• Aucun résultat trouvé

The role of association of ions in ionic liquid/molecular solvent mixtures on metal extraction

N/A
N/A
Protected

Academic year: 2021

Partager "The role of association of ions in ionic liquid/molecular solvent mixtures on metal extraction"

Copied!
7
0
0

Texte intégral

(1)

Cite this: Phys. Chem. Chem. Phys., 2017, 19, 28834

The role of association of ions in ionic liquid/

molecular solvent mixtures on metal extraction†

J.-M. Andanson, *aN. Papaiconomou,bcdP.-A. Cable,aM. Traı¨kia,aI. Billardcdand P. Hussona

Several mixtures of butyl acetate (BA) and an ionic liquid (1-methyl-3-octylpyrrolidinium bis(trifluoromethyl-sulfonyl)imide, [C1C8Pyrro][NTf2], 1-octyl-pyridinium bis(trifluoromethylsulfonyl)imide, [C8Pyr][NTf2],

1-butyl-3-methyl-imidazolium bis(trifluoromethylsulfonyl)imide, [C1C4Im][NTf2] or 1-octyl-3-methyl-imidazolium

bis(trifluoromethylsulfonyl)imide, [C1C8Im][NTf2]) have been characterized by measuring density, viscosity and

conductivity, from pure BA to pure ILs at 298.15 K. The ionicity of these mixtures has been determined on the basis of electrical conductivity and NMR spectroscopy. IR spectra of these mixtures were used to examine the interactions between ions. In parallel, Pt(IV) was extracted from the acidic aqueous phase towards mixtures of BA and [C1C8Im][NTf2] over the entire composition range. A drastic modification in the

distribution coefficient of Pt(IV) was observed at ca xAB= 0.8. A drop was also observed in the ionicity of the

extracting phase (IL + BA) at a similar composition. The variation of the distribution coefficient is ascribed to changes in the interactions in the mixtures, which in turn could induce changes in the extraction mechanism.

1. Introduction

Liquid–liquid extraction (LLE) is an efficient technique to recover metals.1In most traditional LLE experiments, the starting phase is an acidic aqueous phase and the extracting phase is composed of a molecular neutral solvent (kerosene, chloroform, etc.) in which a solute is dissolved (tributylphosphate, calixarenes, phosphine oxides, crown-ethers etc.). Most molecular solvents are unable to extract metal ions by themselves, and thus solutes are often viewed as crucial additives that are able to expel these undesirable water molecules surrounding the metal complexes.1 However, some extraction systems are different. For example, tributylphosphate (TBP), being liquid at room temperature, is a very efficient and neat solvent for metal ions extraction. However, because TBP is too efficient as a pure solvent to allow back extraction, it is used as a solute dissolved in the solvent dodecane for actinide separation2,3 (the usual proportion is 1.1 M TBP in dodecane, i.e. a molar fraction of 0.26 for TBP). The addition of a ‘‘modifier’’ in order to avoid the formation of a third phase driven by the solute is another example, wherein TBP, the modifier, is added to a diglycolamide,

the solute-extracting agent, into TPH (hydrogenated tetrapropylene), the solvent.4As this modifier does not have any specific effect on the extraction itself, the term co-solvent is sometimes used.

Ionic liquids (ILs), a new class of solvents only composed of ions and with melting temperatures below 100 1C, have brought about an entirely new perspective in metallic ion extraction. It is now well-established that various metallic ions can be efficiently extracted by neat ILs.5–8 Therefore, pure solvents efficient for extracting metal complexes cannot be considered as an exception anymore but a common phenomenon. A puzzling aspect of ILs comes from extraction experiments involving a molecular solvent, 1,2-dichloroethane, a phosphine oxide and an IL, 1-butyl-3-methyl-imidazolium hexafluorophosphate, [C1C4 Im]-[PF6].9 The molecular solvent or the IL alone is inefficient for extracting lanthanide ions and so is the IL as a solute in dichloro-ethane. Phosphine oxide as a solute in dichloroethane or in IL allows extraction of all the lanthanide series. Finally, mixing all the three compounds leads to extraction efficiencies higher than any of the other combinations. Turanov’s group investigated other systems of the same type, but employing other molecular solvents (nitro-benzene, chloroform, etc.), other ligands (diamide or other organo-phosphorous compounds), another IL, 1-butyl-3-methyl-imidazolium bis(trifluoromethylsulfonyl)imide, [C1C4Im][Tf2N], and other metals (Ca, Sr, and Ba).10–12 They discussed the results on the basis of chemical interactions, such as a synergistic effect between the classical extracting agent and the IL, indicating that the IL acts as a source of hydrophobic anions in the extraction system.10

aUniversite´ Clermont Auvergne, CNRS, SIGMA Clermont,

Institut de Chimie de Clermont-Ferrand, F-63000 Clermont–Ferrand, France. E-mail: j-michel.andanson@uca.fr

bUniv. Savoie, LEPMI, F-73000, Chambe´ry, France cUniv. Grenoble-Alpes, LEPMI, F-38000, Grenoble, France dCNRS, LEPMI, F-38000, Grenoble, France

†Electronic supplementary information (ESI) available. See DOI: 10.1039/ c7cp05886a Received 29th August 2017, Accepted 9th October 2017 DOI: 10.1039/c7cp05886a rsc.li/pccp

PCCP

PAPER

(2)

The different extraction mechanisms observed with various extracting phases, as summarized above, could be induced by different molecular organizations. A succession of different molecular structures, covering the entire composition range from the neat IL to the pure molecular component, has been reviewed in literature.13In a pure IL, each ion is surrounded by several counterions creating an ionic 3-dimensional network. The addition of a molecular solvent perturbs this organization up to the point where the network is disrupted (percolation limit) and ionic clusters are formed. At lower IL concentrations, these aggregates become smaller and smaller up to the for-mation of isolated ions surrounded by the solvent at the lowest concentrations. However, it is difficult to observe this structural evolution experimentally,14,15for which molecular simulation is an efficient tool. As an example, the organization of mixtures of ethyl-methylimidazolium ethylsulfate and water as a function of the composition was reported by Bernades et al.16

The objective of the present study is thus to correlate the existence of different extraction regimes with modifications of the molecular organization of the extracting phase. To gain a more profound understanding of the interactions in ILs/molecular solvent mixtures and to quantify their impact on the extraction of metal ions, we performed the LLE experiments for Pt(IV) present in

an aqueous solution of hydrochloric acid towards water-immiscible mixtures of butyl acetate (BA) and 1-methyl-3-octylimidazolium bis(trifluoromethanesulfonyl)imide, [C1C8Im][NTf2]. This system was chosen because extraction of Pt(IV) using this pure IL has

previously been extensively studied in some recent articles.17–19 Various mixtures of compositions ranging from pure IL to pure BA were considered. The extraction efficiency was evaluated for each extracting mixture. In parallel, the molecular structure and organization of the extracting phase were assessed through the measurement of the association of the IL and spectroscopic investigations. Other mixtures composed of BA and a hydrophobic IL have also been studied for comparison. Two ILs containing the same octyl chain but with various cations (1-methyl-3-octyl-pyrrolidinium bis(trifluoromethylsulfonyl)imide, [C1C8 Pyrro]-[NTf2], 1-octyl-pyridinium bis(trifluoromethylsulfonyl)imide, [C8Pyr][NTf2]) and one with a shorter alkyl chain (1-butyl-3-methyl-imidazolium bis(trifluoromethylsulfonyl)imide, [C1C4 Im]-[NTf2]) were considered. All these ILs were chosen because they are fully miscible with BA, are immiscible with water and present a reduced viscosity. The addition of BA to the IL is expected to modify the interactions in the system, as evidenced via the ionicity and spectroscopy results, and thus influence the mechanism of extraction of Pt(IV), which is under the stable [PtCl6]2form under our chemical conditions.20

2. Materials and methods

2.1 Materials

The four ILs were purchased from IoLiTec with a minimum stated purity of 98 wt%. Their chemical structures are presented in Fig. 1. BA was purchased from Fluka (499% purity). Before use, all IL samples were degassed and dried under vacuum for

24 hours. Their water content was measured by coulometric Karl Fischer titration (Mettler Toledo DL31) just after this conditioning step. Typical water contents below 50 ppm were measured. 2.2 Methods

The IL/BA mixtures were prepared by weighing each component, taking care to limit the headspace above the liquid mixture, which would lead to differential evaporation. The composition of the mixture was determined gravimetrically with an uncertainty of 0.0001 g. The error on the mole fraction was estimated to be 2 104.

The density was measured using a vibrating-tube densimeter (DMA 5000M) following procedures described previously.21The precision of the density measurement is5  105g cm3. The viscosity was measured using a microviscosimeter (Lovis 2000ME, Anton Paar) based on the falling ball principle. The procedure followed was previously described.22The uncertainty in the viscosity is2%.

The electrical conductivity was measured by impedancemetry using an impedance analyzer (7260, Material Mates) and a sealed conductivity cell in borosilicate glass with two platinum electrodes (Material Mates). Details of the procedure were previously described.22 The uncertainty of the conductivity measurements is1%.

ATR-FTIR spectra were recorded using a Nicolet 380 FT-IR spectrometer equipped with a single reflection diamond ATR cell from Specac and a DTGS detector. IR spectra were averaged for 64 scans with a resolution of 2 cm1. All spectra were measured successively with the same background in order to precisely observe the evolution of the position of the naCF3 band of the NTf2anion (asymmetric stretching mode) around 1200 cm1. Each top of the naCF3 band was fitted with a Gaussian function and the center of this function was used to define the precise position of the band.

The pulsed-field gradient spin-echo NMR technique was used to measure the self-diffusion coefficients of cation, anion and BA by observing the13C nucleus after decoupling the1H nucleus. A Bruker Avance 500 spectrometer operating at 500.13 MHz for1H and 125.3 MHz for13C was used. The mixture was not diluted with a deuterated solvent to avoid any modification of its structure. It was Fig. 1 Names and formula of the four ILs studied.

(3)

placed in a NMR tube with an insert filled with deuterated DMSO. Details of the experimental procedure were previously described.23

In view of the extraction experiments, eight solutions consisting of different mixtures of BA and [C1C8Im][NTf2] were prepared as detailed above. An initial aqueous solution containing 1 M HCl and 103M H2PtCl6was also prepared. Extraction was then carried out in a tube by mixing 2 mL of an aqueous solution with 2 mL of a mixture BA/[C1C8Im][NTf2]. The tube was left on a shaker for 12 h and then centrifuged at 3000g for 20 minutes. After 2 h of settling, the aqueous phase was analysed by UV-Vis spectroscopy (Cary 50, Varian) to estimate the initial and final concentrations of PtCl62in the aqueous phases.

The Beer–Lambert’s law was applied at the maximum absorption wavelength of the PtCl62 metal complex (lmax = 262 nm) using calibration solutions ranging from 10 to 100 mM. A correlation factor of 0.999 for the Beer–Lambert’s law was obtained.

The value of the distribution coefficient, D, is calculated as follows: D¼ PtCl6 2  Init  PtCl62  Final PtCl62 ½ Final Vw VOrg (1)

where [PtCl62] refers to the concentration of PtCl62in water and the superscripts Init and Final refer to the values measured prior and after the extraction experiment, respectively. Vwand VOrg stand for the volumes of water and mixtures of IL/BA before the extraction, respectively. When extraction experiments were carried out using pure IL (xBA= 0) or pure butyl acetate (xBA= 1), volume ratios Vw/VOrgwere 2.4 and 1, respectively. In all other cases, Vw/VOrgwas 1.5. The maximum D value measurable in this study by UV-Vis spectroscopy was assumed to be 2 102. Given the uncertainty of the spectroscopic measurement, the relative uncertainty for D is5%.

3. Results

3.1 Viscosity

The viscosity of IL/BA mixtures was measured as a function of the composition as illustrated in Fig. 2. The experimental data are given in ESI† (Table S1). At 298.15 K, the pure ILs have viscosities varying from 50 mPa s for [C1C4Im][NTf2] to 138 mPa s for [C1C8Pyrro][NTf2]. The increase in the alkyl chain on the imidazolium ring leads to an increase in the viscosity (from 50 to 90 mPa s from 4 to 8 carbon atoms on the imidazolium ring) as already reported in literature.24–26Changing the cationic ring also influences the viscosity: the IL with a pyrrolidinium cation is more viscous than the one with a pyridinium cation, and more viscous than the one with an imidazolium ring. The addition of BA to the IL sharply decreases the viscosity of the system as observed for most molecular solvents added to an IL.27–29 The present experimental data exhibit acceptable deviations with the set of data measured by Malek et al.30on the system [C1C4Im][NTf2] + BA. No other data were found in literature for comparison.

3.2 Conductivity

Despite the fact that neat ILs are only composed of ions, their conductivities are often low because of high viscosity and ionic association. Mixing an IL with a molecular component creates new interactions and modifies the association between ions. The electrical conductivity was measured as a function of the composition for all the IL/BA mixtures from the pure IL to the pure BA as illustrated in Fig. 3. The conductivities measured for the mixtures containing [C1C4Im][NTf2] are systematically higher than those of the other ILs, which is coherent with a smaller viscosity because of the shorter alkyl chain. The three pure ILs containing an octyl chain present comparable electrical conductivities: 0.82 mS cm1for [C1C8Pyrro][NTf2], 0.97 mS cm1 for [C8Pyr][NTf2] and 1.31 mS cm1for [C1C8Im][NTf2]. This is also true when they are mixed with BA. The observed order of electrical conductivity exactly matches the inverse order of viscosity, indicating that comparable ionic associations probably occur in these mixtures. Fig. 2 Viscosity of the ILs–BA mixtures as a function of the composition at 298.15 K. Lines are guides to the eyes.

Fig. 3 Electrical conductivity of the mixtures ILs–BA as a function of the composition at 298.15 K. Lines are guides to the eyes.

(4)

The conductivity presents a maximum as a function of the composi-tion. This was previously reported in literature with other molecular solvents23and is explained by a balance between the decrease in charge density and an increase in the mobility of the ions during dilution. For all the considered mixtures, the maximum is observed for a BA mole fraction of around 0.70–0.75 (corresponding to circa 0.4 mass fraction of BA).

3.3 Extraction of Pt(IV) using [C1C8Im][NTf2]

The distribution coefficients for PtCl62obtained in this study are plotted against the mole fraction of BA in the extracting phase, as shown in Fig. 4. It appears that the distribution coefficient decreases with xBA. The value of 23 observed in pure [C1C8Im][NTf2] is slightly higher than previous values reported in the literature,19 but this is most probably related to the concentration of HCl that was slightly lower than that used in previous studies. The fact that no extraction occurs in pure BA is in agreement with the fact that Pt(IV) is not extracted towards a molecular solvent when no extracting agent is used.1

When reporting the distribution coefficients as a function of the composition, several domains can be observed. First, D decreases from 23 in pure IL to approximately 10 at xBA= 0.5. Between 0.5 and 0.8, an average value of 10 0.3 is obtained. In this region, and within experimental uncertainties, the amount of BA appears to have no major influence on the extraction of PtCl62. Finally, a sharp decrease for D is depicted between xBA= 0.8 and 1.

4. Discussion

4.1 Molecular structure and interactions

Angell et al.31and MacFarlane et al.32first revealed that ion pair formation is responsible for the low conductivities observed in ILs. Subsequently, they introduced the concept of ionicity as the relative proportion of free ions in an IL, and this was discussed in terms of a deviation to the Nernst Einstein equation. Walden

plots reported the molar conductivity as a function of fluidity. In these graphs, Angell et al.31defined different domains corresponding to good (highly dissociated) and poor (highly associated) ILs.

Fig. 5 presents the Walden plots obtained for the organic solvent/IL mixtures in this study. The conductivities of pure ILs (lowest Z1) are just below the reference line (KCl 0.01 mol L1), meaning the ILs are not fully dissociated but can be considered as good ILs, according to the classification of Angell. In agree-ment with previous results,33the three ILs with an octyl chain have similar ionicities, while [C1C4Im][NTf2] appears to be more dissociated. The evolution of deviation from the reference line, as a function of composition, allows following the evolution of ionic association when diluting the IL in BA. A progressive increase in ionic association is observed on adding BA, this being even more pronounced for a BA mole fraction above 0.8.

Such a trend was already observed for several other mixtures of ILs with molecular solvents in previously reported studies. The addition of a solvent with low dielectric constant (e.g. chloroform, e = 4.81; dichloromethane, e = 9.08; tetrahydro-furan, e = 7.58) to different imidazolium-based ILs with NTf2 anions leads to an increase in the ionic association, as observed by several authors using Walden plots,34,35FTIR36or NMR37,38 spectroscopy. The addition of an alcohol to an IL based on tetrafluoroborate anions and N-alkylpyridinium cations also increases the ionic association of the system.39 In all these cases, a similar behavior to the present results was obtained: strong ionic association when IL is highly diluted. On the contrary, water added to imidazolium methylsulfate-based ILs tends to promote ionic dissociation with a substantial increase in the ionicity depicted for water mole fractions up to 0.8.23 This can be explained by the important interactions between water molecules and the ions (high dielectric constant of water, H-bonds).

A more quantitative approach to ionicity was developed by Ueno et al.40Ionicity is obtained by dividing the molar conductivity measured by impedancemetry, Limp, by the molar conductivity Fig. 4 Distribution coefficients for PtCl6

2

towards mixtures of [C1C8Im][NTf2] and BA as a function of the mole fraction of BA. Dotted

line is the average of experimental values for xBAover the range 0.5–0.8.

Fig. 5 Walden plots of the ILs–BA mixtures at 298.15 K. Black line corresponds to the ideal dissociated salt (KCl 0.01 mol L1). Lines between experimental points are for clarity.

(5)

estimated from NMR self-diffusion coefficients, LNMR, which is calculated using the Nernst–Einstein equation:40

LNMR¼ NAe2

þ D

kBT

(2) where NAis the Avogadro’s number, e is the electronic charge on each ionic carrier, D+and Dare the diffusion coefficients of the ions, kB is the Boltzmann constant and T is the temperature. Ionicity (Limp/LNMR) can be viewed as the ratio between the mobility of the charged species and that of the non-charged species. The ionicities, calculated for the mixtures of [C1C8Im][NTf2] or [C1C8Pyrro][NTf2] and BA, are presented in Fig. 6.

The two pure ILs present low ionicities (below 0.5), which is coherent with the long alkyl chain on the cations.40 The imidazolium cation leads to lower ionicities than pyrrolidium, which was already observed by Tokuda et al.41for ILs with butyl chains. As already observed in Walden plots, the dilution of the IL with BA decreases the ionicity, particularly for BA mole fractions above 0.8.

Walden plots and molecular conductivity ratio are the two most common approaches to discuss ionicity, even though they only permit to discriminate between two groups of organiza-tion: the charged groups and the neutral species. Real systems can be complex, with the presence of neutral clusters/ion pairs, ionic aggregates or isolated ions. Other experimental techni-ques as spectroscopic technitechni-ques (NMR, IR, UV-Vis) can provide

valuable information on the formation of these more complex structures.13

As can be observed on the IR spectra given in ESI,† the naCF3 (asymmetric stretching mode) around 1200 cm1 is the most intense band of the IR spectrum in NTf2 based ILs.42 The position/shape of this band is particularly sensitive to the environment of the anion. Interaction between an imidazolium cation and the NTf2anion impacts this vibrational mode.43The position of the naCF3band shifts with the concentration of BA as well as other strong bands of the anion at 1350 (naSO2) and 1140 cm1 (nsSO2), but the naCF3 band shifts the most. The shifts observed experimentally for these bands are consistent with previous DFT calculations with C1C8Im cation and NTf2 anion; the bigger shift induced by anion–cation interactions is for the naCF3band.44Cation bands are not as intense in the IR spectra and often overlap with BA bands. This makes analyzing them to observe the evolution of the environment less accurate. Thus, a careful analysis of the position of the naCF3band was conducted for the two ILs that have been studied by pulsed-field gradient spin-echo NMR. The naCF3bands are shifted by approximately 10 cm1on increasing the concentration of BA, as shown in Fig. 6. The evolution of the position of this band is similar to the evolution of ionicity. This indicates that the evolution of this band is mainly due to the decrease in the interactions between the anion and cation rather than the BA molecules influencing the naCF3 band strongly. Analyzing the shifts in IR spectra has the advantage that it is simpler than calculating the ionicity. Furthermore, the results are much less noisy. The evolution of the shift is relatively limited for con-centrations below 0.5 of BA, while this evolution is more significant for concentrations above 0.75 of BA.

These results give a new insight into the concept of solvent and solute in mixtures of ILs with an organic solvent. When xBA is typically below 0.5, the ionicity of the IL in the mixture is close to that of pure IL. One can then consider the mixture as a solution of BA, the solute, in an IL, the solvent, with its reference state being the pure IL. In contrast, at high xBA values, typically above 0.75 for [C1C8Im][NTf2], the ionicity decreases sharply. This is in agreement with the fact that salts are associated in organic solvents exhibiting a low dielectric constant. In this case, the IL therefore, behaves more like an organic salt dissolved in an organic solvent. The IL is thus the solute and BA is the solvent.

As detailed below, having an IL behaving as a solvent or as a solute is expected to have significant consequences on a variety of applications such as the extraction of metal ions from aqueous phases.

4.2 Extraction of PtCl62

Previous reports have discussed the extraction of anionic com-plexes such as PtCl62 towards an organic solvent, such as kerosene or dodecane, containing an extracting agent, such as trioctylamine.45In the case of Aliquat 33646(a mixture of quaternary ammonium salts with mainly octyl and decyl chains), it was shown that the extraction of the targeted anionic complex increases with the concentration of the extracting agent. The extraction Fig. 6 Evolution of ionicity (top) and of the position of thenaCF3band in

the IR spectra of IL–BA mixtures (bottom) as a function of the composi-tion. Lines between experimental points are for clarity.

(6)

mechanism usually accepted is based on an anionic exchange, as follows:

2 R½ 4N½Cl þ PtCl62! R½ 4N2½PtCl6 þ 2Cl: (3)

where [R4N][Cl] stands for Aliquat 336. Pure ILs phases17,19,47,48 have also been investigated as the extracting phase. The extrac-tion of the negatively stable20charged complex [PtCl

6]2towards [C1C8Im][NTf2], for instance, revealed that the distribution coefficient decreased with the concentration of PtCl62initially added into the aqueous phase and of HCl present in the aqueous phase.19 Extraction was then explained by the formation of a hydrophobic ion pair between the anionic complex and the cation of the IL dissolved in water.

2 C½ 1C8Imþ þ PtCl 62! C½ 1C8Im2½PtCl6: (4)

We propose that the two mechanisms detailed above occur simultaneously under our chemical conditions. The first extrac-tion mechanism is predominant when a neat IL is used or when the IL can be described as the solvent, that is, when its ionicity is close to that of the pure IL. On the contrary, the second extraction mechanism is predominant when the IL is the solute. The xBAvalue at which one mechanism supersedes the other is xBA= 0.75, because the clear changes observed in Fig. 4 for the extraction of PtCl62and for the ionicity occur at this specific value.

When xBA4 0.75, the IL can be considered as an (mostly associated) organic salt dissolved in BA. In this case, the extraction mechanism for PtCl62follows eqn (3). That is, the cation of the IL present in the organic phase associates with the metal complex to form a hydrophobic pair. Consequently, D is expected to decrease if the concentration of the IL cation in the organic phase decreases. Such is the case, since D decreases as xBAincreases.

Over the region xBAo 0.75, on the contrary, the IL can be defined as the solvent, as discussed above. In this case, extrac-tion of PtCl62is expected to occur following eqn (4). That is, a cation of the IL dissolved in water associates with the anionic complex forming a hydrophobic pair that is extracted towards the non-aqueous phase. The decrease in D with xBAwould thus be due to fewer cations from the IL being available in the aqueous phase, depending on the xBAvalue. Actually, several studies have reported the great sensitivity of IL cation and anion aqueous solubilities on the exact compositions of the two phases composing the extraction system, (acid nature and concentration for the aqueous phase, nature and concentration of the extractant in the organic phase).49–54

5. Conclusion

The originality of the present study is to propose a link between the ionicity concept and the extraction efficiency of the liquid in a particular application. Physico-chemical measurements (density, conductivity, viscosity, NMR and IR spectroscopy) of several mixtures between BA and ILs allow going beyond the notion of ‘‘solvent’’ and ‘‘solute’’. Our results give access to the

identification of a change in the ionicity (NMR and electrical conductivity), which is correlated with a change in the inter-molecular interactions as observed via IR results. This change impacts the behavior of these mixtures, by modification of the extraction mechanism of an anionic complex, PtCl62, taken as an example of a general trend. It would be interesting to check if such measurements would correlate with extraction results (to be obtained) involving more complex mixtures, composed of a molecular solvent, an IL and an extracting agent under different proportions.

From more of an application perspective, our results show that a neat IL phase is not the only way to obtain reasonable extraction. Although the system under study here was not optimized in view of efficiency, it is clear that adding BA, a rather benign molecular solvent, decreases viscosities and also diminishes the cost of the entire process, without making it too inefficient. Moreover, the rather large range of molar fractions (ca. 0.3–0.75), within which the extraction is constant, offers a robust aspect that should not be neglected in view of industrial applications.

Conflicts of interest

There are no conflicts to declare.

References

1 J. Rydberg, M. Cox, C. Musikas and G. R. Choppin, Solvent extraction principles and practice, revised and expanded, Dekker, New York, 2004.

2 K. L. Nash and G. R. Choppin, Sep. Sci. Technol., 1997, 32, 255. 3 J. Mathur, M. Murai and K. Nash, Solvent Extr. Ion Exch.,

2001, 19, 357.

4 G. Modolo, H. Asp, C. Schreinemachers and H. Vijgen, Solvent Extr. Ion Exch., 2007, 25, 703–721.

5 C. Deferm, M. V. D. Voorde, J. Luyten, H. Oosterhof, J. Fransaer and K. Binnemans, Green Chem., 2016, 18, 4116–4127.

6 T. Vander-Hoogerstraete, S. Wellens, K. Verachtert and K. Binnemans, Green Chem., 2013, 15, 919–927.

7 N. Papaiconomou, I. Billard and E. Chainet, RSC Adv., 2014, 4, 48260–48266.

8 I. Billard, in Handbook on the Physics and Chemistry of Rare Earths, ed. J. C. G. Bu¨nzli and V. K. Percharsky, Elsevier, 2013, vol. 43, ch. 256.

9 A. N. Turanov, V. K. Karandashev and V. E. Baulin, Radio-chemistry, 2008, 50, 266–273.

10 A. N. Turanov, V. V. Karandashev and V. E. Baulin, Solvent Extr. Ion Exch., 2008, 26, 77–99.

11 A. N. Turanov, V. K. Karandashev and V. E. Baulin, Solvent Extr. Ion Exch., 2010, 28, 367–387.

12 A. N. Turanov, V. K. Karandashev and V. E. Baulin, Russ. J. Inorg. Chem., 2008, 53, 970.

13 H. K. Stassen, R. Ludwig, A. Wulf and J. Dupont, Chem. – Eur. J., 2015, 21, 8324.

(7)

14 S. T. Keaveney, K. S. S. M. Hale, J. W. Stranger, B. Ganbold, W. S. Price and J. B. Harper, ChemPhysChem, 2016, 17, 3853–3862.

15 D. Ivanov, N. Petrov, O. Klimchuk and I. BIllard, Chem. Phys. Lett., 2012, 551, 111–114.

16 C. E. S. Bernades, M. E. M. d. Piedade and J. N. C. Lopes, J. Phys. Chem. B, 2011, 115, 2067.

17 S. Ge´nand-Pinaz, N. Papaiconomou and J. M. Leveque, Green Chem., 2013, 15, 2493–2501.

18 N. Papaiconomou, L. Svecova, C. Bonnaud, L. Cathelin, I. Billard and E. Chainet, Dalton Trans., 2015, 44, 20131–20138.

19 N. Papaiconomou, C. Iojioiu, L. Cointeaux, E. Chainet and I. Billard, ChemistrySelect, 2016, 1, 3892.

20 J. Kramer and K. R. Koch, Inorg. Chem., 2007, 46, 7466–7476. 21 A. Andresova, J. Storch, M. Traika, Z. Wagner, M. Bendova

and P. Husson, Fluid Phase Equilib., 2014, 371, 41–49. 22 T. A. Faadeva, P. Husson, J. A. DeVine, M. F. C. Gomes,

S. G. Greenbaum and E. W. Castner, J. Chem. Phys., 2015, 143, 064503.

23 J. M. Andanson, M. Traı¨ka and P. Husson, J. Chem. Thermodyn., 2014, 77, 214.

24 P. Wasserscheid and T. Welton, Ionic liquids in synthesis, Wiley, 2008.

25 G. Yu, D. Zhao, L. Wen, S. Yang and X. Chen, AIChE J., 2012, 58, 2885–2899.

26 R. L. Gardas and J. A. P. Coutinho, AIChE J., 2009, 55, 1274–1290.

27 J. N. C. Lopes, M. F. C. Gomes, P. Husson, A. A. H. Padua, L. P. N. Rebelo, S. Sarraute and M. Tariq, J. Phys. Chem. B, 2011, 115, 6088–6099.

28 Y. Wang, D. Chen and X. OuYang, J. Chem. Eng. Data, 2010, 55, 4878–4884.

29 A. Battacharjee, C. Varanda, M. G. Freire, S. Matted, L. M. N. B. F. Santos, I. M. Marrucho and J. A. P. Coutinho, J. Chem. Eng. Data, 2012, 57, 3473–3482.

30 N. I. Malek and S. P. Ijardar, J. Chem. Thermodyn., 2016, 93, 75.

31 C. A. Angell, N. Byrne and J. P. Belieres, Acc. Chem. Res., 2007, 40, 1228.

32 D. R. MacFarlane, M. Forsyth, E. I. Ozgorodina, A. P. Abbott, G. Annat and K. Fraser, Phys. Chem. Chem. Phys., 2009, 11, 4962.

33 H. Tokuda, K. Ishii, M. A. B. H. Susan, S. Tsuzuki, K. Hayamizu and M. Watanabe, J. Phys. Chem. B, 2006, 110, 2833–2839.

34 D. R. Saeva, J. Petenuci and M. M. Hoffmann, J. Phys. Chem. B, 2016, 120, 9745.

35 E. Cade, J. Petenuci and M. M. Hoffmann, ChemPhysChem, 2016, 17, 520.

36 T. Ko¨ddermann, C. Wertz, A. Heintz and R. Ludwig, Chem-PhysChem, 2006, 7, 1944.

37 J. D. Tubbs and M. M. Hoffmann, J. Solution Chem., 2004, 33, 381.

38 N. T. Sharf, A. Stark and M. M. Hoffmann, J. Phys. Chem. B, 2012, 116, 11488.

39 M. Garcia-Mardones, H. M. Osorio, C. Lafuente and I. Gascon, J. Chem. Eng. Data, 2013, 58, 1613.

40 K. Ueno, H. Tokuda and W. Watanabe, Phys. Chem. Chem. Phys., 2010, 12, 1649.

41 H. Tokuda, S. Tsuzuki, A. B. H. Susan, K. Hayamizu and M. Watanabe, J. Phys. Chem. B, 2006, 110, 19593.

42 I. Rey, P. Johansson, J. Lindgren, J. C. Lasse`gues, J. Grondin and L. Servant, J. Phys. Chem. A, 1998, 102, 3249–3258. 43 O. Ho¨fft, S. Bahr and V. Kempter, Langmuir, 2008, 24,

11562–11566.

44 H. Azizi-Toupkanloo, S. F. Tayyari and P. Nancarrow, J. Iran. Chem. Soc., 2017, 14, 1281–1300.

45 A. Jaree and N. Khunphakdee, J. Ind. Eng. Chem., 2011, 17, 243–247.

46 J. Y. Lee, J. R. Kumar, J. S. Kim and H. S. Yoon, J. Ind. Eng. Chem., 2009, 15, 359–364.

47 S. Ikeda, T. Mori, Y. Ikeda and K. Takao, ACS Sustainable Chem. Eng., 2016, 4, 2459–2463.

48 S. Katsuta, Y. Yoshimoto, M. Okai, Y. Takeda and K. Bessho, Ind. Eng. Chem. Res., 2011, 50, 12735–12740.

49 D. Ternova, M. Boltoeva, L. Cointeaux, C. Gaillard, V. Kalchenko, V. Mazan, S. Miroshnichenko, P. K. Mohapatra, A. Ouadi, N. Papaiconomou, M. Petrova and I. Billard, J. Phys. Chem. B, 2016, 120, 7502–7510.

50 M. Atanassova, V. Mazan and I. Billard, ChemPhysChem, 2015, 16, 1703–1711.

51 V. Mazan, I. Billard and N. Papaiconomou, RSC Adv., 2014, 4, 13371–13384.

52 V. Mazan, M. Y. Boltoeva, E. E. Tereshatov and C. M. Folden, RSC Adv., 2016, 6, 56260–56270.

53 P. G. Rickert, D. C. Stepinski, D. J. Rausch, R. M. Bergeron, S. Jakab and M. L. Dietz, Talanta, 2007, 72, 315–320. 54 K. E. Gutowski, G. A. Broker, H. D. Willauer, J. G. Huddleston,

R. P. Swatloski, J. D. Holbrey and R. D. Rogers, J. Am. Chem. Soc., 2003, 125, 6632.

Figure

Fig. 3 Electrical conductivity of the mixtures ILs–BA as a function of the composition at 298.15 K
Fig. 5 Walden plots of the ILs–BA mixtures at 298.15 K. Black line corresponds to the ideal dissociated salt (KCl 0.01 mol L 1 )

Références

Documents relatifs

The photophysics of 3,3'-diethylthiacarbocyanine iodide (DTCI), as a fluorescence probe, in liquid mixtures of dimethyl sulfoxide (DMSO) with either trimethyl(butyl)ammonium

The separation of cobalt from nickel in weakly acid sulphate solutions problem remained unsolved until it was realized that the separation factor for cobalt

Le monde est dirigé par des hommes puissants Qui agissent en secret dans un cercle influant Très souvent x pont mais aussi franc maçon Ils font partie du pacte qui régit les nations

Thus, decreasing solvent polarity also reduces water-crown complexation [L‚H 2 O (org) ], leading to lower k values. This is probably caused by the combination of a lower solubility

In order to elucidate the extraction mechanism in this extraction system, we aim: (i) to investigate the distribution of nitric acid and water between aqueous acidic phase and

To elucidate further the metal extraction equilibria between an aqueous and organic phase composed of TODGA and TBP extractants diluted in the IL or molecular solvent, we examined

To study the effect of the aqueous pH on the IL constituent ions solubility and HBTA deprotonation, their individual concentrations were measured as a function

Dans un contexte victorien où Pater se donnait pour tâche, à l’instar de Carlyle et de Matthew Arnold, de redéfinir la notion de « culture », non seulement