Sie sind auf Seite 1von 6

Hydrometallurgy 147148 (2014) 164169

Contents lists available at ScienceDirect

Hydrometallurgy
journal homepage: www.elsevier.com/locate/hydromet

Extraction mechanism, behavior and stripping of Pd(II) by


pyridinium-based ionic liquid from hydrochloric acid medium
Yu Tong, Chen Wang, Jing Li, Yanzhao Yang
Key Laboratory for Special Functional Aggregated Materials of Education Ministry, School of Chemistry and Chemical Engineering, Shandong University, Jinan 250100, China

a r t i c l e

i n f o

Article history:
Received 18 October 2013
Received in revised form 14 May 2014
Accepted 19 May 2014
Available online 6 June 2014
Keywords:
Palladium
Extraction
Hexadecylpyridinium chloride
Stripping

a b s t r a c t
The hexadecylpyridinium chloride/chloroform system was employed for the Pd(II) extraction from hydrochloric
acid medium. The effects of extraction time and material concentrations were examined for Pd(II) extraction. The
anion-exchange mechanism of Pd(II) extraction by [Hpy]Cl was conrmed by Job's method, UVVis, infrared
spectrum and 1H NMR analysis. Thermodynamic parameters of the Pd(II) extraction reaction were obtained
from the thermodynamics analysis, which illustrates that higher temperatures show a negative effect on the
Pd(II) extraction. When the mole ratio of [Hpy]Cl to Pd(II) is above 4, the Pd(II) extraction yield (E%) is almost
100%. It also has high selectivity over some metals (Cu(II), Co(II), Ni(II), Fe(III), Al(III) and Sn(IV)). Moreover, oxalate is explored as a reducing agent for [Hpy]+2[PdCl4]2 complexes, by which Pd(II) is stripped in the form of
palladium powders and [Hpy]Cl also can be recycled. Therefore, the method is an efcient, effective and highly
selective approach to extract Pd(II) and recover metal palladium.
2014 Elsevier B.V. All rights reserved.

1. Introduction
Palladium (Pd) is an important noble metal which is widely used in
catalysts, aviation, pharmaceutical, electronic industries, weapons and
so on. Due to its signicance in modern life and the growing demand,
palladium is not only rened from mineral sources but also recovered
and pre-concentrated from secondary sources (e.g., exhausted catalysts
and electronic components). The recovery of Pd from waste solutions or
pre-concentration techniques at low levels has an important value in
economy and environment. For the recovery or rening of palladium,
solvent extraction is a suitable method with obvious advantages, related
to high selectivity, metal purity and energy consumption (Bernardis
et al., 2005). Recently, Pd extractions using ionic liquids as extractants
or media have become a focus point.
Ionic liquids (ILs) are a type of organic salts solely composed of ions
and with the melting points near or below room temperature. The
common cations of ionic liquids include ammonium, phosphonium,
imidazolium and pyridinium cations (Freemantle, 1998). ILs have
some unique characteristics, such as high thermal stability, negligible
volatility and nonammability, so ionic liquids have been extensively
employed as solvent media in recovery or separation of various
metal ions (Coll et al., 2012; Goyal et al., 2011; Jensen et al., 2003; Li
et al., 2007; Regel-Rosocka and Wisniewski, 2011; Sun et al., 2012;

Corresponding author. Tel.: +86 531 88365431; fax: +86 531 88564464.
E-mail address: yzhyang@sdu.edu.cn (Y. Yang).

http://dx.doi.org/10.1016/j.hydromet.2014.05.016
0304-386X/ 2014 Elsevier B.V. All rights reserved.

Whitehead et al., 2007). Over the years, some work about Pd extractions
using ionic liquids has been carried out. The application of commercial
Aliquat 336 ionic liquids as the extractant to extract palladium from nitric acid has been attempted (Giridhar et al., 2006). Phosphonium ionic
liquids, CyphosIL 101 and CyphosIL 104 have been also employed to extract palladium(II) from hydrochloric acid solutions (Cieszynska and
Wisniewski, 2010, 2011, 2012). Pd(II) sorption from HCl solutions has
been tested by using an ionic liquid CyphosIL-101, which is impregnated on Amberlite XAD-7 resin (Navarro et al., 2012). The analytical potential of [Hmim][BF4] have been performed for the determination of
Pd in food additive, sea water, tea and biological samples (Vaezzadeh
et al., 2010). In previous studies, phosphonium or ammonium ionic liquids have been used as extractants or solvents for Pd(II) extraction, but
pyridinium-based ionic liquids are seldom mentioned in the Pd(II) extraction research and little attention is paid to the mechanism of
Pd(II) extraction. Meanwhile, the relevant thermodynamics research
and the Pd stripping are also rarely reported.
Hexadecylpyridinium chloride ([Hpy]Cl) is an analogue of
pyridinium-based ionic liquids, by which we can explore the extraction
mechanism of Pd by pyridinium-based ionic liquids. In the present
work, [Hpy]Cl is exploited as the extractant of Pd(II). The amphiphilic
structure of [Hpy]Cl can signicantly reduce the interfacial tension between aqueous phase and organic phase, which effectively accelerates
the extraction. We focus on the extraction mechanism of Pd(II) by
[Hpy]Cl from hydrochloric acid medium and the relevant thermodynamic parameters. What is more, ethanedioic acid is explored as the reducing agent of Pd[Hpy]Cl complexes, by which Pd can be stripped in
the form of palladium powders and [Hpy]Cl is recycled.

Y. Tong et al. / Hydrometallurgy 147148 (2014) 164169

165

2. Experimental

2.4. Pd(II) stripping and IL regeneration

2.1. Reagents and materials

After the extraction, the Pd(II)-loaded organic phase was separated


from the aqueous phase and then mixed with an equal volume of
H2C2O4 solution (1.00 mol L1). By reux condensation, the reduction
reaction was maintained at 338 1 K, for 4 h. In this process, black palladium powders were formed. After the reduction reaction, Pd precipitation was ltrated from liquid phases and the two liquid phases were
separated by centrifugation. The chloroform phase containing [Hpy]Cl
was recovered and can be reused to extract Pd(II) from new feed solutions again.

Chloroform (AR) was purchased from Kermel Chemical Reagent


Tianjin Co., Ltd. (Tianjin, China) and used without any further purication. [Hpy]Cl, hexadecylpyridinium chloride (N99%), was procured
from Lanzhou Greenchem ILS, LICP. CAS. (Lanzhou, China) and used as
received. The feed solutions were prepared by dissolving metal chlorides in hydrochloric acid solutions: PdCl2, Guangfu Institute of Fine
Chemical (Tianjin, China); SnCl45H2O, FeCl36H2O, CoCl26H2O,
NiCl26H2O, CuCl22H2O and AlCl36H2O, Kermel Chemical Reagent
Tianjin Co., Ltd. (Tianjin, China). Distilled water was used to prepare
the aqueous solutions in all experiments. All the other chemicals used
in this study were of analytical or reagent grade.
2.2. Analytical techniques
Before and after Pd(II) extraction, the Pd(II) concentration in aqueous phase was determined with an atomic absorption spectrophotometer (3150, Precision & Scientic Instrument Shanghai Co., Ltd., Shanghai,
China). Other metal ions (Cu(II), Co(II), Ni(II), Fe(III), Al(III) and Sn(IV))
concentrations in solutions were determined with an inductively
coupled plasma atomic emission spectrometer (IRIS Intrepid II XSP,
Thermo Electron Corporation, Boston, US). Then the concentrations of
metals in organic phase were calculated based on mass balances. By
using UV spectrophotometry (UV-9000, Shanghai Metash Instruments
Co., Ltd., Shanghai, China), the concentration of [Hpy]Cl in aqueous
phase was determined by the discoloration method with bromocresol
green, at pH = 5.8, = 614 nm (Tong et al., 2014), and the concentration of Cl in the aqueous phase was determined by the spectrophotometry method with mercury thiocyanate and ammonium
ferric sulfate at 460 nm. After the Pd(II) extraction, the Pd-loaded organic phase was separated from the aqueous solution, evaporated
completely and washed with distilled water for several times at 320
1 K, 2.5 104 Pa. The obtained sample ([Hpy]ClPd complexes) was analyzed by FT-IR (Tensor27, Bruker Corporation, Karlsruhe, Germany)
and 1H NMR (AV300, Bruker Corporation, Karlsruhe, Germany).
2.3. Extraction of metal ions
For the metal extractions, a required amount of [Hpy]Cl was added
into chloroform to form the organic phase. The organic phase and the
aqueous solution containing Pd(II) or mixed metal ions (Cu(II), Co(II),
Ni(II), Fe(III), Al(III) and Sn(IV)) were added to a glass tube and then
equilibrated mechanically in an orbital shaker for 10 min. The volume
ratio of the aqueous phase to the organic phase was Rw:o = 5. Both of
organic phase and aqueous phase had been pre-saturated. Afterwards,
the organic phase and the aqueous phase were separated by the use of
a centrifuge at 2000 rpm for 2.5 min. The separated organic phase and
aqueous phase were both clear and transparent. Unless stated specially,
all experiments were carried out at 298 1 K. The extraction efciency
(E%) and distribution coefcient (D) of metal ions are expressed as the
following equations:
E%

3. Results and discussion


3.1. Effect of extraction time
Extraction time was preliminarily studied. With the use of the organic phase of 1.88 mmol L1 [Hpy]Cl and the aqueous phase of
10.0 mg L1 Pd(II) in 0.100 mol L1 HCl solution, Rw:o = 5, the effect
of vibration time on Pd(II) extraction by the [Hpy]Cl/chloroform system
was studied. The vibration time varied from 1 to 40 min. The Pd(II) extraction yield (E%) reaches 98.8% within 5 min, then the extraction is independent of the vibration time. Therefore, the vibration time of 10 min
is sufcient in the following extraction experiments. It should be noted
that [Hpy]Cl is a signicant surfactant to reduce the interfacial tension,
which can accelerate the extraction.
3.2. Effect of [Hpy]Cl concentration
The inuence of the initial [Hpy]Cl concentration on Pd(II) distribution ratio (D) and extraction yield (E) was studied. Meanwhile, a classical ionic liquid, 1-hexadecyl-3-methylimidazolium chloride ([C16mim]
Cl) was also tested as a comparison. The results are shown in Fig. 1. Initially, the Pd(II) extraction yield, E increases with [Hpy]Cl concentration.
When the concentration reaches approximately 1.88 mmol L 1 (the
mole ratio of [Hpy]Cl to Pd(II) is about 4), the extraction yield is almost
100%. The results also show that the pyridinium-based ionic liquid
is much better than the imidazolium-based ionic liquid for Pd(II)
extraction.
3.3. Effect of HCl concentration
The extraction behaviors of Pd(II) by the [Hpy]Cl/chloroform system
with various concentrations of hydrochloric acid in the aqueous phase

Cinitial; metal Caq; metal


 100%
Cinitial; metal

Cinitial; metal Caq; metal


Cor; metal
Rw:o 
Caq; metal
C aq; metal

The symbols Cinitial, metal and Caq, metal represent the initial concentration of metal ion and the instantaneous concentration in the aqueous
phase, respectively.

Fig. 1. The extraction yield (E) and distribution ratio (D) of Pd(II) as a function of
[Hpy]Cl or [C16mim]Cl concentration. Aqueous phase, 10.0 mg L1 Pd(II) in 0.100 mol L1
HCl; Rw:o = 5.

166

Y. Tong et al. / Hydrometallurgy 147148 (2014) 164169

were studied. The [Hpy]Cl concentration in the organic phase


was 1.88 mmol L1, while the HCl concentration varied from
0.100 mol L1 to 1.00 mol L1. The results are shown in Fig. 2. The increasing HCl concentration shows a negative effect on the Pd(II) extraction performance. The decrease in Pd(II) extraction is caused by the
competition between PdCl24 and Cl to associate with [Hpy]+ by
anion-exchange reaction expressed by Eq. (1) in Section 3.6. This behavior has been shown in similar Pdextractant systems (Cieszynska and
Wisniewski, 2010, 2012; Navarro et al., 2012).
3.4. Determination of the coordination number
Job's method, also called the method of continuous variation, is
widely used for the investigation and analysis of complexes' formation,
which involves measure of complexes' formation at various mole ratios
of the reactants while maintaining constant the total amount of reactants (Huang et al., 2003; Likussar, 1973; Musier and Hammes, 1988).
A modied method of continuous variation was employed to
investigate the complexation of the extracted Pd(II) with the [Hpy]Cl/
chloroform system. The total amount of [Hpy]Cl and Pd(II) was maintained at 1.68 106 mol, Rw:o = 5, and the initial Pd(II) concentration
in the aqueous phase, Caq,Pd ranged from 0.056 to 0.223 mmol L1.
The results are shown in Fig. 3 by plotting the molar ratios of Pd(II) to
[Hpy]Cl as a function of Pd(II) equilibrium concentrations in the organic
phases (Cor,Pd).
As shown in Fig. 3(b), in the system of 0.100 mol L1 HCl, the maximum equilibrium concentration of Pd(II) in the organic phase is at the
point where the mole ratio of Pd(II) to [Hpy]Cl is 0.5 and the largest
amount of Pd(II)[Hpy]Cl complexes is formed. The data indicate
that Pd(II)[Hpy]Cl complexes with the mole ratio of 1:2 are formed.
As shown in Fig. 3(a), the same results also occur in systems of
0.200 mol L1 HCl and 0.400 mol L1 HCl, which indicates that the coordination number is independent on the HCl concentration.
3.5. The spectrum analysis
We employed a UVVis spectrometer to analyze the chloroform solution of [Hpy]Cl and the [Hpy]ClPd complexes with chloroform as the
blank sample. The Pd(II)-loaded organic phase was also analyzed, while
the same organic phase containing [Hpy]Cl but without Pd(II) was used
as the blank sample. As shown in Fig. 4, the complexes in chloroform
show two ultraviolet absorption peaks (max = 252, 292 nm). The absorption peak at ~ 252 nm belongs to [Hpy] due to the ultraviolet absorption of [Hpy]Cl in chloroform (the curve a in Fig. 4, max =
259 nm). The curve b in Fig. 4 indicates that the UV absorption of the
Pd(II)-loaded organic phase has a peak at 292 nm, which explains that

Fig. 2. The extraction yield (E) and distribution ratio (D) of Pd(II) as the function of HCl
concentration. Aqueous phase, 10.0 mg L1 Pd(II); Rw:o = 5.

Fig. 3. Job's plot for the Pd(II)[Hpy]Cl system, the molar ratio of Pd(II) to [Hpy]Cl ranging
from 0.2 to 2, HCl concentrations in aqueous phases are 0.100, 0.200 and 0.400 mol L1
respectively.

the [Hpy]ClPd complexes has the same absorption peak at 292 nm


due to the Pd(II) loading. The results indicate that hexadecylpyridinium
ligands and central palladium atom contribute to the doublet UV absorption peaks of the [Hpy]ClPd complexes.
To obtain direct evidence of the interaction between [Hpy]Cl and
PdCl24 , IR was employed to analyze the [Hpy]ClPd complexes. The

Fig. 4. UVVis spectra of Pd(II) complexes. (a) 3.76 10-4 mol L1 [Hpy]Cl in chloroform; (b) 1.88 104 mol L1 [Hpy]ClPd complexes in the organic phase; (c) 1.88
104 mol L1 [Hpy]ClPd complexes in chloroform.

Y. Tong et al. / Hydrometallurgy 147148 (2014) 164169

sample preparation was followed by Section 2.2. The IR spectra of [Hpy]


Cl and the [Hpy]ClPd complexes (the molar ratio of Pd(II) to [Hpy]Cl is
1:2) are shown in Fig. 5.
The vibration bands are shown in the IR spectra of [Hpy]Cl: the ring
C\H deformation in plane, 1179; the ring C\N stretch, 1472; the ring
C_C stretch, 1638; aliphatic C\H stretch, 2850; C\H stretch of CH3,
2915; the ring C\H stretch, 3049; the O\H stretch and N\H stretch,
3411. With Pd(II) loaded to the organic phase, the relative strength of
the stretch band at 3411 cm1 has been weakened because of poor
hydrophilicity of the [Hpy]ClPd complexes compared with [Hpy]Cl.
The aliphatic C\H stretch peaks (2850 cm1) have no shift because of
the weaker interaction between PdCl24 and the alkyl side chains of
[Hpy]Cl. However, the ring C\H stretch has clearly been enhanced
and all the absorption bands related to pyridine ring have shifted:
3049 3044; 1638 1634; 1472 1470; 1179 1178. The results
show that the aliphatic C\H stretch is slightly inuenced by the
2
loading of PdCl2
4 , but there is a stronger interaction between PdCl4
and [Hpy]Cl cationic headgroup.
The 1H NMR measurement was also employed to analyze the [Hpy]
ClPd complexes while DMSO was used as the solvent. The 1H NMR
spectra of [Hpy]Cl and the [Hpy]ClPd complexes are shown in Fig. 6.
With the combination of [Hpy]Cl and Pd, the chemical shifts of hydrogen atoms nearest to the nitrogen atom (A and D) change largely:
AA, 9.138 9.116; DD, 4.608 4.598. The hydrogen atoms further away of the nitrogen atom are slightly inuenced by the loading
of PdCl2
4 : C\C, 8.168 8.166; EE, 1.907 1.909; BB, 8.612
8.610. And then the furthest hydrogen atoms are nearly unchanged:
FF, 1.233; GG, 0.854. The results imply a stronger interaction of
PdCl2
with the nitrogen atom on [Hpy]Cl cationic headgroup than
4
that with the alkyl side chains of [Hpy]Cl.

3.6. Extraction mechanism analysis


According to Job's plots and the spectra of the [Hpy]Pd complexes,
1:2 associates are formed between Pd(II) and [Hpy]Cl, meanwhile, a
stronger interaction exists between PdCl24 and [Hpy]. Furthermore,
[Hpy]Cl is composed of [Hpy]+ cation and Cl anion, so the anionexchange mechanism was supposed for Pd(II) extraction. Pd(II) extraction by [Hpy]Cl is deduced as the formation of the neutral complexes
as [Hpy]+2[PdCl4]2. The extraction reaction of Pd(II) by the [Hpy]Cl/
chloroform system from the hydrochloric acid medium can be presented as the following equations:

PdCl4

2
aq

2HpyClor Hpy2 PdCl4 or 2Cl

aq

Fig. 5. Infrared spectra of [Hpy]Cl and Pd(II)[Hpy]Cl complexes.

167

Fig. 6. 1H NMR spectra of [Hpy]Cl and Pd(II)[Hpy]Cl complexes.

so the distribution ratio (D) of Pd(II) can be written as:



Hpy2 PdCl4  or
Pdor

:
Pdaq
PdCl4 aq

To further conrm the validity of the anion-exchange equation, the


effect of Pd(II) extraction on Cl concentration in the aqueous phase
was studied. The Cl concentrations were adjusted by NaCl, and the organic phase is of 1.88 mmol L1 [Hpy]Cl and the aqueous phase is of
10.0 mg L1 Pd(II), Rw:o = 5.
As shown in Fig. 7, the degree of Pd(II) extraction decreases with the
increasing Cl concentration because of the equilibrium shift of Eq. (1).
The data conrm that the extraction of Pd(II) with [Hpy]Cl as the extractant follows the anion-exchange mechanism, as represented by
Eq. (1).
3.7. Thermodynamics of extraction
To investigate the effect of temperature on the extraction, equilibrium constants and thermodynamic parameters were determined at different temperatures. The extraction reaction is shown as Eq. (1), so the
equilibrium constant Ke(T) at different temperatures for the extraction

Fig. 7. The dependence of the extractability of Pd(II) on Cl concentration in aqueous


phase.

168

Y. Tong et al. / Hydrometallurgy 147148 (2014) 164169

reaction is:


2

Hpy2 PdCl4  or  Cl aq



:
Ke T 
2
2
PdCl4
 HpyClor
aq

Substituting Eq. (2) into Eq. (3):


Ke T D 

Cl aq 2
HpyClor 2

Taking logarithm of Eq. (4), The lg Ke(T) values were calculated


using the experiment data and the following equations:

lgKe T lgD 2 lgCl aq 2 lgHpyClor

lgD lgKe T2 lgCl aq 2 lgHpyClor :

The determined lg Ke(T) values are listed in Table 1. Furthermore,


Gibbs energy is G = RT ln Ke(T) and G = H TS, the
G(T) values were calculated and also listed in Table 1. The plot of the
natural logarithm of Ke(T) versus the inverse temperature will yield a
slope proportional to the enthalpy.
ln Ke T

H 1 S

:
R T
R

By plotting inverse temperature versus the natural logarithm of


Ke(T), a line (R2 = 0.98) with a slope of 1940 was obtained, shown in
Fig. 8. Then H = 16.2 kJ mol1 was determined, which indicates
that the reaction of Pd(II) extraction by the [Hpy]Cl/chloroform system
(Eq. (1)) should be exothermic. This theoretical result is consistent with
the experiments.
3.8. Pd(II) extraction from multi-metalion solutions
Natural Pd minerals are almost always related to basic igneous rocks
and closely associated with copper, nickel and iron suldes (Bernardis
et al., 2005). Moreover, this method may be used for the palladium
recovery from the electronic waste which may contain other metals,
particularly copper, cobalt, nickel, iron, aluminum. To investigate the selectivity of the [Hpy]Cl/chloroform system, an aqueous multimetal solution was prepared by adding other chloride metal salts in 0.100 mol L1
HCl solution. The concentration of each metal ion (Cu(II), Co(II), Ni(II),
Fe(III), Al(III) and Sn(IV)) in hydrochloric acid media was 10.0 mg L1.
The extraction behavior for each metal ion is plotted in Fig. 9. Pd(II) is
mainly extracted from the aqueous phase with the extraction yield of
98.8%. The [Hpy]Cl/chloroform system has very low distribution coefcients and extraction percentages for other metal ions: 0.107 and 2.10%
for Cu(II), 0.130 and 2.53% for Co(II), 0.104 and 2.05% for Ni(II), 0.171
and 3.30% for Fe(III), 0.255 and 4.85% for Al(III), 0.444 and 8.17% for
Sn(IV). High selectivity for Pd(II) extraction can be achieved by the
[Hpy]Cl/chloroform.
The results clearly show that the [Hpy]Cl/chloroform is found to be
highly selective for the extraction of Pd(II).

Fig. 8. The natural logarithm of Ke(T) versus the inverse temperature.

3.9. Recovery of Pd and [Hpy]Cl by reductive stripping


In the present study, the Pd(II)[Hpy]Cl complexes in chloroform
phase were stripped by ethanedioic acid reduction. The process was
followed by Section 2.4. The reduction reaction of the [Hpy]2[PdCl4]
complexes and ethanedioic acid may be as the following equation:
Hpy2 PdCl4  H2 C2 O4 Pd 2HpyCl 2HCl 2CO2 :

For the stripping efciency of Pd(II), the chloroform phase was


digested by nitrohydrochloric acid after the stripping process. The
Pd(II) content was determined by AAS and there is hardly any Pd(II)
left in the organic phase after the stripping process, which means that
Pd(II) in Pd(II)[Hpy]Cl complexes is completely reduced to Pd in the
form of palladium powders.
For the recovery of [Hpy]Cl, with the use of the chloroform phase
(initial concentration of [Hpy]Cl is 1.88 mmol L 1) and the aqueous
phase of 10.0 mg L1 Pd(II) in 0.100 mol L1 HCl solution, Rw:o = 5,
the effect of regeneration times on the Pd(II) extraction as well as the
amount of remaining [Hpy]Cl in the chloroform phase were investigated. The recycling rate of [Hpy]Cl after each stripping is presented by S%,
and the initial rate is 100%. The extraction yield (E), distribution ratio
(D) and S are plotted in Fig. 10.
Before the extraction, the initial mole ratio of [Hpy]Cl to Pd(II) is approximately 4:1, and 2:1 associates form between [Hpy]Cl and Pd(II).
After multiple cycles of the extraction-stripping process, the extraction
of Pd(II) is still effective, which means that [Hpy]Cl is effectively regenerated in terms of Eq. (8). The recycling rate of [Hpy]Cl maintains at
95.0% after the repeated back-extraction operation for ve times,
which explains the regeneration performance. Nevertheless, the

Table 1
The equilibrium constants and thermodynamic parameters of the Pd(II) extraction.
T,K

K(T)

G, kJ mol1

TS, kJ mol1

H, kJ mol1

293
298
303
308
313

425,000
384,000
322,000
302,000
282,000

31.6
31.9
32.0
32.3
32.7

15.4
15.7
15.8
16.2
16.5

16.2

Fig. 9. The extraction behavior of Pd(II) from multi-metal solutions. C[Hpy]Cl =


1.88 mmol L1; aqueous phase, 10.0 mg L1 metals in 0.100 mol L1 HCl; Rw:o = 5.

Y. Tong et al. / Hydrometallurgy 147148 (2014) 164169

169

References

Fig. 10. The effect of regeneration times on the Pd(II) extraction and the recycling rate of
[Hpy]Cl.

optimization conditions of the reductive stripping still need to be further explored.


4. Conclusions
The hexadecylpyridinium chloride/chloroform system was investigated for Pd(II) extraction from hydrochloric acid solutions with
[Hpy]Cl as the extractant. The extracted complex is inferred to be
[Hpy]+2[PdCl4]2, according to the mechanism analysis via continuous
variation method, UVVis, infrared spectrum and 1H NMR analysis.
Based on the study of thermodynamics, thermodynamic parameters of
the Pd(II) extraction reaction were obtained. Higher temperatures
show a negative effect on Pd(II) extraction. Under the optimum conditions, the [Hpy]Cl/chloroform system shows a high extractability for
Pd(II) and high selectivity for Pd(II) over base metals (Cu(II), Co(II),
Ni(II), Fe(III), Al(III) and Sn(IV)). Moreover, the Pd stripping and [Hpy]
Cl recovery can be achieved by the reduction reaction using ethanedioic
acid. Therefore, the extraction of Pd(II) by the [Hpy]Cl/chloroform system is considered to be an effective and highly selective approach.
Acknowledgment
This work was supported by the Natural Science Foundation of China
(Grant 21276142).

Bernardis, F.L., Grant, R.A., Sherrington, D.C., 2005. A review of methods of separation of
the platinum-group metals through their chloro-complexes. React. Funct. Polym.
65, 205217.
Cieszynska, A., Wisniewski, M., 2010. Extraction of palladium(II) from chloride solutions
with Cyphos(r)IL 101/toluene mixtures as novel extractant. Sep. Purif. Technol. 73,
202207.
Cieszynska, A., Wisniewski, M., 2011. Selective extraction of palladium(II) from hydrochloric acid solutions with phosphonium extractants. Sep. Purif. Technol. 80,
385389.
Cieszynska, A., Wisniewski, M., 2012. Extractive recovery of palladium(II) from hydrochloric acid solutions with Cyphos(r)IL 104. Hydrometallurgy 113114, 7985.
Coll, M.T., Fortuny, A., Kedari, C.S., Sastre, A.M., 2012. Studies on the extraction of Co(II)
and Ni(II) from aqueous chloride solutions using Primene JMT-Cyanex272 ionic liquid extractant. Hydrometallurgy 125126, 2428.
Freemantle, M., 1998. Designer solvents. Chem. Eng. News 76, 3237.
Giridhar, P., Venkatesan, K.A., Srinivasan, T.G., Vasudeva Rao, P.R., 2006. Extraction of ssion palladium by Aliquat 336 and electrochemical studies on direct recovery from
ionic liquid phase. Hydrometallurgy 81, 3039.
Goyal, R.K., Jayakumar, N.S., Hashim, M.A., 2011. Chromium removal by emulsion liquid
membrane using [BMIM]+[NTf2] as stabilizer and TOMAC as extractant. Desalination 278, 5056.
Huang, C.Y., Zhou, R., Yang, D.C., Chock, P.B., 2003. Application of the continuous variation
method to cooperative interactions: mechanism of Fe(II)ferrozine chelation and
conditions leading to anomalous binding ratios. Biophys. Chem. 100, 143149.
Jensen, M.P., Neuefeind, J., Beitz, J.V., Skanthakumar, S., Soderholm, L., 2003. Mechanisms
of metal ion transfer into room-temperature ionic liquids: the role of anion exchange.
J. Am. Chem. Soc. 125, 1546615473.
Li, Z., Wei, Q., Yuan, R., Zhou, X., Liu, H., Shan, H., Song, Q., 2007. A new room temperature
ionic liquid 1-butyl-3-trimethylsilylimidazolium hexauorophosphate as a solvent
for extraction and preconcentration of mercury with determination by cold vapor
atomic absorption spectrometry. Talanta 71, 6872.
Likussar, W., 1973. Computer approach to the continuous variations method for spectrophotometric determination of extraction and formation constants. Anal. Chem. 45,
19261931.
Musier, K.M., Hammes, G.G., 1988. Assessment of the number of nucleotide binding sites
on chloroplast coupling factor 1 by the continuous variation method. Biochemistry
27, 70157020.
Navarro, R., Saucedo, I., Gonzalez, C., Guibal, E., 2012. Amberlite XAD-7 impregnated with
Cyphos IL-101 (tetraalkylphosphonium ionic liquid) for Pd(II) recovery from HCl solutions. Chem. Eng. J. 185186, 226235.
Regel-Rosocka, M., Wisniewski, M., 2011. Selective removal of zinc(II) from spent pickling
solutions in the presence of iron ions with phosphonium ionic liquid Cyphos IL 101.
Hydrometallurgy 110, 8590.
Sun, X., Luo, H., Dai, S., 2012. Solvent extraction of rare-earth ions based on functionalized
ionic liquids. Talanta 90, 132137.
Tong, Y., Yang, H., Huang, Y., Yang, Y., 2014. Determination of long-chained
alkylimidazolium ionic liquids based on the hypochromic effect. Anal. Methods.
http://dx.doi.org/10.1039/c4ay00594e.
Vaezzadeh, M., Shemirani, F., Majidi, B., 2010. Microextraction technique based on ionic
liquid for preconcentration and determination of palladium in food additive, sea
water, tea and biological samples. Food Chem. Toxicol. 48, 14551460.
Whitehead, J.A., Zhang, J., Pereira, N., McCluskey, A., Lawrance, G.A., 2007. Application of
1-alkyl-3-methyl-imidazolium ionic liquids in the oxidative leaching of sulphidic
copper, gold and silver ores. Hydrometallurgy 88, 109120.

Das könnte Ihnen auch gefallen