Vous êtes sur la page 1sur 9

Electrochimica Acta 147 (2014) 8795

Contents lists available at ScienceDirect

Electrochimica Acta
journal homepage: www.elsevier.com/locate/electacta

Electrochemical behaviors of Dy(III) and its co-reduction with Al(III) in


molten LiCl-KCl salts
Ling-Ling Su a,b , Kui Liu b , Ya-Lan Liu b , Lu Wang b , Li-Yong Yuan b , Lin Wang b , Zi-Jie Li b ,
Xiu-Liang Zhao a, * , Zhi-Fang Chai b,c , Wei-Qun Shi b, *
a
b
c

School of Nuclear Science and Technology, University of South China, HengYang 421000, China
Key Laboratory of Nuclear Radiation and Nuclear Energy Technology, Institute of High Energy Physics, Chinese Academy of Sciences, Beijing 100049, China
School of Radiological & Interdisciplinary Sciences, Soochow University, Suzhou 215123, China

A R T I C L E I N F O

A B S T R A C T

Article history:
Received 15 July 2014
Received in revised form 17 September 2014
Accepted 19 September 2014
Available online 22 September 2014

In this work, the electrochemical behaviors of Dy(III) and its co-reduction with Al(III) on an inert tungsten
electrode was investigated in LiCl-KCl molten salts at the temperature of 773 K by using cyclic
voltammetry (CV), chronopotentiometry (CP) and square wave voltammetry (SWV) techniques. The
results showed that the reduction of Dy(III) ions in LiCl-KCl salts is a reversible diffusion controlled
process through a one-step reaction: Dy(III) + 3e $ Dy(0). The diffusion coefcient of Dy(III) ions was
calculated by both the CV and CP methods. Furthermore, the co-reduction of Al(III) and Dy(III) ions on the
inert tungsten electrode allows Dy(III) ions to be reduced at a more positive potential through forming AlDy alloys. The concentration ratio of Al(III) cations to Dy(III) cations has a large impact on the formation of
Al-Dy alloys. In a Dy(III) ion rich system, three signals attributed to the formation of Al-Dy intermetallic
compounds were observed in CV and SWV analyses, while only two signals corresponding to Al-Dy
intermetallic compounds were observed in the Dy(III) ion poor system. Potentiostatic and galvanostatic
electrolyses performed on an aluminum electrode identied the co-reduction by the formation of one
(Al3Dy) and two Al-Dy alloys (Al3Dy, AlDy), respectively. Finally, the electrolysis products were
characterized by the Scanning Electron Microscopy (SEM) coupled with Energy Dispersive Spectroscopy
(EDS) and X-ray diffraction (XRD) analyses.
2014 Elsevier Ltd. All rights reserved.

Keywords:
molten chlorides
dysprosium
AlCl3
intermetallic compounds
co-reduction

1. Introduction
Partitioning and Transmutation (P&T), is universally accepted to
be one of the key-steps in any future sustainable nuclear fuel
cycles, in which high efcient separations of actinides (An) and
lanthanides (Ln) are generally expected [1]. Ln could account for as
much as 25% in weight of the whole ssion products (FP) [2], and
the strong neutron absorption cross sections of Ln would largely
pull down the transmutation efciency [3]. However, the
physicochemical properties of Ln and An are very similar and
make the separation of An from Ln extremely challenging [4].
The traditional pyrochemical electrorening process, based on
chloride or uoride molten salts, has been regarded to be a
promising alternative for future spent nuclear fuel cycle [57]. The
simple inorganic molten ionic solvents are immune to radiation

* Corresponding author. Tel: +86 010 88233968 ; fax: +86 010 88235294.
E-mail addresses: shiwq@ihep.ac.cn (W.-Q. Shi), 13974753181@163.com
(X.-L. Zhao).
http://dx.doi.org/10.1016/j.electacta.2014.09.095
0013-4686/ 2014 Elsevier Ltd. All rights reserved.

damage and transparent to neutrons. In a typical electrorening


process, a solid stainless steel cathode and a liquid Cd cathode were
used to achieve the recovery of uranium and transuranium
elements, respectively. One of drawbacks of using this liquid Cd
cathode is that the content of Ln in deposited products is relatively
high compared to traditional solvent extraction based processes,
which restrains the propagation of the process in industrial scale.
As for the efcient separation of An from Ln, active solid
aluminium cathode seems to be a promising alternative. It has
been found that the disparity of the deposition potential between
An and Ln is larger on the solid aluminium cathode than that on
other metal electrodes, due to the formation of Al-An alloys [7,8]. In
previous works, a relatively high separation efciency of An over
Ln [7,9,10] and excellent extraction and recovery of An [9,11,12]
have been identied possible on an Al cathode. To establish a
reliable Al cathode based electrochemical pyrochemical process, it
is still quite necessary to investigate the electrochemical properties of representative Ln in AlCl3 contaning melts. Actually, we have
successfully used solid Al cathodes to extract some Ln elements
from melts by forming Al-Ln intermetallic compounds [1316].

88

L.-L. Su et al. / Electrochimica Acta 147 (2014) 8795

On the other hand, dysprosium, as a heavy Ln element, its


electrochemistry in molten salts has not been fully disclosed, only
a few of related studies have been performed [6,1726], although
electrochemical behaviors of it in melts are of importance for the
recovery of dysprosium. In addition, even the electroreduction
mechanism of Dy(III) ions in molten salts is still controversial
according to literature reports. For example, Sala et al. [21] found
that, in LiF-CaF2 molten salts, the reduction of Dy(III) on an inert
molybdenum electrode is a diffusion-controlled one-step process
with three electrons exchanged: Dy(III) + 3e $ Dy(0). Moreover,
Kushkhov et al. [20] reported that the discharge mechanism of
dysprosium chloride in NaCl-KCl melts on a tungsten electrode
was also a three-electron involved process. Similarly, in LiCl-KCl
molten salts, this three electron transferred reduction process of
Dy(III) to Dy metal on inert electrodes were also observed by
Zhang et al. [22], Chang et al. [18], and Konishi et al. [19]. In
contrast, Castrillejo et al. [17] proposed that the electrochemical
reduction of Dy(III) consisted of two consecutive steps: Dy
(III) + 1e $ Dy(II) and Dy(II) + 2e $ Dy(0), on a tungsten
electrode in the same chloride medium at the temperature of
723 K. Some earlier studies also noticed the presence of Dy(II)
species in melts [27]. The complexity of molten salt based
electrochemistry of Dy and different measurement conditions
could account for the discrepancy.
For the extraction of dysprosium, some studies have been
carried out on active electrodes, such as Mg [22,23], Ni [21,24,25]
and Fe [26] electrodes, and corresponding M-Dy alloys have been
prepared. Nevertheless, the report about dysprosium extraction
on an Al electrode was rare. A far as we know, only the formation
of intermetallic compound Al3Dy was observed by Castrillejo
et al. [17] through potentiostatic electrolysis in molten LiCl-KClDyCl3 system. Hence, the basic idea of this work is to reassess the
electrochemical behaviors of dysprosium in LiCl-KCl molten salts
and better understand the formation process and related
mechanisms of Al-Dy alloys through the co-reduction of Dy(III)
and Al(III) ions.

2. Experimental
2.1. General Features
Storage of all reagents and sample preparations were carried out
in a glove box in puried Ar atmosphere where the concentrations
of oxygen content and moisture levels were controlled to be less
than 2 ppm. The experimental cell was carefully prepared. It consists
of an alumina crucible placed in a sealed stainless steel cell inside an
electric furnace and closed by a stainless steel lid cooled inside by
circulating water. A West 3300 programmable device was connected
to control the temperature of the furnace and the temperature
deviation can be maintained below 2 K. A mixture of 100 g
anhydrous LiCl-KCl (44.8:55.2 wt%; AR grade) was introduced into
the alumina crucible and during the experiment the temperature
was monitored in real time with a thermocouple protected by an
alumina tube.
All electrochemical measurements were performed using an
Autolab PGSTAT 302 N potentiostat-galvanostat controlled by a
computer with the Nova 1.10 software package from Metrohm. After
electrolysis, the surface morphology and micro composition of
dysprosium and aluminum of the deposited Al-Dy alloys were
examined by SEM and EDS (HitachiS-4800). To identify the formation
of Al-Dy intermetallic compounds, X-ray diffusion (XRD) (Bruker,
D8 Advance), a surface analysis technology was also performed.
2.2. Preparation and purication of the melts
The anhydrous LiCl-KCl (44.8:55.2 wt. %) mixtures were dried
under vacuum for more than 72 h at 523 K to remove the excessive
water. Then, the melts were rened through pre-electrolysis at
-2.1 V (vs. Ag/AgCl) for 4 h to remove the impurities of other
possible metal cations. Dy2O3 was used as the raw material of Dy
(III) cations following the reaction according to Ref. [28], which
elaborates lanthanide oxide can be chlorided by Al2Cl6 with
lanthanide chloride produced:
Ln2O3 (s) + Al2Cl6 (g) == Al2O3 (s) + 2 LnCl3 (l)

(1)

Fig. 1. A comparison of the CV for pure LiCl-KCl melts (black dotted curve) and LiCl-KCl-DyCl3 (3.73  105 mol cm3) melts (red solid curve). Working electrode: W (surface
area: S0.68 cm2); Temperature: 773 K; Scan rate: 0.1

L.-L. Su et al. / Electrochimica Acta 147 (2014) 8795

Actually, this method has been conrmed in the LiCl-KCl melt


for the preparation of other Ln and An [10,1216,23] chlorides.
In the electrolytic process, anhydrous Dy2O3 and AlCl3 (both, AR
grade) powders were directly added to the LiCl-KCl melt. In LiClKCl melt with the working temperature of 773 K, AlCl3 can be easily
gasied and turn into Al2Cl6, part of them reacts with Dy2O3 to
release Dy(III) ions. This reaction can be represented as:
Dy2O3 (s) + Al2Cl6 (g) == Al2O3 (s) + 2 DyCl3 (l)

(2)

From the thermodynamic data [29], the change of Gibbs energy


of this reaction at 773 K is calculated to be -903.45 kJ mol1. It
reveals that reaction (2) could proceed forward at our experimental temperature. In this work, we rstly explored the electrochemical behaviors of dysprosium on an inert tungsten electrode, hence
after the complete chlorination of Dy2O3, the LiCl-KCl-DyCl3 melt
was puried completely out of AlCl3 by bubbling dry argon
continuously until the ICP-MS analysis of the taken melt sample
shows no remnant Al(III) ion. The concentration of Dy(III) ions in
the LiCl-KCl-DyCl3 melt was measured at the same time. Then the
co-reduction process of Dy(III) and Al(III) were investigated by
increasing the content of AlCl3 in the melts. However, owing to the
volatility of AlCl3, the concentration of AlCl3 we present below is
the initial fractions when AlCl3 were added into the melts.
2.3. Electrochemical electrodes and characterization of cathodic
deposits
A custom-built quartz structure was used to position all of the
electrodes and the thermocouple in molten salt. A silver wire (d
= 1 mm, 99.99%) dipped into the solution of AgCl (1 wt.%) in LiClKCl melts contained in a Pyrex tube was used as the reference
electrode. All potentials were referred to the Ag+/Ag couple. As for
the counter electrode, a 6 mm graphite rod was used. The working
electrode consisted of 1 mm tungsten (W) wire with the lower end
polished by SiC paper. Before each measurement, the working
electrode was cleaned by galvanostatic anodic polarization. The
active electrode surface area was calculated after each experiment
by measuring the immersion depth of the electrode in the molten
salts. As to the electrolysis process, an aluminum plate (Alfa,
99.999%) with thick to be 2 mm was used as cathode. After
electrolysis, the aluminum electrode was abraded and polished by
SiC paper followed by ultrasonic cleaning in ethylene glycol and
ethanol (Sinopharm, 99.8%) in an ultrasonic bath for 15 min and
stored in the glove box before analysis.

89

respectively. The reduction (Ec) occurs in a single sharp peak


mode with a gradual decay, manifesting the deposition of an
insoluble phase [30,31]. The reverse anodic scan shows an
oxidation peak (Ea) with much higher amplitude than the
reduction peak due to the availability of the deposited metal for
the re-oxidation. According to the previous works of Zhang et al.
[22], Chang et al. [18], and Konishi et al. [19], peaks Ea and Ec had
been ascribed to the deposition and dissolution of Dy metal. It is
possible that dysprosium metal would be deposited in a single
direct step by direct reduction of Dy(III) ion into Dy(0).
Furthermore, the reversibility of the reaction of deposition and
dissolution of Dy(III)/Dy(0) was evaluated over a wide scan rate
range from 0.05 to 0.3 Vs1. As shown in Fig. 2a, the peak potential
shifts very slightly with the increasing scan rates. Therefore, the
reduction of Dy(III) to metal should be considered to be a reversible
process. The Nernstian behavior of the reaction at low scan rates
can be further conrmed by plotting the mid-peak potential as a
function of the scan rate. As shown in Fig. 2c, the mid-peak
potential almost remains stable (-1.96 V) at the scan rates of 0.05,
0.1, 0.15 and 0.2 Vs1. In addition, the plot of the cathodic peak
current versus the square root of the sweep rates shows a linear
relationship in Fig. 2b, indicating the process is a diffusion
controlled one. Therefore, it is plausible to use the Berzin-Delahay
equation [32] in this work for a soluble-insoluble couple according
to the theory of linear sweep voltammetry:
Ip = 0.061(nF)3/2C0D1/2V1/2(RT)1/2S

(3)

where n is the number of exchanged electrons; F denotes the


Faraday constant (96,500C mol1); Co represents the solute concentration (mol cm3); D corresponds the diffusion coefcient
(cm2 s1); v designates the potential scanning rate (V s1); T is the
absolute temperature (K) and S corresponds the electrode area
(cm2).
The measurement of the slope of the curve in Fig. 2b yields the
following relation at T = 773 K and C0 = 3.73  105 mol cm3:
Ip
V 1=2

0:068  0:00082AS1=2 V 1=2

(4)

Assuming n = 3, through the combination of Eqs. (3) and (4), the


diffusion coefcient (D) of Dy(III) ions under this condition can be
calculated to be 5.10  106 cm2 s1.

3. Results and discussion


3.1. Electrochemical behavior of Dysprosium Ions on the Tungsten
Electrode
3.1.1. Cyclic Voltammetry
In the present work, investigations of dysprosium began with
CV measurements to establish the nature of the system and the
reversibility of the observed reactions. The typical CV curve of the
pure LiCl-KCl melts is shown in Fig. 1 (dotted curve). The
electrochemical window offered by the LiCl-KCl melts have been
reported to be limited between the reduction of lithium ions
(peak Lc) and the anodic release of chlorine [12,17]. The fact that
there is no other additional peak in its electrochemical window
identies the applicability of the LiCl-KCl melts for our investigations.
Fig. 1 also shows the typical CV of LiCl-KCl-DyCl3 (3.73  105
mol cm3) mixture with the scan rate of 0.1Vs1 on a W working
electrode at the temperature of 773 K. The signals Ea/Ec were
observed in the voltammogram with the reduction peak (Ec) at
-2.04 V and the corresponding anodic peak (Ea) at -1.84 V,

Fig. 2. (a) CVs for 3.73  105 mol cm3 DyCl3 in LiCl-KCl melts at various scan rates.
Working electrode: W (S0.68 cm2); Scan rates: 0.05, 0.1, 0.15, 0.2, 0.25 and 0.3 Vs1.
(b) Plot of the cathodic peak current as a function of the square root of the scan rate.
(c) Mid-peak potential as a function of the scan rate. The dashed curve represents
the average mid-peak potential (-1.96 V vs. Ag+/Ag).

90

L.-L. Su et al. / Electrochimica Acta 147 (2014) 8795


Table 1
Diffusion coefcients of Dy(III) in LiCl-KCl eutectic at 773 K.
Reference

105  D, cm2 s1

Dy(III) Concentration (104 ppm)

Technique

[17]
[6]
[6]
this work
this work

1.47
0.46
0.7
0.51
1.72

2.01
2.0
0.6
0.61
0.57

CP
CP
CP
CV
CP

In addition, it can be found that, there is always a fuzzy preplatform more anodic to the reduction peak Ec in Fig. 1 and 2, which
is very drawn out at all scan rates in Fig. 2, suggesting that this
process could not be diffusion controlled. However, there are still
some discrepancies about the ascription of this pre-platform
[17,21,23]. Ref. [17,23] ascribed it to the reduction of Dy(III) to Dy
(II), whereas Ref. [21] held that the similar pre-platform before the
reduction of Dy(III) ions in LiF-CaF2 melt might be an adsorption
effect of Dy(III) ions to the surface of the working electrode.
Combining the results of following SWV, we prefer to supporting
the pre-platform is attributed to the adsorption effect of Dy(III)
ions to the surface of the W electrode.
3.1.2. Chronopotentiometry
The electrochemical behavior of the redox couple Dy(III)/Dy(0)
was also studied by CP technique. Fig. 3 shows the evolution of the
CPs of DyCl3 in LiCl-KCl melts with the applied current density from
-16 mA to -24 mA on a W electrode. These curves exhibit a single
wave in the same potential range as that observed in the CV curves
and therefore should be associated with the reduction of Dy(III) ions
into metal. In the CP technique, transition time (t) means the time
necessary to observe the complete depletion of the electroactive
species (here the Dy(III) ion), resulting from the diffusion in the layer
of electrolyte at the electrode surface. From Fig. 3a, it can be found
that t decreases with the increase of the applied current density. In
addition, the time-current relationship at a constant value in Fig. 3b
proves the diffusion-controlled process of Dy(III) to Dy(0) and the
validity of Sand's law (Eq. (5)) [32]:

it 1=2 0:5nFSCpD1=2

diffusion coefcient at T = 773 K and C = 3.46  105 mol cm3 was


calculated to be 1.72  105 cm2 s1.
Table 1 gathered the diffusion coefcients of Dy(III) in LiCl-KCl
eutectic at 773 K. Under the same conditions, the diffusion
coefcient calculated by Ref. [17] is approximately three times
higher than that measured by Ref. [6]. The diffusion coefcient of
Dy(III) ion in LiCl-KCl melts determined in this work was 5.1 106
cm2 s1 and 1.72  105 cm2 s1 using CV and CP techniques,
respectively. The complex chemical behaviors of Dy(III) ions in
LiCl-KCl melts, differences in the principles of CV and CP
techniques [33], and imprecise measurement of the wetted length
of the working electrode could account for these discrepancies.
Similar phenomena were observed in the case of cerium and
uranium in Ref. [34,35].
In the above calculation of diffusion coefcient from the results
of CV and CP, the number of exchanged electrons was assumed to
be 3. Actually, the number of exchanged electrons can be deduced
from combining the results of CV and CP, as stated in Ref. [36,37].
By using Eqs. (3) and (5), the following formula was obtained:



Ip
1
0:5 2 D2 C 2 p2 RT
(6)
:
:
:
:
n
1=2 ip1=2
0:61
V
D1 C 21 F
where the concentration (C) and diffusion coefcients (D)
subscripted with 1 and 2 are related to CV and CP, respectively.
According to the results obtained from Fig. 2b Ip =V1=2 0:068 
0:00082AS1=2 V1=2 and Fig. 3b it 1=2 2:5  102 A:S1=2 , the
number of exchanged electrons n3.01 was achieved, almost
entirely consistent with the theoretical expectation.

(5)

where t is the transition time (s), i denotes the applied current (A).
We also assumed the number of exchanged electron n = 3. The

3.1.3. Square wave voltammetry


SWV, with a better accuracy to calculate the number of
electrons exchanged in an electrochemical process, was then
employed to conrm the number of exchanged electrons in this
experiment.
For a single electrochemical reversible process, the differential
intensity measured at each step between the successive pulses
exhibits a Gaussian relationship with the potential. Mathematical
analysis of the Gaussian peak yields a simple equation associating
with the width of the half peak (W1/2) with temperature and the
number of electrons exchanged (n) [38,39]:
W 1=2 3:52

Fig. 3. (a) CPs of LiCl-KCl-DyCl3 (3.46  105 mol cm3) melts at 773 K. Working
electrode: W (S0.68 cm2); Applied current: -16, -18, -20, -22 and -24 mA. (b)
Relationship between the square root of the transition time and the applied current.

RT
nF

(7)

where R is the universal gas constant, T denotes the absolute


temperature in K, n represents the number of exchanged electrons,
and F designates the Faraday's constant.
Fig. 4 shows a typical SWV of DyCl3 (3.73  105 mol cm3) in the
LiCl-KCl melts on a W electrode at the frequency of 20 Hz. A sharp
peak (Ec) associated with the reduction of Dy(III) ion can be clearly
observed in the same potential range as CV and CP. However, peak Ec
is not exactly symmetric as predicted by the theory of nucleation
effect. Due to which, the rise of the current is delayed by the
overpotential caused by the solid phase formation, thus the
increasing part of the differential current is sharper than the
decreasing one. The disturbance of the signal due to the nucleation

L.-L. Su et al. / Electrochimica Acta 147 (2014) 8795


Table 2
Effects of the concentration of DyCl3 on the peak current
ratio.
DyCl3 concentration (105 mol cm3)

iads/idiff

1.761
2.192
2.613

0.518
0.417
0.363

effect could be resolved according to the method introduced in Ref.


[30,40]. That is, only the decreasing part of the peak was taken into
account for the measurement of the width of the half peak, as shown
in Fig. 4 (red dotted curve). After three attempts, the average value of
74 mV for W1/2 was obtained. According to Eqn. (7), n was
determined to be 3.17  0.1 at 773 K, again conrming that the
reduction of DyCl3 in LiCl-KCl melts is a one-step process with three
electrons exchanged.
Similar to the result of CV, there is a pre-platform prior to the
peak Ec in Fig. 4. Interestingly, this pre-platform almost remained
unchanged when we varied the DyCl3 concentration. Table 2 shows
the ratio of iads/idiff at various Dy(III) concentrations. Where iads is
the peak current of the adsorption process and idiff denotes the
intensity of the peak Ec. The ratio of iads/idiff decreases as the
concentration of Dy(III) cations increases. Combining the results of
CV obtained above, the pre-platform should not be a Faraday
electrochemical process associated with the reduction of Dy(III)
ions but could be an adsorption process. One possible explanation
is as follows: the adsorption current can attain a limiting value due
to the adsorption saturation whereas the diffusion current
increases with the increasing concentration of DyCl3. Similar
phenomena had been also observed in the cases of U(IV) [35] or U
(III) [41,42] and Zr(IV) [43] in their alkali metal molten salt. The
adsorption process would largely complicate the electrode
reactions and make them controversial. As for our system, it is
difcult to analyze directly which reaction associated with the preplatform really occurred, due to the little potential difference
between the pre-platform and the reduction peak Ec. Thus big error
would be possible if we analyze the pre-platform by determining
the weak current in the fuzzy platform. Deeper research needs the
combination of transient electrochemical techniques with other
methods such as simulation or spectral analysis. Whatever, the
pre-platform is surface controlled and most likely to be an
adsorption process of Dy(III) ions to the surface of the working
electrode.

Fig. 4. SWV of the LiCl-KCl-DyCl3 (3.73  105 mol cm3) mixture on the W
electrode and the tting of the width (W1/2) of the half peak. Pulse height: 10 mV,
potential step: 5 mV, frequency: 20 Hz.

91

3.2. Co-reduction of Dy(III) and Al(III) ions


3.2.1. Co-reduction principle
The co-reduction refers to a simultaneous reduction of two or
more metallic ions on an inert electrode to form alloys consisting of
at least two metals. Its main characteristic is that a more noble
metal allows the more reactive ones to be reduced at a potential
more anodic than that for the deposition of a pure metal, provided
that they can form intermetallic compounds [44,45]. Actually, the
co-reduction principle has been introduced in molten salt for
nuclear waste reprocessing [10,1316,23,4649] or to act as an
advantageous way to form transition metal-rare earth alloys [19].
In the case of two metallic ions, Rn+ and Np+, where R represents
the more reactive precursor, N corresponds to the less reactive one
and RxNy denotes the alloy formed by the co-reduction process.
Then the co-reduction process can be expressed through the
following reactions:
yNp+ + pye = yN

(I)

yN + xRn+ + nxe = RxNy

(II)

The overall process is:


xRn+ + yNp+ + (nx + py) e = RxNy

(8)

The equilibrium potential of the system R/RxNy can be expressed as


ERn =Rx Ny ERn =R 

RT
InaR inRx Ny 
nF

(9)

where is the equilibrium potential of pure R element, T represents


the absolute temperature in K, n denotes the number of exchanged
electron, F is the Faraday constant (96 500 C) and aR(in RxNy)
designates the activity of R in the intermetallic compound RxNy.
As aR in RxNy is less than one, then the relation ERn =Rx Ny > ERn =R
is clear.
In addition, the co-reduction processes could be inuenced
by the relative concentration ratio of Al(III) to Ln(III) or An(III)
in melts. For example, our previous work [10] found that the
co-reduction of Al(III) with Th(IV) and Eu(III) cations in AlCl3-rich
system allowed a much higher potential disparity between the
formation of Al-Th and Al-Eu alloys. Excellent separation rate
(>99.99%) of Th from Eu was achieved after electrolysis [10].
Moreover, the cathodic deposition product could differ greatly at
various concentration ratios of the two metal cations [12].
Therefore, in this work, the co-reduction of Al(III) and Dy(III)
cations at various concentration ratios was also taken into account
to fully understand the aluminium-dysprosium system, knowing
that ve Al-Dy intermetallic compounds (Al3Dy, Al2Dy, AlDy,
Al2Dy3, AlDy2) are available in the phase diagram [17]. In this case,
dysprosium acts as the more reactive precursor R in the above
equations, while aluminium acts as the less reactive metal N.
3.2.2. Electrochemical behavior
Fig. 5 shows the representative examples of the CVs of LiCl-KClAlCl3 (1.2 or 0.8 wt. %)-Dy2O3 (0.9 wt. %) molten system at 773 K on
the W electrode. Both the two curves in Fig. 5a exhibit large redox
signals labeled with La/Lc and Aa/Ac at almost the same potential
range, respectively. As has been discussed above, signals La/Lc are
associated with the reduction of Li+ cation and its subsequent reoxidation. The reduction peak Ac, at about -1.0 V, and its
corresponding anodic current peak Aa are attributed to the
deposition of aluminum metal and its subsequent dissolution
reaction. The anodic shoulder Ha should owe to the dissolution of
Al-W alloy [12], which underpotentially deposited just prior to the
reduction of Al(III) species. These results were also conrmed by
SWV in Fig. 6 (curve 2 and 3).

92

L.-L. Su et al. / Electrochimica Acta 147 (2014) 8795

Fig. 6. SWVs of LiCl-KCl-DyCl3 (3.73  105 mol cm3) melts (curve 1) and LiCl-KClAlCl3-Dy2O3 (0.9 wt.%) melts with different AlCl3 concentrations 0.8 wt.% (curve 2)
and 1.2 wt.% (curve 3). Working electrode: W (S0.68 cm2); Temperature: 773 K;
Pulse height: 10 mV, potential step: 5 mV, frequency: 20 Hz. Vs1.

Fig. 7. CVs of LiCl-KCl-DyCl3 melts (black dotted curve) and LiCl-KCl-AlCl3-DyCl3


melts (red solid curve) on an Al electrode. Temperature: 773 K; Scan rate: 0.1

Fig. 5. (a) CVs of LiCl-KCl-AlCl3-Dy2O3 (0.9 wt.%) melts with different AlCl3
concentrations 0.8 wt.% (red curve) and 1.2 wt.% (black curve); (b) CVs of the LiClKCl-AlCl3 (0.8 wt.%)-Dy2O3 (0.9 wt.%) melts; (c) CVs of the LiCl-KCl-AlCl3 (1.2 wt.
%)-Dy2O3 (0.9 wt.%) melts at different inversion potentials. Working electrode:
W (S0.68 cm2); Temperature: 773 K; Scan rate: 0.1 Vs1.

However, large differences were observed between signals La/Lc


and Aa/Ac in Fig. 5a. As for the red curve, peak Ec at -2.04 V and Ea at
about -1.83 V can be ascribed to the reduction of Dy(III) to metal and
its subsequent re-oxidation, respectively, according to the above
results of LiCl-KCl-DyCl3 system and Ref. [6,17]. Between peaks Ea/Ec
and Aa/Ac, two broad anodic peaks at about -1.30 and -1.65 V were
observed. The CVs measured subsequently in the same system at
different terminal potentials display a broad anodic peak at

approximate -1.3 V which actually consist of two close peaks Ia and


IIa, similar to their cathodic peaks Ic and IIc (Fig. 5b). In addition, the
SWV in Fig. 6 (curve 2) also revealed the existence of the two close
peaks Ic and IIc at approximate -1.3 V. According to the co-reduction
principle, redox peaks Ia/Ic, IIa/IIc and IIIa/IIIc should be ascribed to the
formation and dissolution of at least three AlxDyy intermetallic
compounds. Moreover, the closer the deposition potential of the
intermetallic compound to that of Dy metal, the higher Dy content
could be formed in the AlxDyy intermetallic compound [12,15].
As for the black curve in Fig. 5a, CV of AlCl3-rich system, much
higher current intensities of peaks Aa and Ac were obviously
observed compared to that of AlCl3-poor system. Besides, other
differences could also be observed in the two curves of Fig. 5a. For
example, the broad anodic peak at -1.3 V in the red curve separated
into two redox peaks Ia/Ic and IIa/IIc in the black curve. The CVs
measured at various cathodic terminal potentials in AlCl3-rich
system (Fig. 5c) and the SWV (curve 3 in Fig. 6) also present two
clearly separated redox peaks Ia/Ic and IIa/IIc. However, the redox
signals IIIa/IIIc and Ea/Ec corresponding to the formation and
dissolution of an intermetallic compound with higher Dy content

L.-L. Su et al. / Electrochimica Acta 147 (2014) 8795

93

Fig. 8. SEM-EDS and XRD results of the potentiostatic electrolysis products of LiCl-KCl-AlCl3 (1.2 wt.%)-Dy2O3 (0.9 wt.%) melts on the Al electrodes: (a) SEM image (deposited
at -1.6 V); (b) XRD pattern (deposited at -1.50 V and -1.60 V); (c) Enlarged SEM image (deposited at -1.6 V); (d) EDS result (deposited at -1.6 V).

and the redox couple of Dy(III)/Dy(0) vanished from both the CV


(Fig. 5c) and SWV (curve 3 in Fig. 6) curves. In the meanwhile, a
new couple of peaks marked as IVa/IVc emerged with its cathodic
and anodic potential at approximate -2.16 and -2.05 V, respectively,
which should be ascribed to be the reduction and oxidation of Al-Li
alloy [12,17]. The reason of the difference between the two curves
in Fig. 5 could be as follows: With the increase of AlCl3
concentration, a much thicker layer of Al was deposited on the
W electrode, which would facilitate the diffusion of Dy metal.
Subsequently, the initially generated intermetallic compound
AlxDyy tends to be transformed into another intermetallic
compounds AlxDyy with high Al content. Therefore, peaks IIIa/
IIIc and Ea/Ec which correspond to the formation/dissolution of an
intermetallic compound with high Dy content and the redox
couple Dy(III)/Dy(0), respectively, could not be observed. In
addition, when the deposited AlxDyy intermetallic compounds
were not fully mantle the Al-covered electrode, Al-Li alloys would
have the chance to be formed [15].
Electrochemical behaviors of LiCl-KCl melts containing both
Al(III) and Dy(III) cations were also investigated on an Al electrode.
Fig. 7 provides a comparison about the CVs of LiCl-KCl-DyCl3 and
LiCl-KCl-AlCl3-DyCl3 melts using Al as the working electrode. The
CV of LiCl-KCl-DyCl3 melts without Al(III) cations (black dotted
curve) is consistent with Ref. [17]. Peaks Ic/Ia are ascribed to the
formation and dissolution of Al-Dy alloys on the Al electrode. The
red solid curve in Fig. 7 shows a typical co-reduction behavior of
Al(III) and Dy(III) ions on the Al electrode, which is very similar to
that obtained in LiCl-KCl-DyCl3 melts, although the peaks become
more bulky. This could be caused by the formation of different AlDy alloys through the co-reduction of Al(III) and Dy(III) cations at
more cathodic potential [50].

3.3. Preparation and characterization of the Al-Dy alloys


To conrm the co-reduction of Dy(III) and Al(III) ions and
examine the formation of AlxDyy intermetallic compounds at
various concentration ratio of Al(III) and Dy(III), both potentiostatic and galvanostatic electrolyses were carried out on a
tungsten electrode. However, only a very small amount of AlDy alloys that adhered to the W electrode could be obtained even
the experiment was repeated for several times. This phenomenon
is probably caused by the small cathode current and the high melt
point of the Al-Dy alloys. Therefore, we further used an Al plate
electrode with the size of 1.5 cm  1.5 cm  0.2 cm for electrolysis.
To prepareAl-Dy alloys at more anodic potential, potentiostatic
electrolysis at -1.4 V, -1.5 V and -1.6 V, each for 3 h, respectively,
was performed. Fig. 8 shows the XRD patterns and the crosssection SEM images coupled with EDS analysis of the cathodic
deposits of potentiostatic electrolysis. It turns out that the
electrolysis at -1.4 V achieved nothing but Al metal, while
electrolysis at -1.5 V and -1.6 V produced a uniform layer covering
on the Al plate electrode (Fig. 8a and c). By XRD analyses (Fig. 8b),
the composition of the deposition layer obtained at -1.5 V was
conrmed to be Al metal and the intermetallic compound Al3Dy
with crystallographic structure of rhombohedral lattice (R-3 m)
(PDF in XRD data base: 180020). When electrolysis was
performed at more negative potential of -1.6 V, the intermetallic
compound Al3Dy with crystallographic structures of R-3 m and
hexagonal lattice (P63/mmc) (PDF in XRD data base: 656363)
could be both obtained.
It is well known that potentiostatic electrolysis has the
advantage of controlling the composition of the compound
produced by the cathodic reaction. According to the co-reduction

94

L.-L. Su et al. / Electrochimica Acta 147 (2014) 8795

Fig. 9. SEM (a,c)-EDS (d) and XRD (b) results of the galvanostatic electrolysis products of LiCl-KCl-AlCl3 (1.2 wt.%)- Dy2O3 (0.9 wt.%) melts on the Al electrode. Current: -50 mA;
Time: 2.5 h; Temperature: 773 K.

behaviors above, at least two kinds of AlxDyy intermetallic


compounds (DyAl3 and DyAl2) could be formed by potentiostatic
electrolysis at -1.5 and -1.6 V, since two very close pairs of
redox peaks were observed in CVs and SWV. However, only one
kind of intermetallic compound Al3Dy was acquired, which was
against with our expectation. The main reason might be the low
current density at our experimental concentration. Under this
condition, even though the intermetallic compound DyAl2 was
formed, the formation rate was much slower than its diffusion
rate, then the transformation of DyAl2 into the more stable Alrich phase (DyAl3) on the Al electrode would take palce. Hence,
ultimately only one kind of intermetallic compound DyAl3 was
observed in Fig. 8, similar phenomena had been observed in the
potentiostatic electrolysis for the preparation of Al-Gd alloys
[15]. The fact that intermetallic compound Al3Dy with more
crystallographic structures was obtained by electrolysis at more
negative potential, to some extent, shows the importance of
nucleation overpotential for the growth of Al-Dy alloys onto the
electrode.
To provide a stable current to equably form more AlxDyy
intermetallic compounds, galvanostatic electrolysis with the
current intensity of -50 mA was also carried out for 2.5 h in our
experiment. During the electrolysis, the cathode potential was
controlled within the range of -1.3 V to -1.75 V to prevent the
deposition of pure Dy and in the meanwhile cover the two much
more anodic redox peaks (Ia/Ic and IIa/IIc) associated with the
formation of AlxDyy intermetallic compounds. As shown in the
SEM image in Fig. 9a, a much thicker layer of approximate 40 mm of
the deposits was obtained than that gained by potentiostatic
electrolysis (Fig. 8a). The XRD result in Fig. 9b conrms that the

deposits are composed of intermetallic compounds DyAl3 and


DyAl, although DyAl2 was still not observed, which proves once
again that DyAl2 could not be stable at this temperature and easily
be transformed into DyAl3. The EDS analyses coupled with SEM in
Fig. 8d and Fig. 9d also conrmed the co-existence of Dy and Al in
the deposits of electrolysis.
4. Conclusions
Electrochemical behaviors of Dy(III) cations on an inert W
electrode were studied in molten LiCl-KCl-DyCl3 salts by
combining various electrochemical techniques (i.e. CV, CP and
SWV). The electroreduction of Dy(III) ions on the tungsten
electrode is a single step process with transfer of three electrons.
The reduction shows a reversible behavior for polarization rates
range of 50  V  300mV1, which is controlled by the diffusion of
Dy(III) cations in solution. Accordingly, the diffusion coefcient of
Dy(III) ion in the LiCl-KCl melts was measured by both CV and CP
techniques. The adsorption effect, which is surface based, was
also observed prior to the reduction of Dy(III) to Dy(0).
The concentration ratio of Dy(III) ions to Al(III) ions has a great
inuence on the co-reduction. In a Dy-rich system, three signals
corresponding to the formation of three AlxDyy were observed on
the tungsten electrode. However, when Al(III) cations were
sufcient, only two of which with higher Al content were
observed. SEM-EDS and XRD characterizations identied intermetallic compound DyAl3 was produced by potentiostatic
electrolysis at -1.5 V and -1.6 V, while two intermetallic compounds DyAl3 and DyAl were obtained through galvanostatic
electrolysis at -50 mA.

L.-L. Su et al. / Electrochimica Acta 147 (2014) 8795

Acknowledgements
[23]

This work was supported by the Major Research Plan Breeding


and Transmutation of Nuclear Fuel in Advanced Nuclear Fission
Energy System of the Natural Science Foundation of China (Grants
91226201, 91126006, 91326202, 11275219) and the Strategic
Priority Research program of the Chinese Academy of Sciences
(Grants XDA03010401, XDA03010403 and XDA03010404).

[25]

References

[26]

[1] A. Stanculescu, IAEA activities in the area of partitioning and transmutation,


Nuclear Instruments and Methods in Physics Research Section A: Accelerators,
Spectrometers, Detectors and Associated Equipment 562 (2) (2006) 614617.
[2] C. Nourry, L. Massot, P. Chamelot, P. Taxil, Electrochemical reduction of Gd(III)
and Nd(III) on reactive cathode material in molten uoride media, Journal of
Applied Electrochemistry 39 (6) (2009) 927933.
[3] K. Kinoshita, T. Inoue, S. Fusselman, D. Grimmett, J. Roy, R. Gay, C. Krueger,
C. Nabelek, T. Storvick, Separation of uranium and transuranic elements from
rare earth elements by means of multistage extraction in LiCl-KCl/Bi system,
Journal of Nuclear Science and Technology 36 (2) (1999) 189197.
[4] M. Salvatores, G. Palmiotti, Radioactive waste partitioning and transmutation
within advanced fuel cycles: Achievements and challenges, Progress in Particle
and Nuclear Physics 66 (1) (2011) 144166.
[5] R. Malmbeck, C. Nourry, M. Ougier, P. Sou9
cek, J. Glatz, T. Kato, T. Koyama,
Advanced fuel cycle options, Energy Procedia 7 (2011) 93102.
[6] K. Sridharan, S. Martin, M. Mohammadian, J. Sager, T. Allen, M. Simpson,
Thermal Properties of LiCl-KCl Molten Salt for Nuclear Waste Separation,
Transactions of the American Nuclear Society 106 (2012) 12401241.
[7] P. Sou9
cek, R. Malmbeck, C. Nourry, J.-P. Glatz, Pyrochemical Reprocessing of
Spent Fuel by Electrochemical Techniques Using Solid Aluminium Cathodes,
Energy Procedia 7 (2011) 396404.
[8] O. Conocar, N. Douyere, J. Lacquement, Extraction behavior of actinides and
lanthanides in a molten uoride/liquid aluminum system, Journal of nuclear
materials 344 (1) (2005) 136141.
[9] P. Sou9cek, R. Malmbeck, E. Mendes, C. Nourry, J.-P. Glatz, Exhaustive
electrolysis for recovery of actinides from molten LiClKCl using solid
aluminium cathodes, Journal of radioanalytical and nuclear chemistry 286 (3)
(2010) 823828.
[10] Y.-L. Liu, Y.-D. Yan, W. Han, M.-L. Zhang, L.-Y. Yuan, R.-S. Lin, G.-A. Ye, H. He, Z.-F.
Chai, W.-Q. Shi, Electrochemical separation of Th from ThO2 and
Eu2O3 assisted by AlCl3 in molten LiClKCl, Electrochimica Acta 114 (2013)
180188.
[11] P. Soucek, R. Malmbeck, E. Mendes, C. Nourry, D. Sedmidubsky, J.P. Glatz, Study
of thermodynamic properties of Np-Al alloys in molten LiCl-KCl eutectic, J Nucl
Mater 394 (1) (2009) 2633.
[12] Y.-L. Liu, Y.-D. Yan, W. Han, M.-L. Zhang, L.-Y. Yuan, K. Liu, G.-A. Ye, H. He,
Z.-F. Chai, W.-Q. Shi, Extraction of thorium from LiClKCl molten salts by
forming AlTh alloys: a new pyrochemical method for the reprocessing of
thorium-based spent fuels, RSC Advances 3 (45) (2013) 2353923547.
[13] K. Liu, L.-Y. Yuan, Y.-L. Liu, X.-L. Zhao, H. He, G.-A. Ye, Z.-F. Chai, W.-Q. Shi,
Electrochemical reactions of the Th4+/Th couple on the tungsten, aluminum
and bismuth electrodes in chloride molten salt, Electrochimica Acta 130 (2014)
650659.
[14] K. Liu, Y.-L. Liu, L.-Y. Yuan, X.-L. Zhao, H. He, G.-A. Ye, Z.-F. Chai, W.-Q. Shi,
Electrochemical formation of erbium-aluminum alloys from erbia in the
chloride melts, Electrochimica Acta 116 (2014) 434441.
[15] K. Liu, Y.-L. Liu, L.-Y. Yuan, X.-L. Zhao, Z.-F. Chai, W.-Q. Shi, Electroextraction of
gadolinium from Gd2O3 in LiClKClAlCl3molten salts, Electrochimica Acta
109 (2013) 732740.
[16] K. Liu, Y.-L. Liu, L.-Y. Yuan, H. He, Z.-Y. Yang, X.-L. Zhao, Z.-F. Chai, W.-Q. Shi,
Electroextraction of samarium from Sm2O3 in chloride melts, Electrochimica
Acta 129 (2014) 401409.
[17] Y. Castrillejo, M. Bermejo, A. Barrado, R. Pardo, E. Barrado, A. Martnez,
Electrochemical behaviour of dysprosium in the eutectic LiClKCl at W and Al
electrodes, Electrochimica acta 50 (10) (2005) 20472057.
[18] K.G. Chang, X.P. Lu, F.Y. Du, M.S. Zhao, Determination of the apparent standard
potential of the Dy/Dy (III) system in the LiCl(c)\\KCl eutectic, Chinese Journal
of Chemistry 12 (6) (1994) 509515.
[19] H. Konishi, T. Nohira, Y. Ito, Morphology control of Dy-Ni alloy lms by
electrochemical displantation, Electrochemical and solid-state letters 5 (12)
(2002) B37B39.
[20] H.B. Kushkhov, A.S. Uzdenova, M.M.A. Saleh, A.M.F. Qahtan, L.A. Uzdenova, The
Electroreduction of Gadolinium and Dysprosium Ions in Equimolar NaCl-KCl
Melt, American Journal of Analytical Chemistry 04 (06) (2013) 3946.
[21] A. Sala, M. Gibilaro, L. Massot, P. Chamelot, P. Taxil, A.-M. Affoune,
Electrochemical behaviour of dysprosium (III) in LiF?CaF2 on Mo, Ni and Cu
electrodes, Journal of Electroanalytical Chemistry 642 (2) (2010) 150156.
[22] M.L. Zhang, Y.S. Yang, W. Han, M. Li, K. Ye, Y. Sun, Y.D. Yan, Electrodeposition of
magnesium-lithium-dysprosium ternary alloys with controlled components

[27]

[24]

[28]

[29]
[30]
[31]

[32]
[33]
[34]

[35]

[36]
[37]
[38]
[39]

[40]

[41]

[42]

[43]

[44]

[45]

[46]

[47]

[48]

[49]

[50]

95

from dysprosium oxide assisted by magnesium chloride in molten chlorides,


Journal of Solid State Electrochemistry 17 (10) (2013) 26712678.
Y.S. Yang, M.L. Zhang, W. Han, P.Y. Sun, B. Liu, H.L. Jiang, T. Jiang, S.M. Peng,
M. Li, K. Ye, Y.D. Yan, Selective electrodeposition of dysprosium in LiCl-KClGdCl3-DyCl3 melts at magnesium electrodes: Application to separation of
nuclear wastes, Electrochimica Acta 118 (2014) 150156.
K. Yasuda, S. Kobayashi, T. Nohira, R. Hagiwara, Electrochemical formation
of DyNi alloys in molten NaClKClDyCl3, Electrochimica Acta 106 (2013)
293300.
H. Konishi, T. Nohira, Y. Ito, Morphology Control of Dy-Ni Alloy Films by
Electrochemical Displantation, Electrochemical and Solid-State Letters 5 (12)
(2002) B37.
H. Konishi, T. Nohira, Y. Ito, Formation of DyFe alloy lms by molten salt
electrochemical process, Electrochimica acta 47 (21) (2002) 35333539.
T. Ogawa, K. Minato, Dissolution and formation of nuclear materials in molten
media, Pure and Applied Chemistry 73 (5) (2001) 799806.
G. Papatheodorou, G. Kucera, Vapor complexes of samarium (III) and
samarium (II) chlorides with aluminum (III) chloride, Inorganic Chemistry
18 (2) (1979) 385389.
I. Barin, Thermodynamic Data for Pure Substance, VCH, Weinheim Germany,
1995.
C. Hamel, P. Chamelot, P. Taxil, Neodymium(III) cathodic processes in molten
uorides, Electrochimica Acta 49 (25) (2004) 44674476.
C. Hamel, P. Chamelot, A. Laplace, E. Walle, O. Dugne, P. Taxil, Reduction process
of uranium(IV) and uranium(III) in molten uorides, Electrochimica Acta 52
(12) (2007) 39954003.
A.J. Bard, L.R. Faulkner, Electrochemical methods: fundamentals and
applications, Wiley, New York, 1980.
G.M. Haarberg, T. Store, R. Tunold, Metal deposition from chloride melts: I.
Rates of diffusion in solvent melt, Electrochimica Acta 76 (2012) 256261.
K.C. Marsden, B. Pesic, Evaluation of the Electrochemical Behavior of CeCl3 in
Molten LiCl-KCl Eutectic Utilizing Metallic Ce as an Anode, Journal of the
Electrochemical Society 158 (6) (2011) F111F120.
P. Masset, D. Bottomley, R. Konings, R. Malmbeck, A. Rodrigues, J. Serp, J.P.
Glatz, Electrochemistry of uranium in molten LiCl-KCl eutectic, Journal of the
Electrochemical Society 152 (6) (2005) A1109A1115.
A.J. Bard, L.R. Faulkner, R. Rosset, D. Bauer, Electrochimie: Principes, mthodes
et applications, Masson (1983) .
R.K. Jain, H.C. Gaur, B.J. Welch, Chronopotentiometry: A review of theoretical
principles, J. Elctroanal. Chem. Interfac. Electrochem. 79 (2) (1977) 211236.
L. Ramaley, M.S. Krause, Theory of square wave voltammetry, Analytical
Chemistry 41 (11) (1969) 13621365.
J.L. Settle, Z. Nagy, Metal deposition-dissolution in molten halides-on the
question of measurability of wery fast electrode-reaction rates 132 (7) (1985)
16191627.
M. Chandra, S. Vandarkuzhali, S. Ghosh, N. Gogoi, P. Venkatesh, G. Seenivasan,
B.P. Reddy, K. Nagarajan, Redox behaviour of cerium (III) in LiF-CaF2 eutectic
melt, Electrochimica Acta 58 (2011) 150156.
K. Serrano, P. Taxil, Electrochemical reduction of trivalent uranium ions
in molten chlorides, Journal of Applied Electrochemistry 29 (4) (1999)
497503.
B.P. Reddy, S. Vandarkuzhali, T. Subramanian, P. Venkatesh, Electrochemical
studies on the redox mechanism of uranium chloride in molten LiCl-KCl
eutectic, Electrochimica Acta 49 (15) (2004) 24712478.
C.P. Fabian, V. Luca, T.H. Le, A.M. Bond, P. Chamelot, L. Massot, C. Caravaca,
T.L. Hanley, G.R. Lumpkin, Cyclic Voltammetric Experiment - Simulation
Comparisons of the Complex Mechanism Associated with Electrochemical
Reduction of Zr4+ in LiCl-KCl Eutectic Molten Salt, Journal of the
Electrochemical Society 160 (2) (2013) H81H86.
F. Krger, Cathodic deposition and characterization of metallic or semiconducting binary alloys or compounds, Journal of The Electrochemical Society
125 (12) (1978) 20282034.
S. Massaccesi, S. Sanchez, J. Vedel, Cathodic deposition of copper selenide lms
on tin oxide in sulfate solutions, Journal of The Electrochemical Society 140 (9)
(1993) 25402546.
H. Tang, Y.-D. Yan, M.-L. Zhang, X. Li, Y. Huang, Y.-L. Xu, Y. Xue, W. Han, Z.-J.
Zhang, AlCl3 aided extraction of praseodymium from Pr6O11 in LiClKCl
eutectic melts, Electrochimica Acta 88 (2013) 457462.
Y.D. Yan, H. Tang, M.L. Zhang, Y. Xue, W. Han, D.X. Cao, Z.J. Zhang, Extraction of
europium and electrodeposition of AlLiEu alloy from Eu2O3 assisted by
AlCl3 in LiClKCl melt, Electrochimica Acta 59 (2012) 531537.
M. Gibilaro, L. Massot, P. Chamelot, P. Taxil, Co-reduction of aluminium and
lanthanide ions in molten uorides: Application to cerium and samarium
extraction from nuclear wastes, Electrochimica Acta 54 (22) (2009)
53005306.
S. Delpech, G. Picard, J. Finne, E. Walle, O. Conocar, A. Laplace, J. Lacquement,
Electrochemical determination of gadolinium and plutonium solvation
properties in liquid gallium at high temperature, Nuclear Technology 163
(3) (2008) 373381.
Y. Castrillejo, P. Fernndez, J. Medina, M. Vega, E. Barrado, Chemical and
electrochemical extraction of ytterbium from molten chlorides in pyrochemical processes, Electroanalysis 23 (1) (2011) 222236.

Vous aimerez peut-être aussi