Tang 2014

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Electrochimica Acta 119 (2014) 120–130

Contents lists available at ScienceDirect

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

Electrochemical behavior of LaCl3 and morphology of La deposit on


molybdenum substrate in molten LiCl–KCl eutectic salt
Hao Tang a,b , Batric Pesic b,∗
a
Key Laboratory of Superlight Materials and Surface Technology, Ministry of Education, College of Materials Science and Chemical Engineering,
Harbin Engineering University, Harbin 150001, China
b
Department of Chemical and Materials Engineering, University of Idaho, Moscow, Idaho 83844, USA

a r t i c l e i n f o a b s t r a c t

Article history: The electrochemical behavior of LaCl3 dissolved in molten LiCl–KCl eutectic salt was studied in the tem-
Received 12 May 2013 perature range of 693–823 K by using inert electrodes, Mo as the cathode, and high density graphite
Received in revised form as the anode. Cyclic voltammetry, chronopotentiometry and square wave voltammetry were used to
26 November 2013
determine major kinetic parameters. The standard reaction rate constant of the order ≈ 10−3 cm s−1 ,
Accepted 27 November 2013
Available online 17 December 2013
determined by Nicholson method, placed the redox reaction of lanthanum in the quasi-reversible range
per Matsuda-Ayabe criteria for practical concept of electrochemical reversibility. Sand’s equation was
used to determine the diffusion coefficient of La(III) ions at four different temperatures. The effect of
Keywords:
Lanthanum chloride electrochemistry temperature on diffusion coefficient obeyed the Arrhenius law, according to which the activation energy
Standard rate constant of La(III)/La(0) for diffusion of La(III) ions was 33.5 ± 0.5 kJ mol−1 . The exchange current density of La(III)/La(0) redox reac-
Matsuda-Ayabe and Nicholson working tion, evaluated at three different temperatures by linear polarization method on Mo and La substrates,
curves was consistently somewhat higher on the later.
Exchange current density of La(III)/La(0) For each temperature, the equilibrium potential of La(III)/La(0) redox couple was determined by
Morphology of lanthanum electrodeposits using open circuit chronopotentiometry, with subsequent calculation of the apparent standard poten-
Nucleation and growth of lanthanum ∗0
tial, ELa(III/La(0)) ∗0
, and the apparent Gibbs free energy, GLaCl The activity coefficients for LaCl3 , LaCl3 was
dendrites. 3
∗0
determined from the difference of apparent and standard Gibbs free energies, GLaCl − GLaCl
0
.
3 3 (SC)
The nucleation mechanism of lanthanum deposition on a molybdenum substrate according to the
electrochemical model of Scharifker-Hill indicated the instantaneous nucleation with three-dimensional
growth of the hemispherical nuclei. Contrary to this, the SEM studies of electrode surface morphology as a
function of electrodeposition time clearly showed that La nucleation and growth follows the mechanisms
responsible for dendritic growth. For the first time, the transient dendritic morphology events were
possible to record, which is the major contribution of this work.
© 2013 Elsevier Ltd. All rights reserved.

1. Introduction of needed technologies. Because these are based on high temper-


atures in order to convert particular solid salts into liquid state
Although the nuclear energy offers to the modern world two electrolyte, the overall term is also known as pyroprocessing. Most
major environmental benefits, the elimination of CO2 emissions of the research thus far has been performed with the binary LiCl–KCl
and the reduced rate of fossil fuel depletion, it still is faced with salts of eutectic composition, as the most promising electrolyte.
some societal issues, such as the radioactive contamination and the Accumulation of lanthanides, as the main fission products, has
proliferation of weapon grade materials. To alleviate these issues, detrimental effects on the efficiency of nuclear fuel because of
the government together with the nuclear power industry needs strong neutron absorption properties. At the same time, from the
to take necessary measures. One avenue is the development of reprocessing point of view, the lanthanides are the most difficult to
technologies for reprocessing of already used nuclear fuel, popu- separate from the actinides because of similarity of their chemical
larly termed as spent nuclear fuel (SNF). In the last 30 years, the and electrochemical properties. To aide toward the better under-
USA and the partnering countries have zeroed their attention onto standing of involving electrochemical reactions, i.e. the more reli-
the application of electrometallurgical principles for development able in-situ monitoring of lanthanides during reprocessing, further
acquisition of basic electrochemical and thermodynamic data on
lanthanides in molten LiCl–KCl systems is needed. Additional ben-
∗ Corresponding author. Tel.: +1 208 885 6569; fax: +1 208 885 7462. efit from such efforts is the extension of gathered knowledge onto
E-mail addresses: pesic@uidaho.edu, pesic48@gmail.com (B. Pesic). the understanding of electrochemical behavior and morphology

0013-4686/$ – see front matter © 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.electacta.2013.11.148
H. Tang, B. Pesic / Electrochimica Acta 119 (2014) 120–130 121

of actinides because the lanthanides are known as their reliable overnight at the natural rate. This procedure ensured perfectly
surrogate representatives. clean crucibles, with no black residues from previous experiments.
Lanthanum, as the first element in the lanthanide series, has the The temperature of the salts was monitored with a
ionic radius of 103.2 pm, which is closest to that of U(III) than any nickel–chromium thermocouple and kept to ±1 K.
other lanthanide [1]. Electrochemistry of lanthanum chloride has
been studied in molten chlorides on a tungsten electrode by several
2.2. Electrochemical apparatus and electrodes
authors in the past twenty years. Castrillejo et al. [2,3] have stud-
ied the electrochemical and thermodynamic properties of LaCl3
All electrochemical measurements were performed using PAR
in LiCl–KCl melts at 450 ◦ C, and CaCl2 –NaCl melts at 550 ◦ C. They
VSP (Bio-Logic SA) electrochemical workstation with the EC-Lab
have determined the lanthanum diffusion coefficient and apparent
V9.54 software package.
standard potentials as well as the nucleation mechanisms. Van-
The reference electrode was fabricated from a silver wire
darkuzhali et al. [4] utilized several different techniques, such as
(d = 0.5 mm, Alfa Aesar 99.99% purity) inserted into a 3 mm diame-
the electrochemical impedance spectroscopy, to study the electro-
ter mullite tube filled with LiCl–KCl eutectic salts containing AgCl
chemical behavior of La(III)/La(0) redox couple in LiCl–KCl molten
(1 wt.%). The bottom of mullite tube was polished by a metallo-
salts. Lantelme et al. [5], Fusselman et al. [6] and Masset et al. [7]
graphic paper to a thickness of about 0.1 mm. Unless otherwise
have used different methods to measure the apparent standard
stated, all potentials were referred to this Ag/Ag+ couple. The
potentials of La(III)/La(0) at several temperatures. Further, several
counter electrode was a graphite rod (d = 3 mm), which was slightly
reports [8–10] describe the utilization of reactive electrodes for
polished prior to each experiment. The working electrode was a
separation and extraction of actinides and lanthanides in molten
polished molybdenum wire (d = 1 mm, Alfa Aesar 99.99% purity),
salts.
or molybdenum disk (d = 3 mm, Alfa Aesar 99.99% purity). All
Regarding the electrochemistry of lanthanum in molten
electrodes were cleaned ultrasonically with deionized water and
LiCl–KCl systems, the literature review points out to polarized
methanol, subsequently dried, and mounted for experimentation.
views on the reversibility of La(III)/La(0) redox reaction, as dis-
Between the experiments, the working electrode was cleaned by
cussed below. Some discrepancy between the found values for
∗0 anodic stripping at +0.3 V for 30 s, followed by at least a 30 s pause to
the apparent standard potential, ELa(III)/La(0) , and the diffusion
ensure the electrode depolarization due to the concentration gradi-
coefficient of La(III) in LiCl–KCl eutectic melts is also evident. ents. The consistency of molten salts was periodically checked by
No data exist for another important kinetic parameter, the cyclic voltammetry. The active electrode surface area was deter-
exchange current density of La(III)/La(0) in molten salts, the essen- mined by controlling the depth of electrode immersion. This was
tial information toward understanding of electrocrystallization done by controlling the vertical path of the micro positioner onto
phenomena. Information about the morphology of lanthanides which the working electrode holder was mounted.
metal deposits on inert electrodes is at best scant. The morphology and micro-zone chemical analysis of lan-
Particular to this work is the use of molybdenum as the inert thanum deposits on a molybdenum disk was determined with SEM
electrode substrate, the exchange current density determination, and EDS (Tescan-Vega II).
and the SEM verification of La electrodeposits. Further, regarding The uniqueness of the given experimental setup is the visibility
the electrochemical deposition from molten salts, very early stages of molten salt electrolyte during the experimentation. This is very
of metal nucleation and assembling of metal nuclei into dendritic helpful when observing the visible reaction products on each of
morphology is for the first time documented on SEM micro- the electrodes and their transfer across the electrolyte. In the pub-
graphs. The results from this study allow us to compare the lished literature the crucibles with the electrolyte and electrodes
kinetic and thermodynamic data of electrochemical phenomena are buried into the furnace with no provision for visual inspection
with those obtained on tungsten, as the popular electrode of the events.
substrate.

3. Results and Discussion


2. Experimental
3.1. Electrochemical behavior of LaCl3
2.1. Preparation and purification of the molten salt
The reduction mechanism of lanthanum chlorides in eutec-
All experimental steps, such as salt preparation, purification, tic LiCl–KCl molten salts was studied by several electrochemical
and experimentation were done in the glove box (VAC-OMNI-LAB) techniques, such as cyclic voltammetry, chronopotentiometry and
filled with high purity argon gas (less than 1 ppm O2 and H2 O). square wave voltammetry.
The eutectic LiCl–KCl (45:55 wt.%) was prepared from dried Fig. 1 shows the cyclic voltammograms obtained on a molyb-
lithium chloride (Alfa Aesar 99 + %) and potassium chloride (Macron denum electrode in the absence and in the presence of LaCl3 .
Chemicals 99.9%) salts. Exactly 20 grams of the eutectic salt was The dashed curve of purified blank salt has no additional reaction
placed into a straight wall alumina crucible, and transferred into a signal within the examined electrochemical window except the
tightly fit graphite jacket for melting by heating with home- made currents corresponding to the reduction/oxidation of Li. Also, the
electric resistance furnace. Molten salt was then taken out poured background current of the blank salts is smaller than 1.0 mA cm−2
into a cold crucible for crystallization. These steps of melting- (about 0.3 mA), confirming that the prepared molten salts are of
crystallization-melting when performed five times improved the acceptable purity to pursue the electrochemistry of LaCl3 studies.
purity of experimental LiCl–KCl salt, as verified by its perfect trans- After the addition of LaCl3 , the solid line clearly shows a new cou-
parency and low background current, less than 1.0 mA cm−2 . ple of redox reactions. The shapes of cathodic wave A at around
Lanthanum was introduced into the bath in the form of anhy- −2.08 V and the associated anodic peak A at −1.95 V are charac-
drous LaCl3 (Alfa Aesar 99.9%) powder. teristic of formation of a new solid phase on an inert electrode,
After the experiments, the alumina crucibles (Aldrich; 20 ml limited by diffusion, followed by the dissolution [11]. In addition,
capacity), were cleaned by soaking in acid, washing and rinsing the ratio of anodic to cathodic peak currents is larger than one.
with deionized water. Upon drying, the crucibles were thermally Fig. 2 shows a cyclic voltammogram for the reduction of LaCl3
treated for several hours at 800 ◦ C. Slow rate cooling was done in the potential range prior to the reduction of lithium. Only one
122 H. Tang, B. Pesic / Electrochimica Acta 119 (2014) 120–130

Fig. 1. Cyclic voltammograms in the absence and in the presence of LaCl3 in molten Fig. 3. (a) A series of voltammograms (corrected with ohmic resistance drop) related
LiCl–KCl salts. LaCl3 = 5.42 × 10−5 M cm−3 , T = 733 K, scan rate = 100 mV s−1 . to the reduction of La(III) (1.10 × 10−4 M cm−3 ) at 693 K on a Mo electrode in molten
LiCl–KCl–LaCl3 salts at different scan rates. Inset: Variation of the difference between
cathodic and anodic peak potentials after ohmic drop correction with the logarithm
redox reaction couple observed implies that the reduction of lan- of the sweep rate.
thanum is a one-step reaction. The inset is the I–t graph developed
from Fig. 2. The integrated values of the cathodic and anodic reac- 0.40 , which is the same value as measured by Cassayre et al.
tions on this I–t curve are −82.14 and 75.43 mAs, respectively, [13], but lower from 0.72 , found by Vandarkuzhali et al. [4]. It
implying that the deposited product is not completely dissolved is customary to test for reversibility of a reaction by checking the
in the anodic process. There are several possible reasons for the linearity of the peak current (ip ) vs. square root of scan rate (1/2 ).
slight observed difference, such as fall-off of loosely attached However, this is a non-discriminatory method because the straight
electrodeposit, shielding effect of wire geometry, effect of uncom- line dependence is achieved for both cases, fully reversible, i.e.
pensated resistance, and the integration of background current in Randles-Sevcik equation [12], or fully irreversible, Nicholson–Shain
cathodic direction. Anyhow, the A/A redox couple corresponds to theory [14]. For quasi-reversible reactions the relationship ip vs.
the deposition and stripping of lanthanum metal on a molybdenum 1/2 is non-linear, in effect making the non-linearity more distinc-
electrode. tive diagnostic than the linearity for the other two mechanisms.
One of the most sought answers regards the reversibility of When the linearity of peak-current density (j) versus square root of
a reaction system, usually obtained by the application of peak- the scan rate (v1/2 ) is tested, open symbols in Fig. 4, the relationship
voltammetry theories. When cyclic voltammograms of lanthanum does show bending with the increase of scan rate, indicating the
chloride were recorded at different scan rates, at 693 K, it was quasi-reversible reaction mechanisms. Matsuda and Ayabe [15] are
observed that with the increase of scan rate the cathodic poten- the first to provide theoretical treatment for the case when the elec-
tial shifted to slightly more negative values and that the cathodic tron transfer rate, compared to the mass transfer rate, is insufficient
current increased in succession, indicating the possibility of par- to maintain Nernstian equilibirium (electrochemical reversibility).
ticipation of a charge transfer step in the overall electrode reaction They have shown that diagnostic parameter  does not change
rate control [12]. Before ensuing further analysis, to avoid an exper- with kinetic parameters, n, D, T and v, but standard rate constant,
imental artifact from the contribution of electrolyte resistance, the ko , does. For quasi-reversible systems the range of  is given by
CVs had to be redrawn, Fig. 3, by compensating for the potential equation (1),
change in question, popularly termed as the IR drop.
Electrochemical impedance spectroscopy was used to measure k0 1
10−2(1+˛) ≤  = ≤ 15 (1)
the ohmic resistance of the electrolyte. The obtained value was (DnF/RT ) v1/2

Fig. 2. Cyclic voltammogram of LaCl3 on a Mo electrode in molten LiCl–KCl–LaCl3


salts. LaCl3 = 1.09 × 10−4 M cm−3 ; T = 723 K; Scan rate =100 mV s−1 . Inset: the evolu- Fig. 4. Variation of the cathodic peak current density with the square root of scan
tion of I-t from the given cyclic voltammogram. rate.
H. Tang, B. Pesic / Electrochimica Acta 119 (2014) 120–130 123

Fig. 5. a-b Nicholson method working curve for determination of standard rate constant (a); Matsuda-Ayabe working curve for transition between quasi-
reversible/irreversible reaction mechanism (b). Standard rate constants determined by Nicholson method are placed on Matsuda-Ayabe plot for determination of a reaction
reversibility type.

To test for quasi-reversible vs. irreversible switchover, as a func- range on the given graph, these values are still quite away from the
tion of scan rate , the best approach is to first produce a working reversible/quasi-reversible borderline, defined by the requirement
curve considering lower limit, 10−2(1 + ˛) , of  in equation (1). The  ≥ 15, i.e. ko ≥ 0.43 v1/2 .
working curve in question, equation (2), It should be pointed out that the methods of Nicholson and
of Matsuda-Ayabe are highly susceptible to uncompensated resis-
ko = 10−2(1+˛) (DnF/RT ) v1/2
2
(2)
tance, i.e. voltage drops, creating the room for significant errors.
for n = 3, and D = 1.7 × 10−5 cm2 s−1 , determined separately below Additionally, the underlying theories developed for soluble/soluble
at T = 733 K, becomes systems can drastically change for the insoluble reaction products.
The effects of nucleation overpotential and deposit morphology
ko = 2.84 × 10−5 v1/2 (3) are the most notable. To avoid these complications, the practical
Equation (3) plot as a function of scan rate, v (V s−1 ), is presented approach is to compare peak-voltammetry experimental behavior
in Fig. (5b). with the predictions from simulations, which has its own chal-
To make the use of Fig. 5b, when determining the reaction lenges. Consequently, one should expect considerable difference
reversibility behavior (a function of scan rate), the standard rate in reported results with respect to lanthanum redox reactions in
constant ko needs to be known. molten salts, as given below.
While the method of Matsuda and Ayabe uses a known standard Glatz et al. [18] suggested that the reduction of La3+ was a
rate constant ko for determining the range of electrochemical reversible and diffusion controlled process for scan rate smaller
reversibility reaction type, the method of Nicholson [16] links than 150 mV s−1 . Kuznetsov [19] claimed that the La(III)/La0 system
the difference between the cathodic (Epc ) and anodic peak (Epa ) is reversible up to a polarization rate of 1.0 V s−1 . However, Cas-
potentials to a function , equation (4), to determine ko of a quasi- trillejo et al. [3] stated that reduction of La(III) is quasi-reversible.
reversible reaction. The numerical simulation results are not consistent either.
According to Lantelme and Berghoute [5] the electrochemical reac-
k0 1
= (4) tion of La(III)/La0 system is rapid, while Vandarkuzhali et al. [4]
1/2 v1/2
( DnF/RT ) calculated that the rate of La(III) reduction in molten LiCl–KCl at
In Nicholson method, also, the first step is to produce a working 798 K is slow, with ko = 4.1 × 10−6 cm s−1 . They concluded that
curve, n
Epp vs. log plot. The
Epp and data, compiled from the reduction of La(III) ion on W-electrode is quasi-reversible for
tables in [16,17], are presented in Fig. 5a. The working curve in scan rates up to 150 mV s−1 , and irreversible for scan rates higher
Fig. 5a can be interpreted as a transitional (quasi-reversible) state than this. Fabian et al. [20] stated, based on numerical simulation,
between very fast reactions (reversible: large log( ), small separa- that La3+ + 3e ↔ La0 is quasi-reversible “beyond doubt”, with the
tion of peaks), and very slow reactions (irreversible: small log( ), standard rate constant ko in the range 5.7 × 10−3 to 3.17 × 10−2 cm
large separation of peaks). Its practical purpose is to determine the s−1 in the temperature range 450 − 550 ◦ C. Although not related to
standard reaction rate constant, ko , from the measured separation La chemistry, but similar, the numerical simulation Kim et al. [21] of
of peaks for cathodic, Epc , and anodic, Epa , sweeps. The separa- Ce(III) + 3e ↔ Ce0 reaction in LiCl–KCl produced 7.5 − 8.4 × 10−3 cm
tion of peaks
Epp plotted as a function of log(v) is presented as s−1 for standard rate constant.
the inset in Fig. 3. One straight line relationship confirms that for Fig. 6 presents a current reversal chronopotentiogram obtained
the reaction La3+ + 3e ↔ La0 there is no mechanism change in the in LiCl–KCl–LaCl3 . The transient curves show that the transi-
studied 50 − 700 mV s−1 scan range. The procedure to calculate the tion time for anodic reaction approximately equals that for
standard rate constant is given in Fig. 5a for a particular tempera- the cathodic reaction. This result verifies the formation of a
ture, T = 733 K. When the found standard rate constants for 50 mV solid insoluble phase on the electrode during the reduction
s−1 and 500 mV s−1 scan rates are placed on Matsuda-Ayabe work- process [22].
ing curve plot in Fig. 5b, bordering quasi-reversible/irreversible Square-wave voltammetry has been employed to determine the
reaction mechanisms, it becomes obvious that the respective val- number of electrons involved in the electrochemical reaction. For a
ues of ko = 1.5 × 10−3 cm s−1 and ko = 2.75 × 10−3 cm s−1 fall deep reversible soluble-soluble system, the square wave voltammogram
into the quasi-reversible predominance region. Although out of is a perfect Gaussian peak. The half-peak width (W1/2 ) is a function
124 H. Tang, B. Pesic / Electrochimica Acta 119 (2014) 120–130

Fig. 6. Current reversal chronopotentiogram of LaCl3 on a Mo electrode in molten Fig. 8. Lanthanum chronopotentiograms on Mo electrode in molten LiCl–KCl–LaCl3 .
LiCl–KCl eutectic salt. Current = −20 m A, T = 733 K, electrode area = 0.34 cm2 , LaCl3 = 1.10 × 10−4 M cm−3 ; T = 693 K; Electrode area = 0.353 cm2 .
LaCl3 = 8.17 × 10−5 M cm−3 .
3.2. Calculation of the diffusion coefficient and exchange current
density of La(III) in LiCl–KCl eutectic melts
of temperature and the number of electrons transferred [23,24].
This relationship is Fig. 8 illustrates chronopotentiograms recorded on a molyb-
denum electrode for LaCl3 (1.10 × 10−4 mol cm s−3 ) in LiCl–KCl
W1/2 = 3.5RT/nF (5) eutectic melts at 693 K at different current intensities. The family of
curves showing a voltage plateau at about −2.10 V, corresponding
where R is the universal gas constant, T is the absolute tempera- to the reduction of La(III) ions into metal, is in agreement with the
ture, n is the number of exchanged electrons, and F is the Faraday’s potential observed in the cyclic voltammograms. After this plateau,
constant. the cathodic potential rapidly increased to reach the limiting value
Fig. 7 shows a series of square-wave voltammograms with corresponding to the deposition voltage of lithium. The transition
different frequencies for the LaCl3 (1.09 × 10−4 mol cm−3 ) in the time for lanthanum reduction was determined by using the method
LiCl–KCl molten salts on a molybdenum electrode at 723 K. Only described in Ref. [26]. Fig. 9 shows that the plots of j vs. −1/2 at
a single reduction peak observed in the electrochemical window four different temperatures exhibit a linear relationship, passing
investigated indicates that the electrodeposition of lanthanum is a through the origin. This indicates that Sand’s equation [13], derived
single-step reaction. The asymmetric shape of this wave is caused for mass transfer controlled electrochemical reactions, is applica-
by nucleation phenomena according to which the nucleation over- ble to the deposition of lanthanum on a molybdenum electrode
potential causes that the right section of the peak is narrower than in LiCl–KCl eutectic. The diffusion coefficient of La(III) was calcu-
the section on the left [25]. Furthermore, the validity of Eq. (5) was lated from the slopes of current density vs. ␶−1/2 plots per Sand’s
verified by plotting the peak current versus the square root of the equation, equation (6)
frequency [23,24]. The straight line through the origin (see inset in
I 1/2 = 0.5nFC0 S 1/2 D1/2 (6)
Fig. 7) confirms that Eq. (5) can be employed to calculate the num-
ber of electrons in the frequency range studied. The found value is where I is the applied current (A), C0 the bulk concentration of La(III)
2.75 which indicates that the number electrons transferred during ion (mol cm−3 ), D is the diffusion coefficient (cm2 s−1 ), F is the Fara-
La(III) reduction is equal to three. day’s constant, S is the surface area of working electrode (cm2 ), n
is the number of exchanged electrons, and is the transition time
(s).

Fig. 7. Net-current square-wave voltammograms for the reduction of La(III) at a


Mo electrode. Pulse height = 25 mV; Potential step = 1 mV; Frequency = 10, 20, and
30 Hz, LaCl3 = 5.42 × 10−5 M cm−3 , T = 723 K. Inset: the linear relationship of j versus Fig. 9. Linear relationship of I versus −1/2 for the chronopotentiometric data given
f1/2 for square-wave voltammograms. in Fig. 8.
H. Tang, B. Pesic / Electrochimica Acta 119 (2014) 120–130 125

Table 1
Diffusion coefficients of La(III) ions in the molten LiCl–KCl eutectic salt at several
different temperatures.

T (K) 693 733 763 813

D × 105 /cm2 s−1 1.22 1.70 2.10 2.88

The slopes of the lines in Fig. 9 were used to calculate the


diffusion coefficients, Table 1, which are further used to pro-
duce the Arrhenius plot, Fig. 10. The resulting Arrhenius slope
produced the activation energy of 33.5 ± 0.5 kJ mol−1 . Somewhat
increased dependence on temperature for the mobility of La(III)
ions can be explained by its complexation with chloride ions,
LaCl6 3− and La2 Cl11 5− being the predominant species, Iwadate et al.
[27]. Also, the activation energy for diffusion of La(III) ions esti-
mated in this work was close to the values obtained by Castrillejo
Fig. 11. Linear polarization plot for LaCl3 in LiCl–KCl eutectic molten salt on Mo-
[2] and Lantelme [5], who found EA to be 32.9 and 30.0 kJ mol−1 ,
electrode (solid line), and Mo-electrode coated with lanthanum metal (dashed line)
respectively. The small difference may be attributed to the differ- at 723, 773 and 823 K.
ence in experimental methods used for the estimation of diffusion
coefficients.
The exchange current density is an important reaction kinet-
ics parameter, linked to the reversibility of a reaction. It is also substrate coated with La metal is somewhat larger than that on
linked to the nucleation characteristics, and morphology of elec- Mo substrate. The crystal structure of molybdenum and lanthanum
trodeposits. For these reasons, it was desired to quantify the order are body-centered cubic (bcc), and hexagonal close-packed (hcp),
of magnitude of exchange current density of the La(III)/La(0) cou- respectively. Therefore, the reason for small exchange current den-
ple in the LiCl–KCl eutectic molten salts. The linear polarization sity on a Mo electrode could possibly be due to the lattice mismatch
method, which utilizes the data from very low overpotentials, was effects. The exchange current densities of cerium and uranium were
employed in the present investigation. The exchange current den- estimated by Marsden and Pesic [28] and Choi et al. [29]. Their val-
sity was determined on two different metal surfaces. One was ues at 773 K fall in the region of 0.04–0.1 A cm−2 for cerium, and
the fresh molybdenum surface, the other was molybdenum sub- 0.02–0.06 A cm−2 for uranium. Compared with our experimental
strate coated with freshly deposited lanthanum metal. At very low data, the exchange current density of La(III) in LiCl–KCl eutectic is
overpotentials, Butler-Volmer equation [12] can be simplified to a closer to uranium than that of Ce(III).
following expression:

j = j0 (nF/RT ) (7) 3.3. Equilibrium potential and apparent standard potential


Fig. 11 shows the linear polarization data at ±15 mV overpo-
tentials and the scan rate of 0.166 mV s−1 on a Mo substrate (solid EMF measurements were performed to determine the equi-
line) and Mo substrate coated with lanthanum metal (dotted line), librium and the apparent standard potentials [30]. For each
at 723, 773 and 823 K. The slope of fitted line can be employed temperature, the molybdenum electrode was coated with lan-
to estimate the exchange current density j0 . The coating was per- thanum film by applying a cathodic potential for 10 s, and the
formed at a 100–150 mV cathodic overpotential versus La(III)/La(0) open-circuit potential (OCP) was recorded versus the Ag/Ag+ ref-
couple for 10 s. Because the surface area of molybdenum substrate erence electrode, Fig. 12. A very stable plateau was obtained each
coated with lanthanum metal couldn’t be accurately determined time, allowing the measurement of the corresponding equilibrium
only geometric area of Mo-electrode was used for conversion to potential Eeq . The measured potential for a metal in equilibrium
current density. The calculated values for exchange current den-
sity j0 are listed in Table 2. The exchange current density on Mo

Fig. 12. Open-circuit potential transient curve for a Mo electrode coated by lan-
Fig. 10. Arrhenius plot: The logarithm of the diffusion coefficient as a function of thanum metal by electrodeposition from LaCl3 (1.07 × 10−4 M cm−3 ) in the LiCl–KCl
inverse of temperature. eutectic at −2.10 V vs. Ag/Ag+ for 10 s at 783 K.
126 H. Tang, B. Pesic / Electrochimica Acta 119 (2014) 120–130

Table 2
Estimation of exchange current density of La electrodeposition on barren Mo and Mo substrate covered with thin La coating.

Temperature (K) LaCl3 concentration (mol cm−3 ) j0 on Mo (A cm−2 ) j0 on Mo coated with La (A cm−2 )
−4
723 1.09 ×10 0.023 0.030
773 1.07 ×10−4 0.034 0.040
823 1.05 ×10−4 0.040 0.057

with its metal chloride, in this case the La(III)/La(0) couple, is deter- Table 4
The comparison of standard enthalpy and entropy of LaCl3 in molten LiCl–KCl.
mined by the Nernst relationship, equation (8).
∗0 ∗0
aLa(III) Source HLaCl (kJ mol−1 ) SLaCl (kJ mol−1 )
eq RT
ELa(III)/La(0) (vs. (Ag/Ag + )/V ) = ELa(III)/La(0)
0 3 3
+ ln (8)
nF aLa(0) Present study −1074.46 0.2393
Lantelme [5] −1059.96 0.231
0
where ELa(III)/La(0) is the standard potential at a hypothetical super- Fusselman [6] −1053.9 −0.1979
Castrillejo [3] −1045.3 −0.1856
cooled liquid reference state of unit mole fraction and unit activity,
Masset [7] −1071.6 0.1375
aLa(III) and aLa(0) are the activity of La3+ ions and of metallic La,
respectively. Unit activity was assumed for pure metal.
∗0
The apparent standard potential, ELa(III)/La(0) of the La redox cou- their data, our experimental data are different by about 10 mV. The
ple is defined as possible reasons are attributed to the different reference electrodes
used by different authors.
∗0 0 RT
ELa(III)/La(0) = ELa(III)/La(0) + ln La(III) (9) The standard potential relationship, equation (13), can be used
nF
to compute the standard Gibbs free energy of the reaction
where La(III) = aLa(III) /XLa(III) is the activity coefficient of La(III) and
XLa(III) is the molar fraction of La(III) in the salt. Combining Eq. (8) La + 3/2Cl2 → LaCl3 (14)
∗0
and Eq. (9), ELa(III)/La(0) can be expressed as according to the relationship
∗0 ∗0
∗0
ELa(III)/La(0)
eq
(vs. (Cl2 /Cl− )/V ) = ELa(III)/La(0) (vs. (Ag/Ag + )/V ) GLaCl = −3FELa(III)/La(0) (kJmol−1 ) (15)
3

∗0
where GLaCl is the Gibbs free energy of the dissolved metal
RT
− ln XLa(III) − EAg/Ag + (vs. (Cl2 /Cl− )/V ) (10) 3
nF chloride, calculated from the experimentally-determined standard
∗0
potential, ELa(III)/La(0) . Gibbs free energy is also a function of tem-
The potential data have been calculated versus the Cl2 /Cl− ref- perature, following the form
erence electrode, according to the following equation: ∗0 ∗0 ∗0
GLaCl = HLaCl − TSLaCl (kJmol−1 ) (16)
3 3 3
− 0 RT
EAg/Ag + (vs. (Cl2 /Cl )/V ) = EAg + ln XAg + (11) ∗0 ∗0
+
nF from which the enthalpy HLaCl and entropy SLaCl can be
3 3
∗0
obtained. When in the present study the values of GLaCl were
Based on extrapolation to infinite dilution of the data of Yang 3
and Hudson [31] at low AgCl concentrations in the LiCl–KCl eutectic plotted versus temperature the following linear relationship was
salts, and other references [6,28], the potential of the reference elec- obtained, equation (17),
trode used in this work (0.0039 mole fraction AgCl) versus Cl2 /Cl− ∗0
GLaCl = −1074.46 + 0.2393T (kJmol−1 ) (17)
is given by the expression 3

∗0 ) and entropy
EAg/Ag + (vs. (Cl2 /Cl− )/V ) = −1.0910 − 1.855 × 10−4 T (K) (12) Therefore, the standard enthalpy (HLaCl
3
∗0 ) values for LaCl in LiCl–KCl eutectic molten salts are
(SLaCl 3
Table 3 summarizes the experimental data obtained from Mo 3

electrode coated with La by open circuit potentiometry. Fig. 13 −1074.46 and 0.2393 kJ mol−1 , respectively. The data from other
shows the variation of standard potential as a function of temper- authors are listed in Table 4 for comparison with the present study.
ature computed from the experimental data given in Table 3. The The activity coefficient of LaCl3 in LiCl–KCl eutectic melts was
experimental data of the present study (shown by the line) can be determined from the difference between the Gibbs free energy of
described by
∗0
ELa(III)La(0) (vs. (Cl2 /Cl− )/V ) = −3.712 + 8.267 × 10−4 T (K) (13)

Several other authors have studied the standard potential of


La(III)/La(0) system. At 750 K, the data reported by Lantelme [5] is
−3.0997 V. At 723 K, the data stated by Fusselman is −3.1462 V [6]
and Castrillejo −3.124 V [3]. At 733 K, the data reported by Masset
is −3.126 V [7], and most recently, at 773 K, the standard potential
found in the report by Sridharan is −3.11 V [32]. Compared with

Table 3
Experimental results for the apparent and standard potential of La(III)/La(0) redox
couple in LiCl–KCl eutectic molten salts.

T (K) Eeq vs. (Ag/Ag+ )/V E*0 vs. (Ag/Ag+ )/V E*0 vs. (Cl2 /Cl− )/V

693 −2.031 −1.919 −3.139


723 −2.005 −1.889 −3.114
753 −1.979 −1.858 −3.089
783 −1.954 −1.828 −3.064
Fig. 13. Variation of the standard potential of La(III)/La(0) with temperature in the
813 −1.929 −1.798 −3.040
LiCl–KCl eutectic molten salt.
H. Tang, B. Pesic / Electrochimica Acta 119 (2014) 120–130 127

Table 5 growth of lanthanum nuclei occurs. While with the cathodic over-
Activity coefficient of LaCl3 at five different temperatures in the LiCl–KCl eutectic
potential increase the value of jm becomes larger the corresponding
melts.
tm decreases. Finally, the current decay with time signifies that the
T (K) G∗0 (kJ mol− ) G(SC)
0
(kJ mol− ) ␥ nucleation and growth processes are controlled by mass transfer
693 −908.60 −899.41 0.203 (Zone III) [2].
723 −901.36 −892.43 0.226 A non-dimensional model developed by Scharifker and Hills
753 −894.13 −885.47 0.251 [36,37] has been widely used to analyze the entire chronoam-
783 −886.89 −878.56 0.278
perograms to determine the type of nucleation mechanisms,
813 −879.94 −871.68 0.295
instantaneous, or progressive, equation (19) and (20), respectively

formation derived from the electrochemical measurements and the  j 2 1.9542


  t
2
Gibbs free energy of formation for pure compounds in the super- = t
1 − exp −1.2564 (19)
jm tm
tm
cooled state, according to the equation:
∗0
2.303RT log LaCl3 = GLaCl 0
− GLaCl (18)  j 2    t 2  2
3 3 (SC) 1.2254
= t
1 − exp −2.3367 (20)
0
where GLaCl is the Gibbs free energy of formation from reac- jm tm
tm
3 (SC)
tion of the pure compounds in the supercooled state. The Gibbs free where j is current density, jm the peak current density, t the
energy of formation of LaCl3 in the supercooled state, calculated by time, and tm the time of peak current. In Fig. 15 the experimental
using the thermochemistry software [33], is summarized in Table 4.
2

The computed activity coefficients, Table 5, are larger than the data chronoamperograms are re-plotted in j j versus tm t
coordi-
m
from Castrillejo [3] and Masset [7]. The differences are mainly due nates, and compared with the theoretical lines described by Eqs.
0
to the literature values for GLaCl The activity coefficient values (19) and (20). It is obvious that the experimental chronoamper-
(SC) 3
ograms fit the model for instantaneous nucleation for all studied
found in the literature are not consistent, as reviewed by Marsden
overpotentials before the value of t/tm of about 2.5. The results
and Pesic, [34].
imply that lanthanum nuclei are generated at the same time on all
active sites available at the start of electrolysis. However, further
3.4. Determination of the nucleation phenomena during
growth of instantaneously formed La nuclei is highly sensitive on
electrodeposition of lanthanum on a Mo substrate in LiCl–KCl
overvoltage difference of only a few millivolts. Castrillejo et al. [2]
eutectic melts
found that nucleation of lanthanum on tungsten was instantaneous
with the growth of nuclei also dependent on overvoltage.
Electrodeposition of a metal on an inert electrode occurs
Although the morphology of electrodeposits during nucleation
through nucleation and growth stages. The nucleation rate defines
and initial growth of nuclei is possible to predict by using the
the crystal grains of the metal deposit, while the growth rate deter-
appropriate electrochemical models, such as those of Scharifker
mines the general appearance and structure of the metal deposit
and Hills, as given above, their validity still needs the visual confir-
[35].
mation. In other words, the electrode surface has to be inspected by
Chronoamperometry is a technique that can provide valuable
microscopy, like atomic force or scanning electron microscopy. To
information about the kinetics of electrodeposition. When a solid
do so, regarding the molten salt systems, there are three technical
phase is deposited on a foreign inert substrate, it needs overpoten-
hurdles to surpass: The electrode surface has to be (1) free from salt
tial to form isolated nuclei, and the resulting j-t transients can be
solidified upon electrode withdrawal, with the nuclei as the only
used to provide the information whether the nucleation and growth
features available for imaging, (2) sufficiently smooth to be able
mechanisms are progressive, or instantaneous.
to distinguish the nuclei from mechanically introduced imperfec-
Fig. 14 shows the j-t curves of La(III) obtained at various poten-
tions such as scratches from polishing, and (3) isolated from oxygen
tial pulses on a Mo electrode. The current in the first part (Zone
and moisture in the lab atmosphere during its transfer from the
I) decreases very fast. This process is related to charging the elec-
trode double layer. Then, the current quickly increases to reach the
maximum value jm (Zone II), where the formation and subsequent

Fig. 15. Comparison of the dimensionless experimental data for several overpoten-
Fig. 14. Potentiostatic current-time curves for lanthanum on Mo electrode in tials derived from the current-time transients (Fig. 14) with the theoretical models
LiCl–KCl eutectic molten salt at various overpotentials: (a) −1.997 V; (b) −2.000 V; for instantaneous (dashed line) and progressive (dotted line) nucleation mecha-
(c) −2.003 V; (4) −2.005 V. LaCl3 = 1.09 × 10−4 M cm−3 , T = 733 K. nisms.
128 H. Tang, B. Pesic / Electrochimica Acta 119 (2014) 120–130

Fig. 16. SEM examination of morphology of La electrodeposited on a Mo-disk electrode at −2.04 V and 723 K as a function of deposition time: (a) 0.5 s, (b) 2.5 s, (c) 5 s, (d)
10 s and (e) 20 s.

glove box to the SEM chamber. Among these, producing the elec- The concept of using solvents to remove the solidified salt from
trode surface free of salt is the most difficult to achieve, further the electrode surface was abandoned after several dozen of combi-
explaining the absence of photomicrographs in the literature nation of various organic mixtures produced unsatisfactory results.
depicting the electrode surface during nucleation. Even when a cocktail of organic solvents produced satisfactory
H. Tang, B. Pesic / Electrochimica Acta 119 (2014) 120–130 129

LiCl–KCl–LaCl3 salt dissolution rate, its reactivity with electrode- the surface, for which longer deposition times would have been
posited La made it ineligible. required.
Because, experimentally, it is very difficult to polish a wire Back to the Scharifker-Hill model and the data in Fig. 15.
type of electrode to remove the mechanical scratches and surface Regarding the molten salt systems, the published literature on
imperfections, which can have profound effects on the structure electrodeposition of lanthanides and actinides resorts to the
and homogeneity of the metal deposit [38], the Mo disk elec- Scharifker-Hill equation as an indicator of involving nucleation
trode was used instead. Mo disk electrode was prepared from a mechanisms, but commonly, without the microscopic verification
3 mm D × 10 mm L Mo cylinder, connected to stainless steel wire of the features. The published chronoamperometric data indiscrim-
(1.58 mm D × 5.5 mm L) by spot welding. After shielding the cylin- inately followed the models for the 3D hemispherical nucleation
der wall, the open end provided the circular geometry. The open and growth mechanisms, as our data did in Fig. 16. Obviously this
end was carefully polished (1.2 micron SiC and 1.0 micron and 0.05 would have been a wrong conclusion, if left unchecked, as Fig. 16a-e
micron Al2 O3 were used in succession) to a mirror like condition. clearly shows that it is the dendrites that govern the morphol-
It was found that the freshly and well-polished surface of Mo did ogy. As the reminder, the derivation of instantaneous nucleation
not retain molten salt upon its withdrawal. This was an enabling model requires that all nucleation sites on the surface are instanta-
breakthrough toward the use of SEM for morphological studies, neously activated by a switched potential, distinctly spaced away
eliminating the need of using solvents for salt removal. The last from each other, hemispherical, each with its own diffusion zone.
step, the transfer of electrode from the glove box to the SEM stage, If so, why Scharifker-Hill model, derived for hemispherical nuclea-
involved the use of a specially prepared electrode cover. The elec- tion and growth, describes so well the La nucleation and growth
trode, covered while in the glove box, was taken out and transferred with dendritic morphology? The answer lies in understanding that
onto the SEM stage, whose chamber was continually vented with Scharifker and Hill actually simplified 3D hemispherical geome-
argon gas. Combined, the removal of electrode cover, after mount- try to a 2D problem by considering the equivalent geometric area
ing on the stage, and the closure of SEM chamber took less than of plane towards which linear diffusion transfers same amount of
2-3 seconds, sufficiently short time to avoid possible reactions with material that would be transferred to hemispherical diffusion of
oxygen and moisture in the lab atmosphere. growing center.
The SEM images of lanthanum during the first seconds of Fig. 16b shows that the dendritic arrangement of La clusters of
electrodeposition are presented in Fig. 16a–e. The quality of pho- ad-atoms is concentric, with radial growth across the electrode.
tomicrographs is exceptional, considering no further treatment was In other words, the 2D geometric area prerequisite needed by
needed after the electrode left the molten salt electrolyte, and Scharifker-Hill model is already met. The preceding interpreta-
highly informative, with the obvious conclusion that the nucleation tion explains why the equation (15) developed for instantaneous
and growth of La electrodeposits follow the mechanisms applicable nucleation fits the experimental data from Fig. 14 in this work and
to dendritic morphology. The full coverage of electrode surface, the in the literature on electrodeposition of actinides and lanthanides
planar 2D growth, is achieved in less than 20 seconds, Fig. 16e. from molten salts.
The electrode surface after only 0.5 seconds of deposition was The final note to make is that the above results with respect
inspected to determine if the dendrite origination sites could be to dendritic growth of lanthanum electrodeposits can readily be
identified. Fig. 16a shows numerous sites around which the den- extrapolated onto the electrocrystallization mechanisms of ura-
drite precursors are lined up for the upcoming arrangement in the nium and plutonium in molten salts because lanthanum is an
directions corresponding to dendritic morphology. The dendrites excellent surrogate for these two actinide metals.
are just about to be formed. The white features on all these micro-
graphs are the salt that was not removed on its own; their size and
position is in correlation to the sites with the highest concentration 4. Conclusions
of La metal. After 2.5 seconds, Fig. 16b, the clusters of lanthanum
adatoms, as the dendrite precursors, are positioned into direction It was verified, for the first time, that electrodeposition of lan-
defined by dendritic growth. Away from the already defined den- thanum from molten salts follows the nucleation mechanisms that
drites, the density of dendrite precursors is highest in the positions lead to dendritic morphology. The circular dendrite nucleation cen-
closest to the tips of growing dendrites. Further away from the cen- ters grow radially across the electrode surface until the 2D diffusion
ters, their density is lower but clearly visible. The inspection of very zones overlap, after which no further nucleation growth happens.
important and informative Fig. 16b also indicates that the distribu- This is the first chronological presentation of nuclei formation and
tion of dendrite nucleation centers are uniformly spaced and sized, growth in molten salts regardless of a metal being studied.
and that the dendrites around each center will grow radially across The confirmed 2D diffusion zones are the explanation why the
the electrode surface, creating the conditions for 2D linear diffu- Scharifker-Hill model developed for the instantaneous nucleation
sion. It is difficult to predict the sites for the initiation of La nuclei. mechanisms is also applicable to electrodeposition of lanthanum.
Once formed, the dendrite precursors appear to be arranged along The results from the morphology studies of electrodeposition of
the surface imperfections, particularly the scratches from polish- La on Mo substrate are applicable to the nucleation mechanisms of
ing. In the absence of scratches, the next energetically preferred actinides and other lanthanides metals.
sites are the lattice surface imperfections. In all these situations, The reduction of LaCl3 in LiCl–KCl eutectic on a molybdenum
the adatoms are repeatedly incorporated into the lattice of a den- electrode is a one-step, three-electron transfer, reaction. The redox
drite by repeated one-dimensional nucleation. As stated by Popov reaction La3+ + 3e ↔La0 is of quasi-reversible nature, according to
et al. [39], the electrochemical as well the crystallographic condi- Nicholson [16] and Matsuda-Ayabe [15] analytical protocols. The
tions under which dendritic deposits are formed can be precisely diffusion coefficients, calculated from Sand equation in the tem-
determined. The unresolved issue is what causes the dendritic pre- perature range of 693–813 K, showed temperature dependence in
cursors to appear at regularly spaced locations along the dendrite compliance with the Arrhenius law. The activation energy for dif-
stem. fusion of La(III) ions in LiCl–KCl eutectic was 33.5 ± 0.5 kJ mol−1 .
With longer deposition times, the dendrites continue to expand The exchange current density of La(III)/La(0) was estimated
toward the full surface coverage, Fig. 16c–e. Again, the growth by linear polarization method on molybdenum and lanthanum
is characterized by spreading across the electrode surface with surfaces at three different temperatures. The obtained values are
very little evidence for 3D growth, i.e. protrusion away from between 0.023 and 0.057 A cm−2 .
130 H. Tang, B. Pesic / Electrochimica Acta 119 (2014) 120–130

The equilibrium potential was measured between 693 and 813 K [14] R.S. Nicholson, I. Shain, Theory of stationary electrode polarography, Analytical
from the recorded open-circuit potential of La(III)/La(0). At 723 K, Chemistry 36 (1964) 706.
[15] H. Matsuda, Y. Ayabe, Zur theorie der Randles-Sevcikschen kathodenstrahl-
the apparent standard potential of La(III)/La(0) system in LiCl–KCl polarographie, Zeitschrift fur elektrochemie 59 (1955) 494.
eutectic melts is −3.114 V vs. Cl2 /Cl− . The standard Gibbs free [16] R.S. Nicholson, Theory and application of cyclic voltammetry for measurement
energy, as well as the enthalpy and entropy of LaCl3 in LiCl–KCl of electrode reaction kinetics, Analytical Chemistry 37 (1965) 1351.
[17] S.P. Perone, Evaluation of statioanary electrode polarography and cyclic
eutectic melts, were deduced from the apparent standard poten- voltammetry for the study of rapid electrode processes, Analytical Chemistry
tials for La(III)/La(0). The activity coefficients of LaCl3 in LiCl–KCl 38 (1966) 1158.
eutectic, calculated in the temperature range of 693–813 K, are [18] J.-P. Glatz, R., Malmbeck, C., Pernel, C., Scheppler, J. Serp, Electrochemical
behavior of lanthanides in LiCl–KCl eutectic melts, ITU-CRIEPI, European Com-
between 0.203 and 0.295.
mision, JRC, Institute for Transuranium Elements, Project N◦ : FIS5-1999-00199,
September 2003.
Acknowledgments [19] S.A. Kuznetsov, H. Hayashi, K. Minato, M. Guane-Escard, Electrochemical tran-
sient techniques for determination of uranium and rare-earth metal separation
coefficients in molten salts, Electrochimica Acta 51 (2006) 2463.
Hao Tang gratefully acknowledges the financial support from [20] C. P. Fabian, V., Luca, P., Chamelot, L., Massot, C., Caravaca, G. R. Lumpkin, Exper-
China Scholarship Council, and the acceptance/support by the imental and Simulation Study of the Electrode Reaction Mechanism of La3+ in
LiCl–KCl Eutectic Molten Salt, Journal of The Electrochemical Society 159 (2012)
Chemical and Materials Engineering Department at the University
F63.
of Idaho, USA. [21] T.-J. Kim, D.-H. Ahn, S.-W. Paek, Y. Jung, Study on electrodeposition of Ce(III) at
a tungsten electrode in a LiCl-KCl molten salt solution, International Journal of
Electrochemical Science 8 (2013) 9180.
References
[22] H.B. Herman, A.J. Bard, Cyclic chronopotentiometry. Diffusion controlled elec-
trode reaction of a single component system, Analytical Chemistry 35 (1963)
[1] H.C. Aspinall, Chemistry of the f-Block Elements, CRC, 2001. 1121.
[2] Y. Castrillejo, M.R. Bermejo, A.M. Martinez, P. Diaz Arcas, Electrochemical [23] J.G. Osteryoung, R.A. Osteryoung, Square wave voltammetry, Analytical Chem-
behavior of Lanthanum and yttrium ions in two molten chlorides with dif- istry 57 (1985) 101A.
ferent oxoacidic properties: the eutectic LiCl–KCl and the equimolar mixture [24] L. Ramaley, M.S. Krasue, Theory of square wave voltammetry, Analytical Chem-
CaCl2 –NaCl, Journal of Mining and Metallurgy 39 (1-2 (B)) (2003) 109. istry 41 (1969) 1362.
[3] Y. Castrillejo, M.R. Bermejo, E. Barrado, A. Martinez, P. Diaz Arocas, Solubili- [25] K. Serrano, P. Taxil, Electrochemical reduction of trivalent uranium ions in
zation of rare earth oxides in the eutectic LiCl–KCl mixture at 450 C and in the molten chlorides, Journal of Applied Electrochemistry 29 (1999) 497.
equimolar CaCl2 –NaCl melt at 550 C, Journal of Electroanalytical Chemistry 545 [26] R.W. Laity, J.D.E. McIntyre, Chronopotentiometric diffusion coefficients in fused
(2003) 141. salts I. Theory, Journal of the American Chemical Society 87 (1965) 3806.
[4] S. Vandarkuzhali, N. Gogoi, S. Ghosh, B.P. Reddy, K. Nagarajan, Electrochemical [27] Y. Iwadate, H. Matsuura, A. Kajinami, K. Takase, N. Ohtori, N. Umesaki, R.
behaviour of LaCl3 at tungsten and aluminium cathodes in LiCl–KCl eutectic Fujita, K. Mizuguchi, H. Kofuji, M. Myochin, Local structure analyses of molten
melt, Electrochimica Acta 59 (2012) 245. lanthanum trichloride-Alkali chloride ternary systerms: Aproaches from fun-
[5] F. Lantelme, Y. Berghoute, Electrochemical studies of LaCl3 and GdCl3 dissolved damentals to pyrochemical reprocessing, Electrochemistry 77 (2009) 736.
in fused LiCl–KCl, Journal of the Electrochemical Society 146 (1999) 4137. [28] K.C. Marsden, B. Pesic, Evaluation of the electrochemical behavior of CeCl3
[6] S.P. Fusselman, J.J. Roy, D.L. Grimmett, L.F. Grantham, C.L. Krueger, C.R. Nabelek, in molten LiCl–KCl eutectic utilizing metallic Ce as an anode, Journal of the
T.S. Storvick, T. Inoue, T. Hijikata, K. Kinoshita, Y. Sakamuraet, K. Uozumi, T. Electrochemical Society 158 (2011) F111.
Kawai, N. Takahashial, Thermodynamic properties for rare earths and ameri- [29] I. Choi, B., Serrano, S., Li, S., Herrmann, S. Phongikaroon, Paper 9045 presented
cium in pyropartitioning process solvents, Journal of the Electrochemical at Proceedings of Global 2009, Paris, France, Sept. 6–11, 2009.
Society 146 (1999) 2573. [30] L. Cassayre, J. Serp, P. Soucek, R. Malmbeck, J. Rebizant, J.P. Glatz, Electrochem-
[7] P. Masset, R. J.M. Konings, R., Malmbeck, J., Serp, J.-P. Glatz, Thermochemical istry of thorium in LiCl–KCl eutectic melts, Electrochimica Acta 52 (2007) 7432.
properties of lanthanides (Ln = La, Nd) and actinides (An = U, Np, Pu, Am) in the [31] L. Yang, R.G. Hudson, Some investigations of the Ag/AgCl in LiCl–KCl eutectic
molten LiCl–KCl eutectic, Journal of Nuclear Materials 344 (2005) 173. reference electrode, Journal of the Electrochemical Society 106 (1959) 986.
[8] M. Kurata, Y. Sakamura, T. Hijikata, K. Kinoshita, Distribution behavior of [32] K. Sridharan, Thermal properties of LiCl–KCl molten salts for nuclear waste
uranium, neptunium, rare-earth elements (Y, La, Ce, Nd, Sm, Eu, Gd) and separation, NEUP Final Report, Project No. 09-780, Nov. 30, (2012).
alkaline-earth metals (Sr,Ba) between molten LiCl–KCI eutectic salt and liquid [33] HSC Chemistry 6.0, Chemical reaction and equilibrium software with extensive
cadmium or bismuth, Journal of Nuclear Materials 227 (1995) 110. thermochemical database and flowsheet simulation, Outokumpu Research Oy.,
[9] H. Moriyama, H. Yamana, S. Nishikawa, S. Shibata, N. Wakayama, Y. Miyashita, 1974–2006.
K. Moritani, T. Mitsugashira, Thermodynamics of reductive extraction of [34] K. C. Marsden, B. Pesic, Rare earth elements in fused LiCl-KCl eutectic-A crit-
actinides and lanthanides from molten chloride salt into liquid metal, Journal ical literature review, 51st Annual Conference of Metallurgists, COM 2012,
of Alloys and Compounds 271–273 (1998) 587. September 30-Oct. 3, 2012, Niagara Falls, Canada.
[10] J. Serp, M. Allibertb, A.L., Terriera, R., Malmbecka, M., Ougiera, J., Rebizanta, J.- [35] E. Budevski, G. Staikov, W.J. Lorenz, Electrocrystallization: Nucleation and
P. Glatza, Electroseparation of actinides from lanthanides on solid aluminum growth phenomena, Electrochimica Acta 45 (2000) 2559.
electrode in LiCl–KCl eutectic melts, Journal of the Electrochemical Society 152 [36] G. Gunawardena, G. Hills, I. Montenegro, B. Scharifker, Electrochemical nuclea-
(2005) C167. tion: Part I. General considerations, Journal of Electroanalytical Chemistry 138
[11] D. Pletcher, R. Greef, R. Peat, L.M. Peter, J. Robinson, Instrumental Meth- (1982) 225.
ods in Electrochemistry, Southampton Electrochemistry Group, University of [37] B. Scharifker, G. Hills, Theoretical and experimental studies of multiple nuclea-
Southampton, Horwood, London, 2001. tion, Electrochimica Acta 28 (1983) 879.
[12] A.J. Bard, L.R. Faulkner, Electrochemical Methods- Fundamental and Applica- [38] G. Staikov, W.J. Lorenz, The role of crystal imperfections in electrochemical
tions, Wiley, New York, 2001. phase formation and growth, Canadian Journal of Chemistry 75 (1997) 1624.
[13] L. Cassayre, J. Serp, P. Soucek, R. Malmbeck*, J. Rebizant, J.-P. Glatz, Electro- [39] K.I. Popov, S.S. Djokic, B.N. Grgur, Fundamental Aspects of Electrometallurgy,
chemistry of thorium in LiCl-KCl eutectic melts, Electrochimica Acta 52 (2007) Kluwer Academic/Plenum Publishers, 2002, ISBN 0-306-47269-4, pp. 78-100.
7432–7437.

You might also like