Você está na página 1de 5

Electrochimica Acta 115 (2014) 499–503

Contents lists available at ScienceDirect

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

Electrodeposition of zinc-cobalt alloys from choline chloride–urea


ionic liquid
Qingwei Chu a,b , Jun Liang a,∗ , Jingcheng Hao a,1
a
State Key Laboratory of Solid Lubrication, Lanzhou Institute of Chemical Physics, Chinese Academy of Sciences, Lanzhou 730000, PR China
b
Graduate University of Chinese Academy of Sciences, Beijing 100039, PR China

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

Article history: The electrodeposition behavior of zinc-cobalt (Zn-Co) alloy was investigated in choline chloride/urea (1:2
Received 24 August 2013 molar ratio) deep eutectic solvent containing 0.11 M ZnCl2 and 0.01 M CoCl2 . Cyclic voltammetry revealed
Received in revised form 30 October 2013 that Co reduced preferably with respect to Zn and anomalous codeposition of Zn-Co did not occur in this
Accepted 30 October 2013
solvent. Chronoamperometric investigations combined with field emission scanning electron microscopy
Available online 11 November 2013
(FE-SEM) indicated that the electrodeposition of Zn-Co alloys followed the mechanism of instantaneous
nucleation. Energy dispersive spectroscopy (EDS), grazing incidence X-ray diffraction (GI-XRD) and SEM
Keywords:
results showed that the deposition potential influenced the compositions, phase structure and surface
Electrodeposition
Ionic liquid morphology of the Zn-Co alloys.
Zn-Co alloy © 2013 Elsevier Ltd. All rights reserved.
Codeposition mechanism.

1. Introduction the problems associated with hydroxide generation and hydrogen


evolution that often occur in aqueous baths can in principle be
In recent years, great interest has been shown in the possi- eliminated during the electrodeposition process [15].
bilities offered by the electrodeposition of alloys, because they Ionic liquids (ILs) are ionic materials that have a melting point
can improve the mechanical and chemical properties of metals below 100◦ C. Ionic liquid is a promising electrolyte for electrodepo-
[1]. Electrodeposition of Zn-Co alloy coatings has drawn a lot of sition of various metals, since it possesses a number of interesting
attention owing to their better anti-corrosion properties than and advantageous characteristics, such as wide electrochemical
pure Zn deposits [2–5]. Furthermore, other properties such as windows, wide working temperature range, high electrical conduc-
ductility, hardness and weldability of zinc are also improved. The tivity, and good thermal and chemical stability [16,17]. Ionic liquids
electrodeposition of Zn-Co alloys has been widely practiced in have generated significant interest in electrochemical studies due
aqueous plating systems [3–8]. As characterized by Brenner, the to their remarkable properties.
co-deposition behavior of Zn-Co in aqueous solution is anomalous The electrodeposition of zinc-cobalt alloys has been investi-
process, in which less noble Zn deposits preferably with respect gated in a lewis acidic zinc chloride-1-ethyl-3-methylimidazolium
to the more noble Co, leading to the Zn/Co ratio in the deposit is chloride ionic liquids containing cobalt (II) [18]. It was found that
higher than that in the solution [9]. This anomalous process can the reduction of cobalt in this ionic liquid occurred at a poten-
be explained by the so-called ‘hydroxide suppression mechanism’ tial more positive than the reduction of zinc and the deposits can
[10–12], i.e., a zinc hydroxide layer is formed on the electrode be varied from Co-rich to Zn-rich by increasing deposition over-
surface, which can suppress cobalt reduction [13]. In addition, it potential. However, the hygroscopic nature of this ionic liquid
would be unavoidable to evolve hydrogen during the electrode- has delayed progress in their practical application [19,20]. Abbott
position process in an aqueous solution electrolyte. Consequently, and co-workers introduced a relatively new class of ionic liquid,
the quality of the deposited layers is generally unsatisfactory [14]. deep eutectic solvent (DES), which is based on eutectic mixtures
Non-aqueous solvents, for example, ionic liquids, provide an ideal of choline chloride (ChCl) with a hydrogen bond donor species
alternative for the electrodeposition of Zn-Co alloys in a practical [21]. Eutectic solvent is a powerful and potential media for the
sense. Owing to the absence of water in the non-aqueous baths, electrodeposition of metals. They are stable in air and water, and
thus electrodeposition is possible in a room atmosphere at ambi-
ent temperatures [22]. Wang et al. [14] prepared Zn-Co alloy from
∗ Corresponding author. Tel.: +86 931 4968381; fax: +86 931 4968163. ChCl-urea ionic liquid and investigated the cathode potential, con-
E-mail addresses: jliang@licp.cas.cn (J. Liang), jhao@sdu.edu.cn (J. Hao). centration ratio of Zn2+ /Co2+ , temperature and deposition time
1
Tel.: +86 531 88366074; fax: +86 531 88564464. on the deposition behavior. However, the nucleation and growth

0013-4686/$ – see front matter © 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.electacta.2013.10.204
500 Q. Chu et al. / Electrochimica Acta 115 (2014) 499–503

mechanism of Zn-Co alloy from ChCl-Urea ionic liquid has not been
systematically investigated and well understood.
In the present study, we explored the electrochemical depo-
sition of Zn-Co alloys from a 1:2 ChCl–urea eutectic-based ionic
liquid. Cyclic voltammetry and chronoamperometry were used
to investigate the electrochemical mechanism and the nucle-
ation/growth process of Zn-Co alloys from the ionic liquid. Zn-Co
alloys electrodeposited by potentiostatic methods on Cu substrates
were characterized. The effect of deposition potential on the com-
position, phase structure and morphology of the deposits was
examined.

2. Experimental

Choline chloride (ChCl) [HOC2 H4 N(CH3 )3 Cl, AR, ≥99.0%], urea


[CON2 H4 , AR, ≥99.0%] and zinc chloride [ZnCl2 , AR, ≥98.0%] were
used as received. Cobalt chloride [CoCl2 ·6H2 O, AR, ≥99.0%] was
dried at 80◦ C for 24 h under vacuum. The ChCl-based ionic liquid Fig. 1. Cyclic voltammograms for Pt disk electrode in 1ChCl/2urea ionic liquid con-
was prepared by mixing the ChCl and urea with a molar ratio of 1:2 taining 0.11 M ZnCl2 + 0.01 M CoCl2 at scan rate of 50 mV/s under 80 ◦ C.
in a beaker at 80◦ C until a homogeneous, colorless liquid formed.
Then 0.11 M ZnCl2 and 0.01 M CoCl2 was added to the liquid with
gentle stirring to obtain navy-blue solutions. in the 2␪=30◦ -60◦ range with a scanning rate of 15◦ /min and a step
Cyclic voltammetry (CV) and chronoamperometry (CA) were of 0.02◦ .
carried out using an Autolab electrochemical workstation. Elec-
trochemical experiments were performed using a three-electrode 3. Results and discussion
cell. A platinum disk (1.0 mm diameter) was assembled as working
electrode, and a zinc plate (10mm × 10 mm) was used as counter 3.1. Voltammetric studies
electrode and reference electrode. The working electrode was pol-
ished with 0.3 mm alumina paste, cleaned in an ultrasonic acetone Typical cyclic voltammograms in 1:2 ChCl–urea ionic liquid con-
bath, rinsed with distilled water and dried before each measure- taining 0.11 M ZnCl2 and 0.01 M CoCl2 on a Pt electrode at 80 ◦ C are
ment. The cyclic voltammetry experiment was performed at 80◦ C shown in Fig. 1. There are two reduction peaks on the cathodic
and scan rates of 50 mV/s. branch of the voltammogram. The small peak at 0.06 V (noted as
The deposition of Zn-Co alloy was carried out potentiostatically “a”) is attributed to the Co2+ to Co reduction process, while the evi-
on an Autolab electrochemical workstation. The working elec- dent reduction peak at–0.37 V (noted as “b”) can be assigned to the
trode was a Cu plate with a working surface area of 126 mm2 Zn2+ to Zn reduction process. EDS analysis shows that the deposit
(7mm × 18 mm). The counter electrode and reference electrode potentiostatically plating at 0.06 V (reduction peak “a”) for 20 min
were Zn plate (100 mm2 , 10mm × 10 mm). All the electrodes were on Cu plate is composed of 36.1wt% Zn and 63.9wt% Co (Fig. 2a).
rinsed with acetone, cleaned with distilled water and then dried in Zn-Co codeposition is achieved through the deposition of Zn upon
air before all measurements. The electrodeposition of Zn-Co alloy (deposited) Co at a potential more negative than that of Co deposi-
was performed using a constant potential mode at 80◦ C for 20 min tion. This necessarily results in Zn content lower than that of Co. The
under stirring. deposit potentiostatically plating at −0.37 V (reduction peak “b”)
A field emission scanning electron microscope (FESEM, JSM- for 20 min on Cu plate is composed of 93.5wt% Zn and 6.50wt% Co
6701F) and a scanning electron microscopy (SEM, JSM-5600LV) (Fig. 2b). Combined with the result of CV and component analysis,
were employed for the observations of the deposits morphology. it can be found that the electrodeposition of Zn-Co in 1:2 ChCl–urea
Elemental compositions of the Zn-Co deposits were observed by solution is not an anomalous codeposition, i.e. more noble metal Co
energy dispersive spectroscopy (EDS, Kevex). Phase structure of the depositing preferentially with respect to the less noble one Zn.
deposits was determined by Rigaku D/MAX2400 grazing incidence As suggested by Gómez and Vallés [6], the detection of multiple
X-ray diffraction (GI-XRD, incident angle 2◦ ) using Cu-K␣ radiation peaks during the electrochemical oxidation of Zn-Co alloys can be

Fig. 2. EDS spectra of Zn-Co deposits obtained by potentiostatic deposition from 1ChCl/2urea ionic liquid containing 0.11 M ZnCl2 + 0.01 M CoCl2 at 0.06 V (a) and -0.37 V (b).
Q. Chu et al. / Electrochimica Acta 115 (2014) 499–503 501

Fig. 3. Linear sweep voltammetry for Pt disk electrode in 1ChCl/2urea ionic liquid Fig. 4. Current-time transients resulting from chronoamperometric experiments
containing 0.11 M ZnCl2 (dotted line); 0.01 M CoCl2 (solid line). that were performed at a Pt electrode in 1ChCl/2urea ionic liquid containing 0.11 M
ZnCl2 and 0.01 M CoCl2 at different applied potentials from -0.3 V to -0.5 V at 80 ◦ C.

attributed to the dissolution of the metals in the alloy via different


cobalt contents. Thus, the information regarding the characteristics
of the components of the alloy and the structure of the deposited
phases can be got from the voltammetric response. When the scan
reversed, three oxidation peaks (I, II and III), as shown in Fig. 1, are
observed in the potential range of 0.32 to 0.68 V. The dissolution
peaks for pure Zn and pure Co are at 0.24 V and 0.82 V, respectively
(Fig. 3), and dissolution peaks for Zn–Co deposits are somewhere
in between those two potentials. Based on the results, as well as
on the data reported by other authors for the Zn–Co alloys having
similar characteristics [4,23], the three anodic dissolution peaks
correspond to the dissolution of the constituents of two phases,
i.e, Zn-rich and Co-rich Zn-Co alloy. The anodic peaks (I) and (II)
correspond to the dissolution of zinc from different phases of Zn–Co
alloys. The peak (III) corresponds to the dissolution of Co from their
phases.
Chronoamperometry is widely used to study nucleation
phenomena in aqueous or molten media [24]. Deposition of met-
als/alloys onto a foreign substrate generally involves some type Fig. 5. Comparison of experimental data obtained from chronocurrent transients
with instantaneous and progressive nucleation models in 1ChCl/2urea ionic liquid
of three dimensional nucleation process accompanied with hemi-
containing 0.11 M ZnCl2 and 0.01 M CoCl2 .
spherical growth of the developing nuclei. Three-dimensional
nucleation/growth can be described as either ‘instantaneous’ or
‘progressive’ [19]. Instantaneous nucleation tends to lead to for- and growth of Zn/Co nuclei, until a peak appeared. This rise in cur-
mation of an alloy phase (a solid solution), whereas progressive rent culminates in a broad maximum, im , at the position of which on
nucleation to a phase which has a small amount of one species the time axis, tm , depends upon the magnitude of the potential step.
inserted in the matrix of the other major species [25]. Finally the current converge to the limiting current corresponding
Current transients of chronoamperometric experiments for var- to linear diffusion to a planar electrode as per the Cottrell’s law. The
ious potentials are shown in Fig. 4. The initial regime of each im increases while the tm shortens with an increase in the applied
transient is characterized by a sharp decrease in current which potential. This can be explained by the decrease in the time required
corresponds to the formation of the first nuclei on the electrode. for the diffusion layer to overlap, owing to an increased nucleation
Then the current increases which is attributed to the formation density [19].

Fig. 6. SEM micrographs of the Zn-Co alloy deposited from 1ChCl/2urea ionic liquid containing 0.11 M ZnCl2 and 0.01 M CoCl2 at (a) -0.3 V, (b) -0.40 V and (c) -0.50 V for 10s.
502 Q. Chu et al. / Electrochimica Acta 115 (2014) 499–503

electrolysis, and progressive (Eq. (2)) nucleation, that new crystals


are continuously created throughout electrolysis, are represented,
by
(1)(i/im )2 = 1.9542 (t/tm )−1 {1 - exp[− 1.2564(t/tm )]} 2
2 −1 2 2
(2)(i/im ) = 1.2254 (t/tm ) {1 − exp[−2.3367(t/tm ) ]} where
im and tm are the current and the time, as respective peak
coordinates.
Fig. 5 shows the experimental current transients plotted in
current-time coordinates, along with the lines for instantaneous
and progressive nucleation (black and blue lines), described by
Eq. (1) and (2). It is clear that the Zn-Co nucleation follows the
instantaneous three-dimensional nucleation. Combined with the
cyclic voltammetry results, it reveals that the Zn-Co codeposition
produces solid solutions in ChCl-Urea ionic liquid.
The SEM micrographs of the Zn-Co alloy deposited on the Cu
electrode surface at short depositing time are used to further
Fig. 7. Composition of the electrodeposited Zn-Co alloys, determined with EDS, as a
function of the deposition potential from 1ChCl/2urea ionic liquid 0.11 M ZnCl2 and
investigate the nucleation mechanism (Fig. 6). The presence of
0.01 M CoCl2 at 80 ◦ C. growth centers with similar sizes can be observed at different
applied potential, which is consistent with instantaneous three-
Scharifker and Hills proposed a convenient method to identify dimensional nucleation.
the nucleation mechanism [26]. The most commonly used models
for describing the current-time transient are progressive and 3.2. Characterization of deposits
instantaneous nucleation models. The theoretical transients for
limiting cases of instantaneous (Eq. (1)), in which all the Zn/Co The composition and microstructure of the Zn-Co alloy elec-
nuclei are created at the same moment at the beginning of the trodeposited onto copper plate from 1:2 ChCl–urea ionic liquid

Fig. 8. XRD full pattern of the Zn-Co deposits plated at different potentials for 20 min (a) and XRD fitting patterns of Zn-Co deposits plated at -0.3 V (b), -0.4 V (c) and -0.5 V
(d) for 20 min.
Q. Chu et al. / Electrochimica Acta 115 (2014) 499–503 503

Fig. 9. SEM micrographs of the Zn-Co deposits obtained from 1:2 ChCl-urea ionic liquid containing 0.11 M ZnCl2 and 0.01 M CoCl2 at (a) -0.3 V, (b) -0.4 V and (c) -0.5 V for
20 min.

containing 0.11 M ZnCl2 and 0.01 M CoCl2 using different potentials codeposition and the nucleation mechanism of Zn-Co alloy is an
for 20 min were characterized. instantaneous process. EDS and XRD studies clearly demonstrate
An abridged summary of the composition of the electrode- that the compositions and phase structure of the Zn-Co alloys are
posited Zn-Co alloy, determined with EDS, as a function of the significantly dependent on the deposition potentials. It is found
deposition potential is shown in Fig. 7. It can be seen that the Co that the Co content in the deposits and the ␥-phase of Zn–Co alloys
content decreases when the deposition potential becomes more decrease while the Zn content and the ␩-phase increase with the
negative. This is mainly ascribed to that the Zn deposition rate deposition potential becomes more negative. The morphology of
increases with more negative potential. The composition reference the Zn-Co deposits is also potential dependent and the diameter
line (CRL), representing the percentage of Co in the deposits equals of the grain clusters increases for increasingly negative values of
to that in the bath [10], is also shown in Fig. 7. It is obvious that the potential.
Co content in deposits is always above the CRL, which means that
the codeposition of Zn and Co is a normal type in 1:2 ChCl–urea Acknowledgments
ionic liquid, consistent with the results of the voltammetric study.
Fig. 8 shows the GI-XRD patterns of Zn-Co alloy electrodeposited The financial support from National Natural Science Foundation
plated onto copper plate at different potential for 20 min. It can be of China (Grant No. 51305432) and the “Hundred Talents Program”
seen from Fig. 8a that two diffraction peaks, viz., a broad peak at 2␪ of Chinese Academy of Sciences (J. Liang) was gratefully acknowl-
in the range between 41 to 44◦ and a peak at 2␪=50◦ , are observed edged.
for all deposits. The peak at 2␪=50◦ is the diffraction peak from
copper substrate. The broad diffraction peak at 2␪ in the range References
of 41 to 44◦ can be fitted to a Zn-rich ␩-phase and a cubic Zn-Co
␥-phase by a profile-fitting procedure as shown in Fig. 8b-d. Consid- [1] S. Basavanna, Y. Arthoba Naik, J. Appl. Electrochem. 39 (2009) 1975–1982.
[2] C.N. Panagopoulos, D.A. Lagaris, P.C. Vatista, Mater. Chem. Phys. 126 (2011)
ering the fitting results, an increase of the ␩-phase and decrease of 398–403.
the ␥-phase are observed when the potential becomes more neg- [3] E. Gómez, X. Alcobe, E. Vallés, J. Electroanal. Chem. 505 (2001) 54–61.
ative. The increase of the ␩-phase means that the Co content in [4] J.B. Bajat, S. Stanković, B.M. Jokić, S.I. Stevanović, Surf. Coat. Tech. 204 (2010)
2745–2753.
the deposit decreases with the potential becoming more negative.
[5] J.B. Bajat, V.B. Miskovic-Stankovic, M.D. Maksimovic, D.M. Drazic, S. Zec, Elec-
This observation is in good agreement with the aforementioned trochim, Acta 47 (2002) 4101–4112.
EDS analysis. [6] E. Gómez, E. Vallés, J. Electroanal. Chem. 397 (1995) 177–184.
The surface morphologies of Zn-Co alloys electrodeposited in [7] M. Mouanga, L. Ricq, P. Berçot, J. Appl. Electrochem. 38 (2007) 231–238.
[8] M. Mouanga, L. Ricq, P. Bercot, Surf. Coat. Tech. 202 (2008) 1645–1651.
the 0.11 M ZnCl2 and 0.01 M CoCl2 plating bath for 20 min at the [9] A. Brenner, Electrodeposition of Alloys, Principles and Practice, Academic Press,
potential of −0.3 V, -0.4 V and −0.5 V are shown in Fig. 9. The Zn-Co New York, 1963, pp. 194, Ch. 30.
deposit obtained from −0.3 V is made up of spherical clusters with [10] K. Higashi, H. Fukushima, T. Urakawa, J. Electrochem. Soc. 128 (1981) 2081.
[11] T. Tsuru, S. Kobayashi, T. Akyama, H. Fukushima, S.K. Gogia, R. Kammel, J. Appl.
diameter ranging from 1 to 2 ␮m (Fig. 9a). The clusters distribute Electrochem. 27 (1997) 209–214.
independently each other, making the surface of the deposit highly [12] Z.F. Lodhi, J.M.C. Mol, W.J. Hamer, H.A. Terryn, J.H.W. De Wit, Electrochim. Acta
porous. The grain clusters with size of 2 to 3 ␮m appear on the 52 (2007) 5444–5452.
[13] J.L. Ortiz-Aparicio, Y. Meas, G. Trejo, R. Ortega, T.W. Chapman, E. Chainet, P. Ozil,
coating surface at the potential of −0.40 V (Fig. 9b). With further Electrochim. Acta 52 (2007) 4742–4751.
increasing the potential to -0.5 V, the diameter of the grain clusters [14] X. Wang, H. Li, J. Wang, X. Fu, J. Gu, Mater. Protection 43 (2010) 30–33.
increases to 3 to 4 ␮m (Fig. 9c). With the potential increasing, the [15] M.Z. An, P.X. Yang, C.N. Su, A. Nishikata, T. Tsuru, Chin. J. Chem. 26 (2008)
1219–1224.
growth rate increases and thus the diameter of the grain clusters [16] C.N. Su, M.Z. An, P.X. Yang, H.W. Gu, X.H. Guo, Appl. Surf. Sci. 256 (2010)
increases. 4888–4893.
Based on the EDS, XRD and SEM results, it is clear that the [17] C.D. Gu, Y.H. You, Y.L. Yu, S.X. Qu, J.P. Tu, Surf. Coat. Tech. 205 (2011)
4928–4933.
composition, phase structure and surface morphology of the Zn-
[18] P.Y. Chen, I.W. Sun, Electrochim. Acta 46 (2001) 1169–1177.
Co alloys prepared in ChCl-Urea ionic liquid can be controlled by [19] H.Y. Yang, X.W. Guo, X.B. Chen, S.H. Wang, G.H. Wu, W.J. Ding, N. Birbilis,
varying the deposition potential. Electrochim. Acta 63 (2012) 131–138.
[20] F. Endres, Chem. Phys. Chem. 3 (2002) 144–154.
[21] A.P. Abbott, D. Boothby, G. Capper, D.L. Davies, R.K. Rasheed, J. Am. Chem. Soc.
4. Conclusions 126 (2004) 9142–9147.
[22] A. Bakkar, V. Neubert, Electrochem. Commun. 9 (2007) 2428–2435.
The electrodeposition behavior of Zn-Co alloys from 1:2 [23] G. Trejo, R. Ortega, Y. Meas, E. Chainet, P. Ozil, J. Appl. Electrochem. 33 (2003)
373–379.
ChCl–urea ionic liquid containing 0.11 M ZnCl2 and 0.01 M CoCl2 [24] K. Serrano, P. Taxil, J. Appl. Electrochem. 29 (1999) 505–510.
was studied. Cyclic voltammogram and chronoamperometry anal- [25] Y.P. Lin, J.R. Selman, J. Electrochem. Soc. 140 (1993) 1304–1311.
ysis show that the electrodeposition of Zn-Co alloy is a normal [26] B. Scharifker, G. Hills, Electrochim. Acta 28 (1983) 879–889.

Você também pode gostar