39.91; P, 10.78; calc.: C, 33.81; H, 5.32; N, 9.86; F, 40.11; P, 10.90.
1-Butyl-3-methylimidazolium bis[trifluoromethyl)sulphonyl]imide. A 500-mL, one-necked, round- bottomed flask was equipped with a magnetic stirrer and charged with lithium bis(trifluoromethylsulfonyl)imide (143.693 g, 0.5 mol, 1 equiv.) and 1-butyl-3-methylimidazolium chloride (87.25 g, 0.5 mol, 1 equiv.) dissolved in distilled water (80 mL). The reaction mixture was stirred at room temperature for 2 h affording a two-phase system. Then dichloromethane (200 mL) was added. The lower organic layer was separated and washed with distilled water (5 × 20 mL) until the aqueous fraction observed to be free of chloride (AgNO3). The dichloromethane solution was mixed with activated charcoal, stirred for 2 h, filtered and dried over anhydrous magnesium sulfate. After 1 h, the suspension was filtered and the volatile material was removed by rotary evaporation. The resulting colourless viscous liquid was dried under reduced pressure (0.5 mm Hg) at 70 °C for 12 h. Yield 162.6 g (77.5%). 1H-NMR (250 MHz, DMSO-d6): 0.89 (3H, t, NHCH2CH2CH2CH3), 1.25 (2H, m, NCH2CH2CH2CH3), 1.76 (2H, p, NCH2CH2CH2CH3), 3.84 (3H, s, NCH3), 4.15 (2H, t, NCH2CH2CH2CH3), 7.72 (2H, m, C(5)H, C(4)H), 9.09 (1H, s, C(2)H); Analysis: found: C, 28.67; H, 3.63; N, 9.94; F, 27.30; S, 15.23; O, 15.23; calc.: C, 28.64; H, 3.61; N, 10.02; F, 27.18; S, 15.29; O, 15.26.
1-(2-Ethylhexyl)-3-methylimidazolium bis[trifluoromethyl)sulphonyl]imide was kindly supplied by Dr. V.E. Baulin (A.N. Frumkin Institute of Physical Chemistry and Electrochemistry of RAS). 1H-NMR (200 MHz, CDCl3) 0.85 (m, 6H, 2CH2CH3), 1.25 (m, 8H, 4CH2), 1.75 (m, 1H, CH2CHCH2), 3.92 (s, 3H, CH3N), 4.00 (d, J=7.0 Hz, 2H, CH2N), 7.25(s, 1Harom), 7.36 (s, 1Harom), 8.65 (s, 1Harom); Analysis: found: C, 35.40; H, 4.85; F, 24.00; N: 8.81; S: 13.52; calc.: C, 35.36; H, 4.88; F, 23.97; N: 8.84, S: 13.49.

3.3. Extraction

For a sample preparation the weighted amount of solid CsNO3 was dissolved in a distilled water to obtain the initial solution of salt (1.5·10-2 mol·dm-3), which was further used for the preparation of the diluted samples. For adjusting pH values of aqueous phase, 0.01–3 M HNO3 (analytical reagent grade) was used.
Cesium extraction process in the absence of crown ethers was performed by adding of water- saturated RTIL samples (0.5 mL for [TOMA][Sal], [THA][DHSS], and [BMIM][PF6]; 1.0 mL for [BMIM][N(Tf)2], [EtHMIM][N(Tf)2] and 2.0 mL for [HMIM][N(Tf)2]) to 1.5 × 10-3 mol·dm-3 (2 × 10-4, 5 × 10-4, 5 × 10-3, 1 × 10-2, and 5 × 10-2 mol·dm-3 only at neutral pH, without addition of HNO3) CsNO3 aqueous solutions (5.0 mL for [TOMA][Sal], [THA][DHSS], and [BMIM][PF6]; 3.0 mL for [BMIM][N(Tf)2], [EtHMIM][N(Tf)2] and 2.0 mL for [HMIM][N(Tf)2]). The biphasic systems were shaken for 2 h (adequate time for equilibrium state establishment for all RTILs studied at room temperature, 22 °C). After centrifugation for 15 min aqueous and organic phases were separated. An aqueous phase was analyzed for equilibrium pH (pH-meter pH-410; combined glass microelectrode ESLK-13.7, Aquilon, Russia) and cesium content (flame photometry, FPA-2; Zagorsk Optical and Mechanical Plant, Russia). The recovery (R(Cs), %) and the distribution constant (D(Cs)) of cesium were found according to eqns. (6) and (7):
R(Cs), % = [1 – Cw(Cs)/Cw0(Cs)]·100%,
D(Cs) = Co(Cs)/Cw(Cs) = [Vw/Vo][R(Cs), %]/[100 – R(Cs), %],
where Cw0(Cs) and Cw(Cs) are the initial and equilibrium concentrations of Cs+ in an aqueous phase, respectively, Vw и Vo correspond to the total volumes of aqueous and organic phase, respectively (mL). The ratio Vw:Vo was 10 ([TOMA][Sal], [THA][DHSS], and [BMIM][N(Tf)2]), 3 ([BMIM][PF6], [EtHMIM][N(Tf)2]) and 1 ([HMIM][N(Tf)2]). To compensate the matrix effects the calibration standards were prepared by adding standard solutions of Cs+ to the aqueous solutions pre-equilibrated with the corresponding RTILs. Otherwise the error increased sufficiently.
For cesium extraction study in the presence of crown ethers, the precise amounts of 18C6 or DB18C6 were dissolved in the RTILs saturated by a distilled water. Initial concentration of 18C6, C(18C6)0 was 1.5 × 10-1 mol·dm-3 for all RTILs studied. The same concentration was used for DB18C6 in [BMIM][N(Tf)2]. Due to the solubility problems, the content of DB18C6 in [TOMA][Sal] and [THA][DHSS] was lower (C(DB18C6)0 = 5 × 10-2 mol·dm-3). This concentration was obtained after heating of these RTILs with the amounts of DB18C6 at 80–90 °C. Thus, these solutions are likely to be oversaturated, although no precipitation after cooling to room temperature was observed.
To provide an equal ratio C(CE)0:C(Cs+)0 = 100 and ν(CE)0:ν(Cs+)0 = 10 for all RTILs studied, the initial concentration of Cs+ in aqueous phases contacting with [TOMA][Sal] and [THA][DHSS] was adjusted to C(Cs+)0 = 5 × 10-4 mol·dm-3. The volume ratio Vw:Vo = 10 was used for all 9 extraction systems studied (0.5 mL crown ether solution in RTIL and 5.0 mL of cesium nitrate aqueous solution with various pH). The aqueous solutions pre-equilibrated with the corresponding CE/RTIL systems were used for preparation of the calibration standards. All other experimental conditions were the same as for experiments run without crown ether.
Distribution of DB18C6 was studied in a parallel way within the same experiments on cesium extraction. The samples taken from phases of DB18C6/[TOMA][Sal], DB18C6/[THA][DHSS], and DB18C6/[BMIM][N(Tf)2] were dissolved in an absolute ethanol taken in vol/vol ratio 1000 (for DB18C6/[BMIM][N(Tf)2]) and 300 (for DB18C6/[TOMA][Sal] and DB18C6/[THA][DHSS]). Then the resulting solutions were analyzed spectrophotometrically (U-2900 UV-Vis Recording Spectrophotometer, Hitachi, Japan; 1 cm quartz cells) for DB18C6 content using crown ether absorption band λmax = 274 nm (εmax = 5,000 ÷ 5,500) and the reference solutions of the corresponding RTILs in ethanol with the same dilution as for DB18C6/RTILs. Detaining (R(CE), %) for DB18C6 has been calculated using equation (8):
R(Cs), % = [Co(Cs)/Co0(Cs)]·100%,
where Co0(CE) and Co(CE) denote initial and equilibrium concentrations of crown ether in organic phase (extract), respectively; while distribution constant (D(CE)) − using equation (7).
Distribution of 18C6 was controlled by the method described elsewhere [27]. Bromothymol Blue (H-BTB) and KNO3 were dissolved in an aliquot taken from an aqueous phase (to adjust 1.2 × 10-2 mol·dm-3 concentration of both of them). Then the extraction of the complex [K(18C6)+·BTB–] into CHCl3 was performed with 2 h contact time and phases volume ratio 1:1; pH of an aqueous phase was kept constant (pH = 6.9) by phosphate buffer for all solutions other then neutral. Then the system was centrifuged for 15 min, an organic phase was separated and analyzed spectrophotometrically (λmax = 400 ÷ 450 nm, εmax = 1,500 ÷ 2,000; the chloroform extract obtained in the blank experiment was used as the reference solution). Detaining (R(CE), %) and distribution ratio (D(CE)) were calculated by the eqns., completely analogous to (1) and (2), taking into consideration that Co(CE) = Co0(CE) − 10·Cw(CE), where Cw(CE) denotes an equilibrium concentration of crown ether in aqueous phase.

4. Conclusions

The data presented reveals that several ionic liquids are suitable for extraction of cesium in the presence of crown ethers. The liquids which show a higher retention of a crown ether and a higher stability of the complexes demonstrate better capabilities to extract Cs+. The former factor dominates evidently over the latter one. Among the crowns the ligand of a choice is the one that demonstrates higher hydrophobicity, and, therefore, a higher detaining in extract, e.g., DB18C6.

Acknowledgements

The authors are grateful to V.E. Baulin who kindly provided [EtHMIM][N(Tf)2] ionic liquid. Our thanks are also due to Russian Foundation for Basic Research (projects no. 07-08-00246 and no. 08-03-91319).

References and Notes

  1. Takeda, Y.; Kawarabayashi, A.; Endō, K.; Yahata, T.; Kudo, Y.; Katsuta, S. Solvent extraction of alkali metal (Li–Cs) picrates with 18-crown-6 into various diluents. Elucidation of fundamental equilibria which govern the extraction-ability and -selectivity. Anal. Sci. 1998, 14, 215–223. [Google Scholar] [CrossRef]
  2. Levitskaia, T.G.; Maya, L.; van Berkel, G.J.; Moyer, B.A. Anion Partitioning and ion-pairing behavior of anions in the extraction of cesium salts by 4,5‘‘-bis(tert-octylbenzo)dibenzo-24- crown-8 in 1,2-dichloroethane. Inorg. Chem. 2007, 46, 261–272. [Google Scholar] [CrossRef] [PubMed]
  3. Abramov, A.A. Extraction of cations by crown-ethers. Moscow Univ. Chem. Bull. 2000, 41, 3–15. [Google Scholar]
  4. Robak, W.; Apostoluk, W.; Maciejewski, P. Analysis of liquid–liquid distribution constants of nonionizable crown ethers and their derivatives. Anal. Chim. Acta 2006, 569, 119–131. [Google Scholar] [CrossRef]
  5. Mohapatra, P.K.; Ansari, S.A.; Sarkar, A.; Bhattacharyya, A.; Manchanda, V.K. Evaluation of calix-crown ionophores for selective separation of radio-cesium from acidic nuclear waste solution. Anal. Chim. Acta 2006, 571, 308–314. [Google Scholar] [CrossRef] [PubMed]
  6. Earle, M.J.; Seddon, K.R. Ionic liquids. Green solvents for future. Pure Appl. Chem. 2000, 72, 1391–1398. [Google Scholar] [CrossRef]
  7. Cocalia, V.A.; Holbrey, J.D.; Gutowski, K.E.; Bridges, N.J.; Rogers, R.D. Separations of metal ions using ionic liquids: The challenges of multiple mechanisms. Tsinghua Sci. Technol. 2006, 11, 188–193. [Google Scholar] [CrossRef]
  8. Dai, S.; Ju, Y.H.; Barnes, C.E. Solvent extraction of strontium nitrate by a crown ether using room-temperature ionic liquids. J. Chem. Soc., Dalton Trans. 1999, 1201–1202. [Google Scholar] [CrossRef]
  9. Visser, A.E.; Swatloski, R.P.; Reichert, W.M.; Griffin, S.T.; Rogers, R.D. Traditional extractants in nontraditional solvents: Groups 1 and 2 extraction by crown ethers in room-temperature ionic liquids. Ind. Eng. Chem. Res. 2000, 39, 3596–3604. [Google Scholar] [CrossRef]
  10. Luo, H.; Dai, S.; Bonnesen, P.V.; Buchanan, A.C., III. Separation of fission products based on ionic liquids: Task-specific ionic liquids containing an aza-crown ether fragment. J. Alloys Compounds 2006, 418, 195–199. [Google Scholar] [CrossRef]
  11. Chen, P.Y. The assessment of removing strontium and cesium cations from aqueous solutions based on the combined methods of ionic liquid extraction and electrodeposition. Electrochim. Acta 2007, 52, 5484–5492. [Google Scholar] [CrossRef]
  12. Chun, S.; Dzyuba, S.V.; Bartsch, R.A. Influence of structural variation in room-temperature ionic liquids on the selectivity and efficiency of competitive alkali metal salt extraction by a crown ether. Anal. Chem. 2001, 73, 3737–3741. [Google Scholar] [CrossRef] [PubMed]
  13. Visser, A.E.; Rogers, R.D. Room-temperature ionic liquids: New solvents for f-element separations and associated solution chemistry. J. Solid State Chem. 2003, 171, 109–113. [Google Scholar] [CrossRef]
  14. Shimojo, K.; Goto, M. Solvent Extraction and stripping of silver ions in room-temperature ionic liquids containing calixarenes. Anal. Chem. 2004, 76, 5039–5044. [Google Scholar] [CrossRef] [PubMed]
  15. Visser, A.E.; Swatloski, R.P.; Reichert, W.M.; Mayton, R.; Sheff, S.; Wierzbicki, A.; Davis, J.H., Jr.; Rogers, R.D. Task-Specific ionic liquids incorporating novel cations for the coordination and extraction of Hg2+ and Cd2+: synthesis, characterization, and extraction studies. Environ. Sci. Technol. 2002, 36, 2523–2529. [Google Scholar] [CrossRef] [PubMed]
  16. Visser, A.E.; Swatloski, R.P.; Reichert, W.M.; Mayton, R.; Sheff, S.; Wierzbicki, A.; Davis, J.H., Jr.; Rogers, R.D. Task-specific ionic liquids for the extraction of metal ions from aqueous solutions. Chem. Commun. 2001, 135–136. [Google Scholar] [CrossRef]
  17. Ajioka, T.; Oshima, S.; Hirayama, N. Use of 8-sulfonamidoquinoline derivatives as chelate extraction reagents in ionic liquid extraction system. Talanta 2008, 74, 903–908. [Google Scholar] [CrossRef] [PubMed]
  18. Lohithakshan, K.V.; Aggarwal, S.K. Solvent extraction studies of Pu(IV) with CMPO in 1-octyl- 3-methylimidazolium hexafluorophosphate (C8mimPF6) room temperature ionic liquid (RTIL). Radiochim. Acta 2008, 96, 93–97. [Google Scholar] [CrossRef]
  19. Dietz, M.L.; Stepinski, D.C. Anion concentration-dependent partitioning mechanism in the extraction of uranium into room-temperature ionic liquids. Talanta 2008, 75, 598–603. [Google Scholar] [CrossRef] [PubMed]
  20. Nishi, N.; Murakami, H.; Imakura, S.; Kakiuchi, T. Facilitated transfer of alkali-metal cations by dibenzo-18-crown-6 across the electrochemically polarized interface between an aqueous solution and a hydrophobic room-temperature ionic liquid. Anal. Chem. 2006, 78, 5805–5812. [Google Scholar] [CrossRef] [PubMed]
  21. Vendilo, A.G.; Djigailo, D.I.; Rönkkömäki, H.; Lajunen, M.; Chernikova, E.A.; Lajunen, L.H.J.; Pletnev, I.V.; Popov, K.I. Thermodynamics of cesium complexes formation with 18-crown-6 in hydrophobic ionic liquids. A correlation with extraction capability. In Macrocyclic Chemistry: New Research Developments, 1st ed.; Nova Science Publishers Inc.: Hauppauge, NY, USA, 2010; in press. [Google Scholar]
  22. Yakshin, V.V.; Vilkova, O.N.; Tsarenko, N.A.; Demin, S.V.; Zhilov, V.I.; Tsivadze, A.Y. Molecular Design of macrocyclic extractants for extraction and separation of alkali and alkaline- earth metals. Russ. J. Coord. Chem. 2006, 32, 83–87. [Google Scholar] [CrossRef]
  23. Kikuchi, Y.; Sakamoto, Y. Complex formation of alkali metal ions with 18-crown-6 and its derivatives in 1,2-dichloroethane. Anal. Chim. Acta 2000, 403, 325–332. [Google Scholar] [CrossRef]
  24. Arnaud-Neu, F.; Delgado, R.; Chaves, S. Critical evaluation of stability constants and thermodynamic functions of metal complexes of crown ethers. Pure Appl. Chem. 2003, 75, 71–102. [Google Scholar] [CrossRef]
  25. IUPAC Stability Constants Database, version 2007; Academic Software: Timble, UK.
  26. Egorov, V.M.; Smirnova, S.V.; Pletnev, I.V. Highly efficient extraction of phenols and aromatic amines into novel ionic liquids incorporating quaternary ammonium cation. Sep. Purif. Technol. 2008, 63, 710–715. [Google Scholar] [CrossRef]
  27. Vilkova, O.M.; Yakshin, V.V. Extraction-photometric determination of the composition of multicomponent crown ether—Podand mixtures. J. Anal. Chem. 2003, 58, 27–30. [Google Scholar] [CrossRef]
Sample Availability: 18C6, DB18C6, [BMIM][PF6], [BMIM][N(Tf)2], and [HMIM][N(Tf)2] are commercially available.
Scheme 1. Structures of the RTILs studied.
Scheme 1. Structures of the RTILs studied.
Molecules 14 05001 sch001
Figure 1. Distribution of cesium, 18C6, and DB18C6 vs. pH in [TOMA][Sal] based systems.
Figure 1. Distribution of cesium, 18C6, and DB18C6 vs. pH in [TOMA][Sal] based systems.
Molecules 14 05001 g001
Figure 2. A plot of [logDCs18C6 − logDCs − logD18C6Cs] vs. logK1RTIL for [TOMA][Sal], [THA][DHSS] and [BMIM][PF6] at 22–25 °C.
Figure 2. A plot of [logDCs18C6 − logDCs − logD18C6Cs] vs. logK1RTIL for [TOMA][Sal], [THA][DHSS] and [BMIM][PF6] at 22–25 °C.
Molecules 14 05001 g002
Figure 3. A plot of [logDCs18C6 − logDCs − logD18C6Cs] vs. logK1RTIL for [TOMA][Sal], [THA][DHSS], [BMIM][PF6], [BMIM][N(Tf)2], [HMIM][N(Tf)2], [EtHMIM][N(Tf)2] at 22–25 °C.
Figure 3. A plot of [logDCs18C6 − logDCs − logD18C6Cs] vs. logK1RTIL for [TOMA][Sal], [THA][DHSS], [BMIM][PF6], [BMIM][N(Tf)2], [HMIM][N(Tf)2], [EtHMIM][N(Tf)2] at 22–25 °C.
Molecules 14 05001 g003
Figure 4. Plots of [logDCsCE − logDCs] vs. logDCECs for TOMAS at 22 °C. (a) CE = 18C6; (b) CE = DB18C6.
Figure 4. Plots of [logDCsCE − logDCs] vs. logDCECs for TOMAS at 22 °C. (a) CE = 18C6; (b) CE = DB18C6.
Molecules 14 05001 g004aMolecules 14 05001 g004b
Table 1. Distribution ratios of cesium (Cw0(Cs) = 1.5 × 10-3 mol·dm-3) between RTIL/aqueous phases in the absence of crown ether as a function of pH at 22 °C.
Table 1. Distribution ratios of cesium (Cw0(Cs) = 1.5 × 10-3 mol·dm-3) between RTIL/aqueous phases in the absence of crown ether as a function of pH at 22 °C.
RTILVo, mLVw, mLAqueous pHlogDCs
[THA][DHSS]0.55.05.941.21
3.361.11
1.590.84
[TOMA][Sal]0.55.05.980.69
4.950.76
3.010.74
1.130.10
[BMIM][PF6]0.55.05.90-0.59
3.33-0.82
1.25-0.80
[BMIM][N(Tf)2]1.03.06.44-0.67
3.35-1.24
1.25-1.31
[EtHMIM][N(Tf)2]1.03.06.34-0.81
2.09-0.62
[HMIM][N(Tf)2]2.02.05.82-1.24
2.84-1.36
0.93-2.21
Table 2. Distribution ratios of cesium [Cw0(Cs) = 1.5 × 10-3 mol·dm-3] and 18C6 [Co0(18C6) = 1.5 × 10-1 mol·dm-3] in the presence of each other between RTIL/aqueous phases as a function of pH at 22 °C (Vo = 0.5 mL; Vw = 5.0 mL).
Table 2. Distribution ratios of cesium [Cw0(Cs) = 1.5 × 10-3 mol·dm-3] and 18C6 [Co0(18C6) = 1.5 × 10-1 mol·dm-3] in the presence of each other between RTIL/aqueous phases as a function of pH at 22 °C (Vo = 0.5 mL; Vw = 5.0 mL).
RTILAqueous pHlogDCs18C6logD18C6Cs
[THA][DHSS]5.920.24-0.12
4.300.23-0.16
3.860.03-0.41
3.39-0.38-0.50
2.480.57-0.15
1.600.75-0.01
[TOMA][Sal]5.840.830.48
4.330.890.44
3.810.900.36
3.310.770.30
2.220.360.21
1.201.110.43
[BMIM][PF6]5.84-0.200.13
4.33-0.450.08
3.81-0.620.06
3.31-1.270.00
2.22-0.740.00
1.20-0.660.08
[BMIM][N(Tf)2]6.881.560.77
5.601.430.65
3.56-0.260.50
2.26-0.270.08
1.27-0.230.23
[EtHMIM][N(Tf)2]6.500.56-0.27
2.150.30-0.45
[HMIM][N(Tf)2]5.600.820.25
4.100.910.17
2.960.610.06
1.40-0.220.08
0.90-0.330.27
Table 3. Distribution ratios of cesium and DB18C6 in the presence of each other between RTIL/aqueous phases as a function of pH at 22 °C (Vo = 0.5 mL; Vw = 5.0 mL).
Table 3. Distribution ratios of cesium and DB18C6 in the presence of each other between RTIL/aqueous phases as a function of pH at 22 °C (Vo = 0.5 mL; Vw = 5.0 mL).
RTILCw0(Cs), mol·dm-3Co0(18C6), mol·dm-3Aqueous pHlogDCsDB18C6logDDB18C6Cs
[THA][DHSS]5.0·10-45.0·10-26.031.562.68
2.851.272.41
2.450.892.20
1.890.742.15
1.060.862.30
0.141.022.68
[TOMA][Sal]5.0·10-45.0·10-26.071.912.62
3.541.862.32
2.771.422.11
2.111.122.05
1.081.522.45
0.131.963.05
[BMIM][N(Tf)2]1.5·10-31.5·10-16.932.282.89
5.092.062.71
3.160.142.57
1.92-0.472.63
0.38-0.602.94
Table 4. Stability Constants of Cesium Complex Formation with 18C6 in RTIL and Extraction Efficiency from Water at 22 °C and pH 5 – 7a.
Table 4. Stability Constants of Cesium Complex Formation with 18C6 in RTIL and Extraction Efficiency from Water at 22 °C and pH 5 – 7a.
SolventlogK1logK2logDCs logDCs18C6logD18C6Cslog[DCs18C6/(DCsD18C6Cs)]
[HMIM][N(Tf)2]4.41.13−1.240.820.251.81
[EtHMIM][N(Tf)2]3.41.16−0.810.56−0.271.64
[BMIM][N(Tf)2]3.41.29−0.671.560.771.46
[BMIM][PF6]2.3 −0.59
−0.59b
−1.17c
−0.20
−0.52d
0.13
0.15e
0.26
[TOMA][Sal]1.45 0.690.830.49−0.38
[THA][DHSS]0.77 1.210.25−0.12−0.84
1,2-dichloroethane7.98f2.58f−7.7g0.8h0.03i6.93
a Stability constants for RTILs are taken from [21] for 25 °C, K1 = [Cs(18C6)][Cs]−1[18C6]−1 K2 = [Cs(18C6)2][Cs(18C6)]−1[18C6]−1; b Ref. [12]; c Ref. [9], pH = 7; d Ref. [9], Co0(18C6) = 1.0 × 10−1 mol·dm-3; e Ref. [9], Co0(18C6) = 1.0 × 10−1 mol·dm-3, blank experiment (without CsNO3); f Ref. [23]; g Ref. [2], picrate salt; h Ref. [3], picrate salt; i Ref. [1].

Share and Cite

MDPI and ACS Style

Vendilo, A.G.; Djigailo, D.I.; Smirnova, S.V.; Torocheshnikova, I.I.; Popov, K.I.; Krasovsky, V.G.; Pletnev, I.V. 18-Crown-6 and Dibenzo-18-crown-6 Assisted Extraction of Cesium from Water into Room Temperature Ionic Liquids and Its Correlation with Stability Constants for Cesium Complexes. Molecules 2009, 14, 5001-5016. https://doi.org/10.3390/molecules14125001

AMA Style

Vendilo AG, Djigailo DI, Smirnova SV, Torocheshnikova II, Popov KI, Krasovsky VG, Pletnev IV. 18-Crown-6 and Dibenzo-18-crown-6 Assisted Extraction of Cesium from Water into Room Temperature Ionic Liquids and Its Correlation with Stability Constants for Cesium Complexes. Molecules. 2009; 14(12):5001-5016. https://doi.org/10.3390/molecules14125001

Chicago/Turabian Style

Vendilo, Andrey Grigoryevich, Dmitry Ivanovich Djigailo, Svetlana Valeryevna Smirnova, Irina Ivanovna Torocheshnikova, Konstantin Ivanovich Popov, Vladimir Georgyevich Krasovsky, and Igor Vladimirovich Pletnev. 2009. "18-Crown-6 and Dibenzo-18-crown-6 Assisted Extraction of Cesium from Water into Room Temperature Ionic Liquids and Its Correlation with Stability Constants for Cesium Complexes" Molecules 14, no. 12: 5001-5016. https://doi.org/10.3390/molecules14125001

Article Metrics

Back to TopTop