Solvent Extraction of Cesium with a New ... - ACS Publications

Oct 16, 2013 - Peng Zhang , Yaling Wang , and Dongxiang Zhang ... Xin Lu , Dongxiang Zhang , Shengnan He , Jian Feng , Alemtsehay Tesfay Reda , Cong ...
0 downloads 0 Views 1MB Size
Article pubs.acs.org/jced

Solvent Extraction of Cesium with a New Compound Calix[4]arenebis[(4-methyl-1,2-phenylene)-crown-6] Anyun Zhang,*,† Ying Dai,† Lei Xu,† and Zhifang Chai†,‡ †

Department of Chemical and Biological Engineering, Zhejiang University, 38 Zheda Road, Hangzhou 310027, P. R. China Institute of High Energy Physics, Chinese Academy of Sciences, P.O. Box 918, Beijing 100049, P. R. China



ABSTRACT: A new macrocyclic calix[4]arene-bis[(4-methyl-1,2phenylene)-crown-6] (CalixBisMePhC) was synthesized. An extraction study of Cs(I), Na(I), K(I), Rb(I), Sr(II), Ba(II), Pd(II), and Ru(III) with CalixBisMePhC/CHCl3 was investigated by examining the effects of contact time, HNO3 concentration, and temperature. The results showed that due to the effective molecular recognition, CalixBisMePhC/CHCl3 had excellent extraction ability and high selectivity for Cs(I) over the tested metals except Rb(I). The maximum distribution ratio for Cs(I) was found at the HNO3 concentration of 4.4 mol·kg−1. The composition of the extracted species of Cs(I) was determined to be CsNO3·0.5CalixBisMePhC· 0.5HNO3 by a slope method. Some valuable parameters for the Cs(I) extraction were obtained. The results demonstrated that the CalixBisMePhC/CHCl3 extraction system is promising to partition Cs(I) from highly level liquid waste.

1. INTRODUCTION In the PUREX (plutonium uranium reduction extraction) process, both uranium and plutonium are recovered using the 30 % tri-n-butyl phosphate/kerosene extraction system.1 Consequently, the process produces an acidic and highly radioactive solution, called high level liquid waste (HLW), containing fission, transuranic, and activation isotopes such as the long-lived minor actinides MAs(III), rare earth metals REs(III), heat generators Cs(I) and Sr(II), noble metals, Mo(VI), Tc(VII), Zr(IV), and so on.1 Recovery of these metals is valuable to the safe disposal of HLW and a reduction in its potential risk to environment and human health. Although plenty of effort has been made,2−13 industrial partitioning of these metals has been problematic so far. It is known that Cs-137 has properties of high energy gamma ray emission (661.9 keV), heat output (0.42 W/g), and relatively long half-life (t1/2 = 30.1 years), and poses threat to final geological disposal.14−18 Cs-135 has an extremely long half-life (t1/2 = 2.3·106 years) and is also an important contributor to the long-term radiological impact on the deep geological repository.19,20 Therefore, separation of radio cesium from HLW will resolve the radiation exposure problem during radioactive waste handling and can also reduce the risk to the vitrified waste storage. Moreover, the separated Cs-137 can be utilized as resources of β-radiation and energy generators in industrial determination and hospital examination. Separation and recovery of Cs(I) from HLW is therefore meaningful. However, so like the separation of long-lived minor actinides, partitioning of Cs(I) has also been a challenge up to now. The author has been focusing on the separation of Cs(I) and Sr(II), and developed an advanced partitioning technology © XXXX American Chemical Society

entitled SPEC (strontium/cesium partitioning from HLW by extraction chromatography) process. Cs(I) and Sr(II) can be successfully separated from simulated HLW using calix[4]arene-crown/SiO2−P and 4,4′,(5′)-di(tert-butylcyclohexano)18-crown-6 (DtBuCH18C6)/SiO2−P as the stationary phase.21 The stationary phases used in the SPEC process were prepared by impregnating and immobilizing supramolecular recognition reagents into the pores of macroporous SiO2−P particles. Lots of effort has been taken to investigate the adsorption ability of silica-based materials and its adsorption behavior as functions of HNO3 concentration, contact time, and temperature. However the extraction behavior has not been understood as well as the composition of the complex between Cs(I) and supramoleculars. Solvent extraction is one of the most important methods to investigate the interaction between Cs(I) and supramoleculars. In the present work, a new supramolecular compound calix[4]-bis[(4-methyl-1,2-phenyl-ene)-crown-6] (CalixBisMePhC) was synthesized using a new synthetic route developed. The extraction properties of Cs(I) and some typical metals were investigated as functions of the HNO 3 concentration, contact time, and temperature. The optimum HNO3 concentration for the Cs(I) extraction were determined. Separation factors of Cs(I) to other tested metals were obtained. The extraction mechanism and the composition of the extracted species of Cs(I) with CalixBisMePhC were discussed using a slope method. The possibility of the Received: August 13, 2013 Accepted: September 28, 2013

A

dx.doi.org/10.1021/je400735z | J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

separation of Cs(I) by a multistage countercurrent extraction was evaluated.

2. EXPERIMENTAL SECTION 2.1. Reagents and chemicals. Alkali metal nitrates MINO3 (MI = Na, K, Rb, and Cs), and alkaline earth metal nitrates MII(NO3)2 (MII = Sr and Ba) used in the present study were of analytical grade. The liquid reagents, palladium nitrate solution and ruthenium nitrosyl nitrate solution, were provided by the Tanaka Noble Metal Co. Inc., Japan, and Strem Chemicals, U.S.A., respectively. Metals solutions in HNO3 medium were prepared by dissolving designed weights of the above nitrates with HNO3 solution. Deionized water was used for preparation of metals solutions. Calix[4]-bis[(4-methyl-1,2-phenylene)-crown-6] (CalixBisMePhC) as shown in Figure 1 was synthesized by multistep Figure 2. Appearance of CalixBisMePhC.

metals. Then equal volumes (5.0 cm3) of the organic phase and the aqueous phase were mixed in a glass-stopped conical flask and horizontally shaken at 120 rpm for designed contact time in a TAITEC MM-10 model thermostatted water bath maintained at temperatures of 293 K, 298 K, 303 K, 313 K, and 318 K. The mixture was then centrifuged, and the organic phase was removed. The concentrations of the tested metals in the aqueous phase were measured using a Varian AA 240 FS model atomic adsorption spectroscope. The distribution ratio, D, was calculated as follows:

Figure 1. Molecular structure of calix[4]-bis[(4-methyl-1,2-phenylene)-crown-6] (CalixBisMePhC).

reactions. Chloroform, a commercial product of analytical grade, was used as a diluent to dissolve CalixBisMePhC. No other additives or modifiers were used due to the excellent solubility of CalixBisMePhC in chloroform. All other reagents used were of analytical grade. 2.2. Synthesis of CalixBisMePhC. CalixBisMePhC was quantitatively prepared according to the following steps: (1) 5,11,17,23-tetra-butyl-25,26,27,28-tetra-hydroxy-calix[4]arene (TBTHCalix[4]arene) was synthesized by the polymerization reaction between p-tert-butylphenol and formaldehyde. The reaction was catalyzed with sodium hydroxide in diphenyl ether under refluxing in nitrogen atmosphere for 5 h at 533 K.22 (2) The preparation of 25,26,27,28-tetra-hydroxy-calix[4]arene (THCalix[4]arene) was completed using phenol to remove four butyl groups from TBTHCalix[4]arene.23 Anhydrous AlCl3 was added as a dryer to eliminate the microamount of water made from the reaction. (3) Bis[1,2-[2′(2″-hydroxy ethoxy)ethoxy]]-4-methylbenzene was prepared by heating and stirring a suspension of 4-methylcatechol, 2-chloroethoxyethanol, and K2CO3 in dimethylformamide at 353 K under an atmosphere of nitrogen overnight.24 (4) The synthesis of bis[1,2-[2′(2″-hydroxyethoxy)ethoxy]]-4-methylbenzene di-ptoluenesulfonate was carried out by reacting bis[1,2-[2′(2″hydr-oxyethoxy)ethoxy]]-4-methylbenzene with p-toluenesulfonyl chloride by the esterification reaction at low temperature.25 (5) CalixBisMePhC was synthesized through 1,3alternate cyclization of THCalix[4]arene with bis[1,2-[2′(2″hydroxyethoxy)ethoxy]]-4-methyl-benzene di-p-toluenesulfonate in acetonitrile.26 It was catalyzed by cesium carbonate under refluxing in nitrogen atmosphere for 17 days. The resultant CalixBisMePhC shown in Figure 2 was then characterized by elementary analysis, 1H NMR, and ESI-MS. 2.3. Extraction Procedure. Prior to extraction experiment, the organic phase was preconditioned via one contact with equal volume of an appropriate acid solution containing no

D = C(o)/C(a)

(1)

where C(o) and C(a) represent the equilibrium concentrations of the tested metals in the organic phase and aqueous phase, respectively. Separation factor of Cs(I) to the other tested metals, SFCs/M, was calculated as follows: SFCs/M = DM1/DM2

(2)

3. RESULTS AND DISCUSSION 3.1. Dependence of Extraction of the Tested Metals on Contact Time. A supramolecular recognition agent generally has strong complexation with a metal ion that fits well in its cavity. CalixBisMePhC in 1,3-alternate conformation is assembled by one calix[4]arene and two crown ethers and contains two cavities capable of recognizing a metal ion. This makes possible the separation and concentration of a metal ion from others. The effect of contact time on the extraction was investigated to well understand the recognition process of CalixBisMePhC. Some typical fission and nonfission elements, Na(I), K(I), Rb(I), Cs(I), Sr(II), Ba(II), Pd(II), and Ru(III) were extracted by CalixBisMePhC/CHCl3 for different contact time in 4.4 mol·kg−1 HNO3 medium. The results are listed in Table1. As shown in Table 1, the distribution ratio of Cs(I), DCs, increased with the increase in contact time and the equilibrium reached in 15 min. Above it DCs maintained a stable level up to 200 min. This indicated that CalixBisMePhC/CHCl3 had fast extraction dynamics for Cs(I). It was ascribed to the effective complexation of CalixBisMePhC toward Cs(I). Meanwhile, the tested Na(I), Sr(II), Ba(II), Pd(II), and Ru(III) with B

dx.doi.org/10.1021/je400735z | J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Table 1. Comparison of the Distribution Ratio of the Tested Metals with the Change in Contact Time Csa

distribution ratio D

time min

Na(I)

K(I)

Rb(I)

Cs(I)

Sr(II)

Ba(II)

Pd(II)

Ru(III)

%

5 10 15 25 40 60 200

0.01 0.04 0.02 0.02 0.02 0.02 0.01

0.12 0.13 0.12 0.17 0.18 0.17 0.18

0.53 0.58 0.59 0.63 0.62 0.59 0.62

1.50 1.64 1.79 1.76 1.79 1.79 1.78

0.04 0.06 0.04 0.05 0.06 0.08 0.05

0.00 0.00 0.01 0.01 0.01 0.01 0.00

0.02 0.02 0.03 0.06 0.04 0.04 0.05

0.04 0.04 0.00 0.02 0.02 0.06 0.03

60.0 62.1 64.1 63.8 64.1 64.2 64.1

a Cs(I) recovery percentage (%). Conditions: T = 298.1 K, [CalixBisMePhC](o) = 5.4·10−3 mol·kg−1, [Metal](a) = 5.0·10−4 mol·kg−1, [HNO3](a) = 4.4 mol·kg−1, phase ratio =1:1. The expanded uncertainty of the result with level of confidence = 0.95.

Table 2. Comparison of the Distribution Ratio of the Tested Metals with the Change in the HNO3 Concentration of in Aqueous Phase mol·kg

−1

0.4 1.0 1.9 2.8 3.6 4.4 5.1

Csa

distribution ratio D

HNO3 concn Na(I)

K(I)

Rb(I)

Cs(I)

Sr(II)

Ba(II)

Pd(II)

Ru(III)

%

0.02 0.04 0.02 0.02 0.02 0.01 0.01

0.10 0.08 0.12 0.14 0.14 0.18 0.17

0.26 0.25 0.32 0.41 0.45 0.62 0.52

0.45 0.50 0.97 1.16 1.51 1.78 1.74

0.15 0.00 0.00 0.01 0.00 0.05 0.04

0.00 0.00 0.00 0.00 0.00 0.00 0.00

0.00 0.00 0.00 0.04 0.01 0.05 0.03

0.00 0.01 0.00 0.00 0.03 0.03 0.04

31.0 33.1 49.3 53.8 60.2 64.1 63.5

Cs(I) recovery percentage (%). Conditions: T = 298.1 K, [CalixBisMePhC](o) = 5.4·10−3 mol·kg−1, [Metal](a) = 5.0·10−4 mol·kg−1; contact time = 200 min; phase ratio = 1:1. The expanded uncertainty of the result with level of confidence is 0.95. a

Table 3. Separation Factor (SF) of Cs(I) to Other Metals at Different HNO3 Concentrationsa HNO3 concn (mol·kg−1)

SFCs/Na

SFCs/K

SFCs/Rb

SFCs/Sr

SFCs/Ba

SFCs/Pd

SFCs/Ru

0.4 1.0 1.9 2.8 3.6 4.4 5.1

35.7 18.4 66.4 29.0 61.0 139 91.7

4.44 6.38 8.20 8.19 11.1 8.99 9.32

1.75 1.97 3.02 2.85 3.34 3.47 3.33

2.91 > 2.91 > 2.91 87.5 > 2.91 114 39.0

> > > > > >

68.8 68.8 68.8 68.8 68.8 68.8 68.8

697 157 > 157 26.6 117 54.3 54.6

> 37.9 37.9 > 37.9 > 37.9 50.8 22.5 38.5

[CalixBisMePhC](o) = 5.4·10−3 mol·kg−1, [Metal](a) = 5.0·10−4 mol·kg−1, T = 298.1 K; contact time = 200 min; phase ratio =1:1; shake speed = 120 rpm. The expanded uncertainty of the result with level of confidence is 0.95. a

It could be concluded that CalixBisMePhC/CHCl3 had an excellent extraction ability and high selectivity toward Cs(I) over all other tested elements in 4.4 mol·kg−1 HNO3. 3.2. Dependence of Extraction of the Tested Metals on HNO3 Concentration. CalixBisMePhC consists of one calix[4]arene and two crown ether frames. Oxygen atoms in the crown ether moieties usually have a strong affinity toward nitric acid through hydrogen bonds. Acidity of HLW generated from the PUREX process was known to be about 2.8 mol·kg−1 HNO3. The acidity may have a certain effect on the recognition of CalixBisMePhC toward metal ions. In other words, chemical complexation between metal ions and CalixBisMePhC and the protonation of CalixBisMePhC with HNO3 would be two competing reactions. Table 2 shows the dependence of extraction of CalixBisMePhC/CHCl3 toward Na(I), K(I), Rb(I), Cs(I), Sr(II), Ba(II), Pd(II), and Ru(III) on HNO3 concentration in the range of 0.4 mol·kg−1 to 5.1 mol·kg−1. As can been seen, at the evaluated HNO3 concentration, distribution ratios of Na(I), Sr(II), Ba(II), Pd(II), and Ru(III) were less than 0.10, showing weak or no extraction. Distribution ratios of K(I) and Rb(I)

distribution ratios less than 0.10 showed almost no extraction with CalixBisMePhC/CHCl3. It was attributed to the weak complexation of CalixBisMePhC toward these metals. Although K(I) and Rb(I) were slightly extracted with the maximum distribution ratios of 0.17 for K(I) and 0.62 for Rb(I), respectively, their influences on the selective extraction toward Cs(I) were not notable. The obvious high selectivity for Cs(I) was based on matching size between Cs(I) and CalixBisMePhC. Since Sr(II) and Ba(II) were not extracted in the tested condition, it was inferred that all alkaline earth metals might not be unfavorable to the extraction of CalixBisMePhC/CHCl3. This is helpful for recognition and extraction of Cs(I) from HLW with the CalixBisMePhC/CHCl3 system. Table 1 also showed the dependence of Cs(I) recovery percentage on contact time. As can be seen, being similar with DCs, the extraction percentage of Cs(I) increased with the increase in contact time and remained at about 64 % after 15 min. C

dx.doi.org/10.1021/je400735z | J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

were, respectively, less than 0.20 and 0.60, showing weak extraction. The values of DCs were obviously greater than distribution ratios of the metals mentioned above. It reflected that these coexistent metals might have no adverse impact on solvent extraction and separation of Cs(I). It was found that DCs increased obviously from 0.4 mol·kg−1 to 4.4 mol·kg−1 HNO3 and then decreased from 4.4 mol·kg−1 to 5.1 mol·kg−1 HNO3. The curve of Cs(I) recovery percentage is similar to that of DCs. The optimum acidity for the Cs(I) extraction in the experimental conditions was therefore determined to be 4.4 mol·kg−1 HNO3. Separation factors (SF) of Cs(I) to the other metals at different HNO3 concentrations were calculated and presented in Table 3. It was reported that conventional complexing agents were difficult to complex Cs(I).15 The present research reflected that in 4.4 mol·kg−1 HNO3 solution, CalixBisMePhC was capable of recognizing and complexing Cs(I). The selectivity of CalixBisMePhC for Cs(I) was identified as matching size between the crown cavity moiety and Cs(I) as well as πbonding interaction with the arene groups and structural reorganization of the molecule of CalixBisMePhC. 3.3. Extraction Mechanism of CalixBisMePhC for Cs(I). CalixBisMePhC is a kind of neutral extractant. If one assumes that the composition of the extracted species of Cs(I) with CalixBisMePhC is mCsNO3·nCalixBisMePhC·xHNO3, then, the extraction equilibrium of Cs(I) would be expressed as the following:

Figure 3. Relationship between log DCs and log[CalixBisMePhC](o).

mCs+(a) + m NO3−(a) + x H+(a) + x NO3−(a) + nCalixBisMePhC(o) ⇌ mCsNO3 ·nCalixBisMePhC·x HNO3(o)

(3)

The apparent extraction constant (Kex) is expressed as

Figure 4. Relationship between log DCs and log[Cs+](a).

Kex = [mCsNO3 ·nCalixBisMePhC·x HNO3](o)

in metal under the experimental conditions, that is m = 1. The effect of the concentration of H+(a) on the distribution ratio of Cs(I) was investigated at fixed other experimental conditions. A plot of log DCs versus log[H+](a) is a straight line with a linear slope close to 0.5, that is x = 1, as shown in Figure 5. It could be determined that a 1:0.5:0.5 type of the extracted species between Cs(I), CalixBisMePhC, and HNO3 was formed. The

/([Cs+]m(a) [H+]x (a)[NO− 3](m + x) (a)[CalixBisMePhC]n (o))

(4)

The DCs is in accordance with the following equation: DCs = m[mCsNO3 ·nCalixBisMePhC·x HNO3](o) /[Cs+](a) (5)

According to eqs 4 and 5, the relationship between the distribution ratio and the extraction constant is described as log DCs = (m − 1) log[Cs+](a) + (m + x) log[NO3−](a) + x log[H+](a) + n log[CalixBisMePhC](o) + log mKex

(6)

Under the conditions of fixed concentration of Cs(I) and other experimental parameters, varying the concentration of CalixBisMePhC in the organic phase yielded the corresponding DCs. A plot of log DCs versus log[CalixBisMePhC](o) yielded a straight line with a linear slope close to 0.5, that is n = 0.5, as shown in Figure 3. Similarly, the effect of Cs(I) concentration on D Cs was examined at constant concentration of CalixBisMePhC and others, and a plot of log DCs versus log[Cs+](a) is shown in Figure 4. The slope of the resultant straight line is close to 0. It showed that the extracted species of Cs(I) and CalixBisMePhC was a kind of mononuclear complex

Figure 5. Relationship between log DCs and log[H+](a). D

dx.doi.org/10.1021/je400735z | J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Figure 6. Extraction mechanism of Cs(I) with CalixBisMePhC/CHCl3.

reaction. It can be further found that there is a linear relationship between distribution ratio and temperature. The recovery percentage of Cs(I) also decreased as temperature increased. Therefore, increasing the operation temperature in experimental conditions was adverse to the Cs(I) extraction. This adverse effect might be related to the change in the structure of CalixBisMePhC. Therefore, the temperature of 298 K was favorable to the extraction of Cs(I) with CalixBisMePhC/CHCl3. A plot of ln DCs vs 1/T was constructed and shown in Figure 8.

composition of the complex of Cs(I) with CalixBisMePhC was therefore determined to be CsNO3·0.5CalixBisMePhC· 0.5HNO3. On the basis of the above discussion, the extraction reaction of Cs(I) with CalixBisMePhC in HNO3/CHCl3 medium is represented as the following: Cs+(a) + NO3−(a) + 0.5H+(a) + 0.5NO3−(a) + 0.5CalixBisMePhC(o) ⇌ CsNO3 ·0.5CalixBisMePhC·0.5HNO3(o)

(7)

The extraction constant of Cs(I) was calculated and the value of log Kex was 1.087. Equation 7 reflected that in HNO3 solution, the extraction of Cs(I) resulted from the effective recognition and complexation of CalixBisMePhC for Cs(I). The extraction mechanism of Cs(I) with CalixBisMePhC is illustrated in Figure 6. 3.4. Dependence of Extraction of Cs(I) on Temperature. The extraction of Cs(I) with CalixBisMePhC in 4.4 mol· kg−1 HNO3 was studied at different temperatures from 298 K to 318 K. The results are shown in Figure 7. As can be seen, increasing the extraction temperature from 298 K to 318 K resulted in an obvious decrease in DCs from 1.80 to 0.61, denoting an exothermic nature in the extraction

Figure 8. Relationship between ln DCs and 1/T.

Thermodynamic parameters of the extraction reaction including the change in enthalpy (ΔH°), free energy (ΔG°), and entropy (ΔS°) shown in Table 4 were calculated using the Table 4. Thermodynamic Parameters of the Solvent Extraction of Cs(I) with CalixBisMePhC/CHCl3a ΔS° −1

ΔG° (KJ·mol−1)

ΔH° −1

−1

(J·mol ·K )

(KJ·mol )

298 K

303 K

308 K

313 K

318 K

−143.82

−44.53

−1.67

−0.95

−0.23

0.49

1.21

[HNO3](a) = 4.4 mol·kg−1, [Cs(I)](a) = 5.0·10−4 mol·kg−1; contact time = 200 min; phase ratio = 1:1. The expanded uncertainty of the result with level of confidence is 0.95. a

Figure 7. Dependence of extraction of Cs(I) on temperature. [HNO3] = 4.4 mol·kg−1, [Cs(I)] = 5.0·10−4 mol·kg−1, contact time = 60 min, phase ratio = 1:1, shake speed = 120 rpm. E

dx.doi.org/10.1021/je400735z | J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Figure 9. A schematic diagram for the extraction of cesium from a simulated HLW.

Arrhenius law d ln D/dT = ΔH°/(RT2) and Gibbs free energy ΔG° = ΔH° − TΔS°. The enthalpy (ΔH°) is negative which means an exothermic extraction process occurs. The entropy (ΔS°) is negative corresponding to a decrease in degree of freedom of the system as the Cs(I) ions are restricted by complex formation. The free energy (ΔG°) is negative confirming the spontaneous nature of this extraction. 3.5. Development of an Extraction Flowsheet for Cesium Separation. On the basis of the fundamental research in the experimental conditions, a schematic diagram of countercurrent extraction for the separation of cesium from a simulated acidic HLW is presented in Figure 9. It consists of a five-stage extraction, a two-stage scrub, and a five-stage strip. The organic phase used for extraction is CHCl3 with 5.4·10−3 mol·kg−1 of CalixBisMePhC. HNO3 with a concentration of 4.4 mol·kg−1 is selected as the acidity of feed solution for the Cs(I) extraction.

91126021, Zhejiang Provincial Natural Science Foundation of China under Contract No. Y4110002, and Zhejiang Provincial Commonweal Technology Applied Research Project under Contract No. 2012C23032. Notes

The authors declare no competing financial interest.



4. CONCLUSION Cs(I) in HLW is one of the most important species because of its high heat emitting and strong radioactivity. The separation of Cs(I) is valuable to the safe treatment and disposal of HLW. However, separation of Cs(I) from HLW has been one of the most challenging works. For this purpose, a new calixbiscrown derivative CalixBisMePhC was prepared. The extraction of Na(I), K(I), Rb(I), Cs(I), Sr(II), Ba(II), Pd(II), and Ru(III) from an aqueous solution with CalixBisMePhC/CHCl3 was studied. It was demonstrated that Cs(I) could be selectively extracted from the aqueous solution with CalixBisMePhC/CHCl3. The DCs increased with increasing HNO3 concentration and had a maximum at around 4.4 mol·kg−1 HNO3. The effect of temperature was studied and the results proved that the extraction reaction was exothermic with the standard enthalpy change ΔH° = −44.53 kJ·mol−1. The composition of the extracted species of Cs(I) with CalixBisMePhC in HNO3 medium was determined as CsNO3· 0.5CalixBisMePhC·0.5HNO3. Some valuable parameters for the extraction of Cs(I) in chemical engineering were obtained that are beneficial to the partitioning of Cs(I) with the CalixBisMePhC-containing extraction system.



REFERENCES

(1) Yusuf, A. Z. Computational screening of ionic liquids as solvents for uranium extraction from spent nuclear fuel. Dissertation, Ahmadu Bello University, Nigeria, 2012. (2) Bascetin, E.; Atun, G. Adsorptive removal of strontium by binary mineral mixtures of montmorillonite and zeolite. J. Chem. Eng. Data 2009, 55, 783−788. (3) Denecke, M. A.; Panak, P. J.; Burdet, F.; Weigl, M.; Geist, A.; Klenze, R.; Mazzanti, M.; Gompper, K. A comparative spectroscopic study of U(III)/Am(III) and Ln(III) complexed with N-donor ligands. C.R. Chim. 2007, 10, 872−882. (4) Serrano-Purroy, D.; Baron, P.; Christiansen, B.; Malmbeck, R.; Sorel, C.; Glatz, J. P. Recovery of minor actinides from HLLW using the DIAMEX process. Radiochim. Acta 2005, 93, 351−355. (5) Ozawa, M.; Suzuki, T.; Koyama, S. I.; Akatsuka, H.; Mimura, H.; Fujii, Y. A new back-end cycle strategy for enhancing separation, transmutation and utilization of materials (Adv. ORIENT cycle). Prog. Nucl. Energ. 2008, 50, 476−482. (6) Wakabayashi, T. Transmutation characteristics of MA and LLFP in a fast reactor. Prog. Nucl. Energ. 2002, 40, 457−463. (7) Tian, M.; Jia, Q.; Song, N.; Quan, X.; Liu, C. Extraction and separation of rare earth elements from nitrate medium with mixtures of sec-octylphenoxyacetic acid and 2,2′-bi- pyridyl. J. Chem. Eng. Data 2010, 55, 4281−4284. (8) Takano, H.; Nishihara, K.; Tsujimoto, K.; Sasa, T.; Oigawa, H.; Takizuka, T. Transmutation of long-lived radioactive waste based on double-strata concept. Prog. Nucl. Energ. 2000, 37, 371−376. (9) Can, M.; Bulut, E.; Ö zacar, M. Synthesis and characterization of pyrogallol-formaldehyde nano resin and its usage as an adsorbent. J. Chem. Eng. Data 2012, 57, 2710−2717. (10) Girnt, D.; Roesky, P. W.; Geist, A.; Ruff, C. M.; Panak, P. J.; Denecke, M. A. 6-(3,5-di-methyl-1-H-pyrazol-1-yl)-2,2′-bipyridine as ligand for actinide(III)/lanthanide(III) separation. Inorg. chem. 2010, 49, 9627−9635. (11) Nawada, H. P.; Fukuda, K. Role of pyro-chemical processes in advanced fuel cycles. J. Phys. Chem. Solids 2005, 66, 647−651. (12) Banerjee, D.; Rao, M. A.; Wattal, P. Separation and recovery of Cs from high active waste simulant using resorcinol formaldehyde polycondensate resin: Batch and column studies. Sep. Sci. Technol. 2013, 48, 133−139. (13) Faghihian, H.; Moayed, M.; Firooz, A.; Iravani, M. Synthesis of a novel magnetic zeolite nanocomposite for removal of Cs+ and Sr2+ from aqueous solution: Kinetic, equilibrium and thermodynamic studies. J. Colloid Interface Sci. 2012, 393, 445−451.

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Tel./Fax: +86 571 87953919. Funding

The present work was financially supported by the National Natural Science Foundation of China under Contract No. F

dx.doi.org/10.1021/je400735z | J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

(14) Raut, D.; Mohapatra, P.; Choudhary, M.; Nayak, S. Evaluation of two calix-crown-6 ligands for the recovery of radio cesium from nuclear waste solutions: Solvent extraction and liquid membrane studies. J. Membr. Sci. 2012, 429, 197−205. (15) Jain, P.; Sahariya, J.; Ahuja, B. L. Compton scattering study and electronic structure of different phases of NH4NO3. Phys. Scr. 2013, 87, 65−102. (16) Boden, S.; Rogiers, B.; Jacques, D. Determination of Cs-137 contamination depth distribution in building structures using geostatistical modelling of ISOCS measurements. Appl. Radiat. Isot. 2013, 79, 25−36. (17) Dash, A.; Ram, R.; Pamale, Y. A.; Deodhar, A. S.; Venkatesh, M. Recovery of 137Cs from laboratory waste using solvent extraction with sodium tetraphenylboron (TPB). Sep. Sci. Technol. 2012, 47, 81−88. (18) Ojovan, M. I; Lee, W. E. Glassy wasteforms for nuclear waste immobilization. Metall. Mater. Trans. A 2011, 42, 837−851. (19) Kandwal, P.; Mohapatra, P. K.; Ansari, S. A.; Manchanda, V. K. Selective caesium transport using hollow fibre-supported liquid membrane containing calix[4]arene-bis-naphthocrown-6 as the carrier extractant. Radiochim. Acta 2010, 98, 493−498. (20) Murali, M. S.; Raut, D. R.; Prabhu, D. R.; Mohapatra, P. K.; Tomar, B. S.; Manchanda, V. K. Removal of Cs from simulated highlevel waste solutions by extraction using chlorinated cobaltdicarbollide in a mixture of nitrobenzene and xylene. J. Radioanal. Nucl. Chem. 2012, 291, 611−616. (21) Xiao, C.; Zhang, A.; Chai, Z. Synthesis and characterization of a new polymer-based supramolecular recognition material and its adsorption for cesium. Solvent Extr. Ion Exch. 2012, 30, 17−32. (22) Böhmer, V. Calixarenes, macrocycles with (almost) unlimited possibilities. Angew. Chem., Int. Ed. 1995, 34, 713−745. (23) Gutsche, C. D.; Lin, L. G. Calixarenes 12: The synthesis of functionalized calixarenes. Tetrahedron 1986, 42, 1633−1640. (24) Casnati, A.; Giunta, F.; Sansone, F.; Ungaro, R.; Montalti, M.; Prodi, L.; Zaccheroni, N. Synthesis, complexation and photophysics of 1,3-alternate calix[4]arene-crowns-6 bearing fluorophoric units on the bridge. Supramol. Chem. 2001, 13, 419−434. (25) Weber, E. Polytopic cation receptors. 2. Synthesis and selective complex formation of spiro-linked “multiloop crown compounds”. J. Org. Chem. 1982, 47, 3478−3486. (26) Pellet-Rostaing, S.; Chitry, F.; Spitz, J. A.; Sorin, A.; FavreRéguillon, A.; Lemaire, M. New water-soluble calix[4]arene-bis(benzocrown-6) for caesium-sodium separation by nanofiltrationcomplexation. Tetrahedron 2003, 59, 10313−10324.

G

dx.doi.org/10.1021/je400735z | J. Chem. Eng. Data XXXX, XXX, XXX−XXX