You are on page 1of 7

Journal of Electroanalytical Chemistry 658 (2011) 1824

Contents lists available at ScienceDirect

Journal of Electroanalytical Chemistry


journal homepage: www.elsevier.com/locate/jelechem

Electrodeposition of Co, Sm and SmCo from a Deep Eutectic Solvent


E. Gmez a,, P. Cojocaru b, L. Magagnin b, E. Valles a
a
b

Departament de Qumica Fsica and Institut de Nanocincia i Nanotecnologia (IN2UB), Universitat de Barcelona, 08028 Barcelona, Spain
Dipartimento di Chimica, Materiali e Ingegneria Chimica Giulio Natta, Politecnico di Milano, 20131 Milano, Italy

a r t i c l e

i n f o

Article history:
Received 1 March 2011
Received in revised form 15 April 2011
Accepted 19 April 2011
Available online 28 April 2011
Keywords:
Electrodeposition
Cobalt
Samarium
Deep Eutectic Solvent

a b s t r a c t
The suitability of 1 choline chloride:2 urea mixture, Deep Eutectic Solvent (DES) for the electrodeposition
of cobalt, samarium and cobaltsamarium system has been studied. Its electrochemical window permits
deposition analysis to be carried out without interference from parallel reactions. Deposition was studied
at 70 C in order to stimulate mass transfer and to lower solution viscosity.
Cobalt deposits according to a nucleation and three dimensional growth mechanism, all its characteristic features do appear for all cases studied. Samarium deposition takes place through a more complex
process in which a rst potential range is found where the species formed limit the conductive character
of the substrate. When potential becomes more negative, normal behaviour is observed. When both
cobalt and samarium are present in the solution, codeposition occurs, at no potential value, any current
diminution of the current was recorded under stirred conditions. Deposits of SmCo show different morphology and composition depending on applied potential. Nodular cobalt-rich deposits are obtained at
low deposition potentials whereas ne grained samarium-rich ones are obtained at more negative deposition potentials.
2011 Elsevier B.V. All rights reserved.

1. Introduction
Electrodeposition of some metals has been restricted in aqueous
solutions at moderate temperatures to those presenting a standard
potential less negative than that of water reduction. Some of these
metals could be deposited, albeit expensively, using either organic
solvents or more drastic conditions such as molten salts [15].
In the last decade research in ionic liquids (ionic materials with
melting point below 100 C), with a wide electrochemical window,
has made it possible to deposit metals with very negative standard
potentials and their alloys [614]. Metallic coatings that up to then
had been impossible to deposit became obtainable. Different generations of ionic liquids have been developed widening the spectrum of depositable metals [1519]. However, many of them do
require a complex synthesis and a very accurate electrochemical
work, since the majority of the ionic liquids present high sensibility
to water or degradation by oxygen.
It has been shown recently that it is possible to create an ionic
uid mixing quaternary ammonium halides with an amide, carboxylic acid or alcohol moiety [2023]. Such mixtures are not,
strictly speaking, room-temperature ionic liquids (RTIL) since in
general contain an uncharged molecular component, so that, the
term Deep Eutectic Solvent (DES) was coined by Abbott [24].
Unlike the ionic liquids these room-temperature eutectic mixtures
Corresponding author.
E-mail address: e.gomez@ub.edu (E. Gmez).
1572-6657/$ - see front matter 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.jelechem.2011.04.015

are easy to prepare in pure state. They are not water reactive and
their biodegradability is proven. Moreover, the economic investment involved is much lower than that incurred using roomtemperature ionic liquids. However, little information is available
about their physicchemical properties [25].
The microelectronic industry employs many kinds of coatings in
specic parts of the devices, magnetic materials being the most popular ones. Important advances has been achieved in the preparation
of soft magnetic materials but the electrodeposition of hard magnetic ones is still restricted to a few ones [2628]. The challenge
in preparing such alloys by electrodeposition is that these alloys
contain metals that have electrocatalytic character to hydrogen
evolution, very negative standard potential, or both. Then, the use
of water-free solvents with a wide electrochemical window is welcome in this eld and its possibilities are worthy of study.
In this work we select to study the cobaltsamarium electrodeposition processing possibilities in DES medium, as this alloy is
potentially a hard magnetic material depending on the metal ratio
[2931]. Few studies of SmCo electrodeposition have been performed in aqueous medium [32] since samarium has a very
negative standard potential leading to simultaneous hydrogen evolution. A study about the preparation of SmCo by electrodeposition
in urea-acetamide-NaBr-MClx melts has been performed by Liu et
al. [33]. Our interest is to analyse the possible electrodeposition of
cobalt, samarium and cobaltsamarium systems using the eutectic
1 choline chloride:2 urea mixture as solvent due to its easy preparation, good working conditions and non aggressive nature.

E. Gmez et al. / Journal of Electroanalytical Chemistry 658 (2011) 1824

19

capillary containing the DES solvent. Stable and reproducible values of the potential were obtained with this reference electrode.
Voltammetric experiments were carried out at 50 mV s 1, scanning at rst to negative potentials. Only one cycle was run in each
voltammetric experiment. Before and during experiments solutions were de-aerated with argon. Work temperature was kept
constant at 70 C to favour low viscosity and high conductivity of
the solvent. Magnetic stirrer was used when the agitation effect
was tested.
Deposits morphology were observed using Hitachi S-2300 scanning electron microscope. Elemental composition was determined
with an X-ray analyser incorporated in Leica Stereoscan S-360
equipment and by X-ray uorescence (XRF).
3. Results

Fig. 1. Cyclic voltammograms in DES solvent at: (a) black: immediately after
desoxigenation, (b) dashed: after two hours of rst scan, and (c) grey: after 50 h of
rst scan.

2. Experimental
Solvent was prepared using choline chloride (from Across
Organics) and urea (from Merck) of analytical grade. The solids,
in the molar proportion 1 choline chloride:2 urea, were warmed
and removed constantly to achieve the liquid state of the deep eutectic solvent (DES). Samarium nitrate from Aldrich, and cobalt
chloride from Fluka both of analytical grade were used as source
of electroactive species. Prior to dissolution, cobalt salt was maintained in a stove at 110 C in order to assure maximum dehydration. Solutions were prepared using the DES solvent.
A cylindrical three electrode cell of one single compartment was
used. Electrochemical experiments were carried out using an Autolab with PGSTAT30 equipment and GPES software. Working electrodes were vitreous carbon rods (from Metrohm). Vitreous
carbon electrode was polished to a mirror nish using alumina of
different grades (3.75 lm and 1.87 lm), cleaned ultrasonically
for 2 min in water and dried with air prior to be immersed in the
solution. The counter electrode was a platinum spiral. The reference electrode was an Ag|AgCl/NaCl 3 M mounted in a Luggin

Cyclic voltammetry was selected to perform the initial electrochemical study of the deposition processes. Firstly, the electrochemical response of the blank solution (DES solvent) was
recorded to establish its electrochemical window (Fig. 1) and to ensure reference electrode reproducibility in the DES medium. Fig. 1
shows the coincidence between voltammetric scans performed
immediate after electrode immersion and those made hours after.
A wide electrochemical window in vitreous carbon electrode was
observed between the reductionoxidation processes of the
solvent.
3.1. Cobalt deposition
A basic study of cobalt deposition was made using the habitual
electrochemical techniques.
Cyclic voltammetric scans were recorded at different negative
limits. By reversing the scan at very low negative overpotential,
nucleation loop was observed (Fig. 2A, curve a). For more negative
potential limits, a well dened reduction peak appeared previous
to a massive reduction current (Fig. 2A, curve b). For all the cathodic limits used, a single oxidation peak was recorded during the
positive scan.
Holding the scan in the potential range of the rst reduction
peak, an increase in the oxidation charge was recorded increasing
the hold time (Fig. 2B, curve b), as corresponds to deposition process. When the hold was made in the potential range of the
massive reduction during the same time, the oxidation charge increase recorded was less important and a widening of the peak
to positive potentials was observed (Fig. 2B, curve c).

Fig. 2. Cyclic voltammograms from CoCl2 0.113 M solution (A) at different cathodic limits, and (B) with potential holds of: (a) t = 0 s, (b) t = 30 s at
1.5 V.

1.25 V and (c) t = 30 s at

20

E. Gmez et al. / Journal of Electroanalytical Chemistry 658 (2011) 1824

Fig. 3. (A) jt transients at 1.2 V from different [Co(II)] solutions: (a) 0.019 M, (b) 0.036 M and (c) 0.113 M, (B) Et transients (a) [Co(II)] = 0.113 M at
the [Co(II)] = 0.036 M solution curves: (b) j = 8 mA cm 2, (c) j = 16 mA cm 2 and (d) j = 32 mA cm 2.

This general behaviour was similar using solutions with different Co(II) concentration, the only difference observed being the advance in the appearance of the reduction current and the increase
of the current recorded as the Co(II) concentration was increased.
For all the solutions tested, the potentiostatic experiments were
made stepping the potential from a potential value at which no
current was detected to different potentials values. All the jt transients recorded evolved to a maximum from which the current

8 mA cm

, and from

prole decays to attain a quasi stationary value. For a xed concentration, the maximum appeared a short deposition times as the potential applied was made more negative. By comparing the jt
curves obtained at xed potential from different cobalt (II) concentrations, a displacement of the maximum to low deposition times
and an increase in the recorded current was observed increasing
the concentration (Fig. 3A). Stirring the solutions, the current decay
after the maximum was reduced.

Fig. 4. (A) Cyclic voltammograms from different Sm(III) solutions: (a) 0.005 M, (b) 0.025 M and (c) 0.045 M, (quiescent conditions). Voltammetric potential holds under
stirred conditions from the [Sm(III)] = 0.025 M solution at: (B) 1.52 V during 60 s and (C) 1.78 and 2.34 V during 60 s.

E. Gmez et al. / Journal of Electroanalytical Chemistry 658 (2011) 1824

Galvanostatic experiments under stirring conditions revealed


that the system initially attained a minimum negative potential value (potential spike) that evolved to less negative values as the
deposition time increased (Fig. 3B, curve a). As the applied current
was made more negative, more negative potentials were associated (Fig. 3B, curves bd).
The shape of both voltammetric and potentiostatic curves and
the effect of the stirring over the voltammetric, potentiostatic
and galvanostatic responses to the Co deposition process in the eutectic 1 choline chloride:2 urea mixture reveals the transport control of the process, as in aqueous solutions [34]. The stirring of the
solution minimises the [Co(II)] depletion near the electrode,
although in this type of solvent vigorous agitation was needed to
appreciate the effect of the stirring arising difcult ion transport.
All the experimental results obtained, nucleation loop, maximum in the jt transient and nucleation spike in the Et transients
evidenced that the characteristic outputs observed in water dissolutions [3537] continue being valid in the DES solvent demonstrating that nucleation and growth process takes place for cobalt
deposition in this eutectic solvent.
3.2. Samarium process
For all the samarium (III) solutions tested, voltammetric experiments revealed similar features in the negative going sweep: at

21

rst a peak-band appeared, followed by a second peak developed


previous to massive current (Fig. 4A). The two reduction peaks
were advanced, enhanced and better dened as the Sm(III) concentration increased. By reversing the scan in none condition nucleation loop was recorded. During the positive scan no oxidation
current was recorded even when very low cathodic limits were
used.
For all Sm(III) concentrations, consecutive cyclic voltammetric
scans evidenced that as the number of scans increases, the onset
of the current appearance occurs at more negative potentials even
when the negative potential limit corresponds to low overpotentials. This is a rather unexpected behaviour. In general, increasing
the number of consecutive scans favours the onset of deposition
process when a previous deposit is present. Therefore, it seems
that in the potential range of the rst peak, the reduction process
is partially blocked for some reduction products.
This result was corroborated by a set of voltammetric experiments holding the scan along the potential range of the rst peak:
both under quiescent and stirred conditions the reduction current
decreased during the hold (Fig. 4B and C). Whereas when the
potential hold was made in the potential range of the second peak
predictable behaviour was recorded: a diminution of current was
observed under stationary conditions but important current increase occurred under stirring conditions (Fig. 4C). In none condition oxidation current was revealed during positive scan.

Fig. 5. (A) jt transients from the [Sm(III)] = 0.025 M solution at different potentials: (a) 1.3 V, (b) 1.4 V, (c) 1.5 V and (d) 1.6 V, the rst part of the curves recorded
under quiescent conditions and from symbol, under stirring. (B) Et transients from 0.045 M Sm(III) at j = 1.6 mA cm 2 using: (a) a polished substrate at stationary
conditions, (b) a polished substrate under agitation and (c) the substrate after experiment b, at stationary conditions. (C) Et transients from [Sm(III)] = 0.045 M (a)
j = 3.2 mA cm 2 under stirring, (b) j = 6.4 mA cm 2 under stirring, (c) j = 6.4 mA cm 2 quiescent solution, (d) j = 9.5 mA cm 2 under stirring, (e) j = 9.5 mA cm 2
quiescent solution, and (f) solvent response at j = 3.2 mA cm 2.

22

E. Gmez et al. / Journal of Electroanalytical Chemistry 658 (2011) 1824

Fig. 6. SEM images of Samarium deposits obtained during a voltammetric hold potential at 1.45 V under stirred conditions from the solution 0.045 M of Sm(III) after: (A)
15 min, (B) 25 min and (C) 80 min, and (D) SEM image of a deposit obtained at 1.9 V during 30 min under stirred conditions in a potentiostatic experiment.

The jt transients corresponded, in all applied potentials, to a


sudden current increase followed by a current decay. As the potential applied was made more negative the current recorded was
greater. The stirring of the solution lead to an increase in the recorded current but was not sufcient to maintain stationary value
and especially when the most negative potentials were applied, a
smooth current diminution was recorded.
Galvanostatic transients were recorded, in both stationary and
stirring conditions and applying wide current densities spectrum,
thus covering stabilization potentials corresponding to the two
reduction peaks. For currents that evolved to potentials corresponding to the rst voltammetric peak, stirring of the solution
lead the potential to more negative values than that corresponding
to quiescent conditions (Fig. 5B, curves a and b). Moreover, in case
that galvanostatic experiment was repeated on a substrate, which
has been maintained into the solution and thus with the deposit already formed present, the potential value achieved was more negative than the corresponding one to the polished substrate (Fig. 5B,
curve c). Nevertheless when the applied currents lead the system
to potentials corresponding to the second voltammetric peak, the
observed behaviour corresponds to a mass control transfer: the potential value achieved under stationary conditions was more negative than the corresponding under agitation (Fig. 5C). Similar
results were obtained when deoxygenation of solution was made
during 10 h prior to the experiment. This discards the possible effect of the oxygen presence.
In all conditions the potential value at which the solvent reduction takes place was very negative compared to the values where
the metallic cation reduced evidencing that in all cases the main
process corresponds to deposit formation (Fig. 5C, curve f).
Deposits prepared under agitation at different conditions were
analysed by scanning electron microscopy. The imaging of deposits
obtained after a voltammetric hold at potential values corresponding to the rst voltammetric peak, revealed that: low deposition
times lead to very ne grained deposits (Fig. 6A). Increasing deposition time the coverage increased and the deposit showed some
cracks and a new growth was observed on the rst deposit
(Fig. 6B). At long deposition times the deposit was cracked
(Fig. 6C). In these experiments low current were owed and the

time consuming to achieve sufcient coverage was important.


The deposits obtained potentiostatically in this zone present similar morphology.
Whereas if the potential applied corresponds to more negative
potential, the coating showed different morphology in which at
deposits were observed (Fig. 6D).
3.3. Samariumcobalt deposition
The effect of cobalt (II) in the samarium (III) deposition process
was studied using a solution containing 0.045 M Sm(III) + 0.018 M
Co(II).
Voltammetric scan evidenced a slight advance of the overall
deposition process (Fig. 7A, curve a) respect to that observed in
the free cobalt (II) solution (Fig. 7A, curve b). Scanning towards
negative potential a peak was detected followed by a close peak
that did not clearly developed since a current increase was observed. Reversing the scan at the current onset a nucleation loop
was recorded followed by an oxidation peak (Fig. 7A, inset as curve
c), it seems that this feature corresponds to cobalt deposition.
Enlarging the cathodic limit to more negative limits, it was not observed in any case the appearance of nucleation loop. Low or nor
oxidation current was detected in the positive scan. Independently
of the potential, with the stirring of the solution the charge involved in the reduction process was enlarged (Fig. 7B).
Galvanostatic experiments were made applying currents sufciently negative to allow codeposition of both metals. At quiescent
conditions no clear nucleation spikes were observed, although the
shape of the curve remembers the typical nucleation feature and
with the time the potential dropped smoothly to more negative
values. However when the solution was stirred the potential increased with the increase of deposition time making in evidence
the favourable stirring effect over the deposition process. The agitation effect was also stated in the potentiostatic curves, positive
difference in the charge recorded under agitation was observed
(Fig. 8A). In all potential range where codeposition occurs, the jt
transients recorded showed a previous current peak followed by
a smooth current increase; more evident when the solution was
stirred (Fig. 8A, curves c and d).

E. Gmez et al. / Journal of Electroanalytical Chemistry 658 (2011) 1824

23

Fig. 7. (A) Comparison of cyclic voltammograms obtained from 0.045 M Sm(III) solution containing: (a) [Co(II)] = 0.018 M and (b) [Co(II)] = 0 M. Inset of the gure shows the
voltammogram recorded reversing the scan at the initio of the deposition and (c) [Co(II)] = 0.018 M. (B) Voltammograms obtained from the 0.045 M Sm(III) + 0.018 M Co(II)
solution under: (a) quiescent and (b) stirred conditions.

Deposits prepared at different potentials under stirred conditions were imaged and their samarium percentage was evaluated.
Fig. 8B shows a detail of the morphology of SmCo deposit obtained
in a potential near to the minimal necessary to induce codeposition. This deposit of 27 wt.% of Sm shows nodular morphology with
rounded grains. Low deposition rate was obtained in this case, after
one hour of deposition no complete coverage was allowed. When
the potential was made more negative, ne grained deposits with

homogeneous coverage were obtained (Fig. 8C). Increasing the


deposition time a new growth over the rst deposit was developed
(Fig. 8D). These deposits contain 75 wt.% of Sm independently of
the thickness.
In order to check the possibilities to deposit on a metallic substrate a Ni/Cu/Au slide was used. By applying similar, or even lower, deposition potentials than those over vitreous carbon, the
composition of the deposits corresponded always to 80 wt.% of Sm.

Fig. 8. (A) jt transients from 0.045 M Sm(III) + 0.018 M Co(II) solution under quiescent conditions at: (a) 1.8 V, (b) 1.9 V, and under stirred conditions at (c) 1.8 V and (d)
1.9 V and SEM pictures of deposits obtained over vitreous carbon at: (B) E = 1.3 V during 3600 s, (C) E = 1.6 V during 900 s and (D) E = 1.6 V during 2500 s.

24

E. Gmez et al. / Journal of Electroanalytical Chemistry 658 (2011) 1824

4. Conclusions
This study demonstrates the DES capability to permit the electrodeposition, in non aggressive conditions, of alloys containing
some transition metal. Cobalt deposition analysis in this DES solvent reveals that cobalt deposits through a typical nucleation and
three dimensional process. All the features related to this process
are observed: loop in voltammetric experiments, maximum in
potentiostatic experiments and the typical spike in the galvanostatic ones. This result conrms also that the nucleation-growth
models developed to describe the metal deposition processes could
be applied in this kind of solvents.
However, samarium deposition in DES solvent shows a less
standard behaviour, two clear reduction regions with diverse electrochemical behaviour being detected. At low overpotentials, a
clearly less conductive coverage than the vitreous carbon used as
substrate is formed during the reduction process. When the reduction progresses the recorded current diminishes, even under stirring conditions. Overcoming this potential range, samarium
deposition takes place as expected because the process is now
not inhibited and solution stirring leads to an increase of the current density. Then, during samarium deposition in this DES solvent,
intermediate samarium species formation can occur that lowers
the conductive character of the electrode. However, the morphology and the composition of the lms prepared at low deposition
potentials do not give any insight into such intermediate Sm species nature, since they do not entirely impede samarium deposit
formation. Work is in progress in order to elucidate the nature of
this lm.
When both cobalt and samarium salts are present in the 1 choline chloride:2 urea eutectic mixture, SmCo deposits can be formed
because the potential deposition range of the alloy is included in
the electrochemical window of the solvent.
From the Sm(III):Co(II) ratio used (around 4:1), SmCo deposits
with Sm percentages that exceed 70 wt.% of samarium are obtained, when metallic substrate is used or a sufciently negative
potential is applied over vitreous carbon. This percentage is higher
than that the corresponding to SmCo5 or Sm2Co17 magnetic alloys.
To decrease the samarium percentage in the deposits, experiments
at less negative potential were made, but then the time necessary
for a deposit to form was so inordinately long as to make the process economically unviable. Moreover, at these low potentials, on
metallic substrate, the deposits obtained contain high samarium
percentage. Having demonstrated the possibility of obtaining
CoSm deposits using a DES solvent, ulterior work will be carried
out to electrodeposit SmCo lms of lower Sm percentages which
could be useful as hard magnetic material.
Although the agitation effect is lesser than the observed in parallel studies at similar Co(II) concentrations in aqueous solutions,
the observed behaviour make in evidence that the transport in this
kind of systems is another parameter to consider.
Acknowledgements
This paper was supported by contract CTQ2010-20726 (subprogram BQU) from the Comisin Interministerial de Ciencia y

Tecnologa (CICYT). The authors wish to thank the Serveis Cienticotcnics (Universitat de Barcelona) for the use of their
equipment.
References
[1] Dawei Wei, Masazumi Okido, Curr. Top. Electrochem. 5 (1997) 2136.
[2] A. Etenko, T. McKechine, A. Shchetkovskiy, A. Smirnov, ECS Trans. 3 (14) (2007)
151157.
[3] M. Gibilaro, L. Massot, P. Chamelot, P. Taxil, Electrochim. Acta 54 (22) (2009)
53005306.
[4] E. Boland, A. Shchetkovskiy, A. Smirnov, Proc. Electrochem. Soc. 19 (2002) 97
802.
[5] D.J. Fray, G.Z. Chen, Mater. Sci. Tecnol. 20 (2004) 295.
[6] A. Lisenkov, M.L. Zheludkevich, M.G.S. Ferreiro, Electrochem. Comm. 12 (6)
(2010) 729732.
[7] Fei Xiao, Fagiong Zhao, Deping Mei, Zhirong Mo, Baizhao Zeng, Biosens.
Bioelectron. 24 (2009) 34813486.
[8] Jing-Fang Huang, I-Wen Sun, Electrochim. Acta 49 (19) (2004) 3251.
[9] A.P. Abbott, K.J. McKenzie, Phys. Chem. Chem. Phys. 8 (2006) 4265.
[10] S. Zein El Abedin, A.Y. Saad, H.K. Farag, N. Bousenko, Q.X. Liu, F. Endress,
Electrochim. Acta 52 (8) (2007) 2746.
[11] Taku Oyama, Takeyoshi Okajima, Takeo Onsaka, J. Electrochem. Soc. 154 (6)
(2007) D322.
[12] F. Endres, D. MacFarlane, A. Abbott, Electrodeposition from Ionic Liquids,
Wiley-VCH Verlag GmbH, Weinheim, Germany, 2008.
[13] Jinwey Tang, Kazuhisa Azumi, Electrochim. Acta 56 (3) (2011) 1130.
[14] Hsin-Yi Huang, Chung-Jui Su, Chai-Liu Kao, Po-Yu Chen, J. Electroanal. Chem.
650 (1) (2010) 1.
[15] T. Welton, Chem. Rev. 99 (1999) 2071.
[16] J.S. Wilkes, M.J. Zaworotko, Chem. Commun. (1992) 965.
[17] J. Fuller, R.T. Carlin, R.A. Osteryoung, J. Electrochem. Soc. 144 (1999) 3881.
[18] A.P. Abbott, G. Capper, D.L. Davies, H. Munro, R. Rasheed, Inorg. Chem. 43
(2004) 3447.
[19] J.Z. Yang, W.G. Xu, P. Tian, L.L. He, Fluid Phase Equilib. 204 (2003) 295.
[20] A.P. Abbott, G. Capper, D.L. Davies, R. Rasheed, V. Tambyrajah, Chem. Commun.
(2003) 70.
[21] A.P. Abbott, G. Capper, B.G. Swain, D.A. Wheeler, Trans. Inst. Met. Finish. 83
(2005) 51.
[22] A.P. Abbott, G. Capper, K.J. McKenzie, K.S. Ryder, J. Electroanal. Chem. 599 (2)
(2007) 288.
[23] A. Florea, L. Anicai, S. Costonci, F. Golgovici, T. Visan, Surf. Interface Anal. 42
(2010) 1271.
[24] A.P. Abbott, D. Bothby, G. Capper, D.L. Davies, R. Rasheed, J. Am. Chem. Soc. 126
(2004) 9142.
[25] Tetsuya Tsuda, Laure E. Boyd, Susumu Kuwabata, Charles L. Hussey, J.
Electrochem. Soc. 157 (8) (2010) F96.
[26] G. Pattanaik, D.M. Kirkwood, X. Xu, G. Zangari, Electrochim. Acta 52 (2007)
27552764.
[27] N.M. Dempsey, P. de Rango (Eds.), in: Proceedings of the 18th Int. Workshop
on High Performance Magnets and their Applications, vol. 2, Section IX
MAGMAS Materials, Annecy, France, 2004.
[28] R. Hasegawa, Physica B 299 (2001) 199.
[29] Y.Q. Guo, W. Li, J. Luo, W.C. Feng, J.K. Liang, J. Magn. Magn. Mater. 303 (2006)
e367.
[30] O. Guteisch, K.-H. Mller, K. Khlopkov, M. Wolf, A. Yan, R. Schfer, T.
Gemming, L. Schultz, Acta Mater. 54 (2006) 997.
[31] A. Walther, D. Givord, N.M. Dempsey, K. Khlopkov, O. Guteisch, J. Appl. Phys.
103 (2008) 043911.
[32] J.C. Wei, M. Schwartz, K. Nobe, ECS Trans. 1 (4) (2006) 273.
[33] P. Liu, Y. Du, Q. Yang, Y. Tong, G.A. Hope, J. Electrochem. Soc. 153 (1) (2006)
C57.
[34] E. Gmez, M. Marn, F. Sanz, E. Valls, J. Electroanal. Chem. 422 (1997) 139.
[35] J.A. Harrison, H.R. Thirsk, in: A.J. Bard (Ed.), Electroanalytical Chemistry, vol. 5,
Marcel Dekker, New York, 1971, p. 67.
[36] S. Fletcher, C.S. Halliday, D. Gates, M. Westcott, T. Lwin, G. Nelson, J.
Electroanal. Chem. 159 (1983) 267.
[37] A. Milchev, M. Irene Montenegro, J. Electroanal. Chem. 333 (1992) 93.

You might also like