You are on page 1of 11

Applied Surface Science 410 (2017) 574–584

Contents lists available at ScienceDirect

Applied Surface Science


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

Full Length Article

Zn–Mn alloy coatings from acidic chloride bath: Effect of deposition


conditions on the Zn–Mn electrodeposition-morphological and
structural characterization
N. Loukil ∗ , M. Feki
Laboratoire de Génie des Matériaux et Environnement (LGME), Université de Sfax, ENIS.B.P.W., 1173-3038, Tunisie

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

Article history: Zn–Mn alloy electrodeposition on steel electrode in chloride bath was investigated using cyclic voltam-
Received 9 December 2016 metric, chronopotentiometric and chronoamperometric techniques. Cyclic voltammetries (CV) reveal a
Received in revised form 3 February 2017 deep understanding of electrochemical behaviors of each metal Zn, Mn, proton discharge and Zn–Mn
Accepted 9 February 2017
co-deposition.
Available online 14 February 2017
The electrochemical results show that with increasing Mn2+ ions concentration in the electrolytic bath,
Mn2+ reduction occurs at lower over-potential leading to an enhancement of Mn content into the Zn–Mn
Keywords:
deposits. A dimensionless graph model was used to analyze the effect of Mn2+ ions concentration on
Electrodeposition
Zn–Mn alloy
Zn–Mn nucleation process. It was found that the nucleation process is not extremely affected by Mn2+
Nucleation concentration. Nevertheless, it significantly depends on the applied potential.
Morphology Several parameters such as Mn2+ ions concentration, current density and stirring were investigated
Structure with regard to the Mn content into the final Zn–Mn coatings. It was found that the Mn content increases
with increasing the applied current density jimp and Mn2+ ions concentration in the electrolytic bath.
However, stirring of the solution decreases the Mn content in the Zn–Mn coatings. The phase structure
and surface morphology of Zn–Mn deposits are characterized by means of X-ray diffraction analysis
and Scanning Electron Microscopy (SEM), respectively. The Zn–Mn deposited at low current density
is tri-phasic and consisting of ␩-Zn, ␨-MnZn13 and hexagonal close packed ␧-Zn–Mn. An increase in
current density leads to a transition from crystalline to amorphous structure, arising from the hydroxide
inclusions in the Zn–Mn coating at high current density.
© 2017 Published by Elsevier B.V.

1. Introduction as Mn is more electronegative than Zn, it dissolves first as Mn2+


ions, causing a slight increase of pH due to oxygen reduction. After
High performance sacrificial electroplated coatings have always its total dissolution from the coating surface, the dissolution pro-
been required in the metal finishing industry. Recently, the auto- cess continues with that of Zn. Because of the higher pH value, Zn2+
motive industry has shown a great interest in alloy coatings Zn–X ions react immediately with the medium and form zinc hydroxide
(X = Fe, Co, Ni, Mn. . .) as an alternative for toxic and high cost chloride ZHC. The same authors [3,4] reported that ZHC inhibits the
cadmium coatings [1]. Compared to pure metal coatings, electrode- activity of the cathodic oxygen reduction, increasing the anticorro-
posited Zn–X alloys have been proved to offer better properties sive properties of Zn–Mn coatings.
depending on the alloying element. Zn–Mn coatings appear to be Several published reports investigated the effect of the Mn
particularly attractive since these deposits exhibit an enhanced cor- content on enhancing the anticorrosive properties of the Zn–Mn
rosion resistance than pure Zn. In fact, Boshkov et al. [2] reported coatings. A Zn–Mn alloy containing approximately 15–40% shows
that the alloying element, Mn, has been proposed to its dual protec- the highest corrosion resistance [5]. Other researchers reported
tive role affording the highest corrosion resistance of these alloys that the optimum behavior of Zn–Mn alloys is achieved for minimal
[2]. Boshkov and co-workers [3,4] explained this dual role of Mn; Mn content of 20% [6].
According to literature, the Mn content co-deposited in the
Zn–Mn alloys can be controlled under several operational param-
eters (solvent choice, bath composition, agitation, current density,
∗ Corresponding author. organic additives, temperature, pH. . .) such that alloy properties
E-mail address: nloukil87@gmail.com (N. Loukil).

http://dx.doi.org/10.1016/j.apsusc.2017.02.075
0169-4332/© 2017 Published by Elsevier B.V.
N. Loukil, M. Feki / Applied Surface Science 410 (2017) 574–584 575

(structure, homogeneity and porosity. . .) change systematically Table 1


Chemical compositions of the electrolytic baths.
[7]. This is mainly due to the induced changes on the mechanism
and kinetics of electrodeposition [8]. Chemical composition Concentration (M)
Regardless of the plating bath type, electrodeposition of Zn–Mn
Bath 0 Bath 1 Bath2
is complicated in view of the relatively large gap between
ZnCl2 – 0.3 0.3
Zn deposition potential (E◦ (Zn2+ /Zn) = −0.76 V/HSE) and Mn (E◦
MnCl2 ·2H2 O – 1 0.3
(Mn2+ /Mn) = −1.18 V/HSE). It should be noted that those two poten- KCl 1.25 1.25 1.25
tials are more negative than that of hydrogen evolution which H3 BO3 0.4 0.4 0.4
remains inevitable. Because of this fact, complexing agents are
usually required to overcome the Zn–Mn electrodeposition prob-
lems. They act by bringing Zn and Mn deposition potentials closer, structure using atomic absorption spectrometry, Scanning Electron
enhancing Mn content in the alloy and limiting hydrogen evolution. Microscopy (SEM) and X-ray diffraction (XRD), respectively.
Zn–Mn alloys are extensively electrodeposited from sulfate
acidic electrolytes containing some complexing agents. Citrate ions 2. Experimental details
[9,10] are namely the most widely used as complexing agent. How-
ever, the deposition of Zn–Mn alloys from these electrolytes shows The chemical compositions of the two used baths are reported in
often harmful drawbacks. The electrolytes with citrate ions have Table 1. Before each experiment, the pH was accurately adjusted to
two major problems [11], namely: (i) bath instability because of 4.5 using dilute hydrochloric acid (HCl) and/or potassium hydrox-
Mn(III)–citrate complex precipitation [12,13] and (ii) low current ide (KOH) solution. The electrolytic baths were prepared with
efficiency [1] in view of the high stability constants of metal com- analytical grade chemicals. The electrochemical measurements
plexes that leads to a higher over-potential for metal reduction were carried out in a conventional three electrode cell, using a
[9,13]. potentiostat autolab 302N coupled to a computer with specific data
Alternative acidic bath based on ammonium chloride has been acquisition software installed. Various electrochemical methods
proposed, since ammonium species favor the co-deposition of Mn. like linear and cyclic sweep voltammetry, chronopotentiometry as
However, ammonium ions generally produce ammonia and nitrite well as chronoamperometry were performed.
which are extensively toxic [11–14]. Furthermore, the removal of The working electrode was a carbon steel electrode with a sur-
metallic ions from the waste streams is difficult. Due to the new face area of 0.8 cm2 . The specimens were cold covered in epoxy
environmental regulations, the use of an acidic aqueous bath devoid resin. Prior to each experiment, the steel substrate surfaces were
of ammonium and citrate ions is compulsory. In this regard, a sim- mechanically polished, degreased in alcohol and then pickled in a
ple KCl electrolyte without strong complexing agent shows several chloride medium.
benefits, compared to other bath types, like high current efficiency The electrodeposition of pure Mn was also studied using the
and good bath stability. Sylla et al. [15] developed an alternative supporting electrolyte (bath 0) containing 1 M of Mn salt. The work-
simple aqueous chloride electrolyte containing ZnCl2, MnCl2 , KCl ing electrode considered for pure Mn electrodeposition was a glassy
and H3 BO3 . Boric acid (H3 BO3 ) has been proposed in view of its carbon electrode. The reference electrode was a saturated Ag/AgCl
good results to inhibit both hydrogen formation and zinc deposi- electrode. All potentials are referred to this electrode. The anode
tion by acting as a buffer and/or adsorbing at the electrode surface was suitably chosen to avoid precipitation in the bath. All experi-
and blocking then the active centers [15]. Hence, the corresponding ments have been conducted at 25 ◦ C.
current efficiency is near 100%. Zn–Mn coatings were electrodeposited galvanostatically under
Moreover, Bucko et al. [16] emphasized that Zn–Mn deposits different current densities for 30 min. The active surfaces of steel
obtained from chloride electrolyte are generally smooth and com- electrodes were of 12 cm2 .
pact in a wider current density range. The overall greater corrosion To determine the Zn and Mn content in each deposit, the latter
stability of Zn–Mn deposits has been shown from acidic chloride was stripped from the working electrode surface in dilute H2 SO4
bath [17]. Owning to these benefits, attention has been focused on solution (1 M) containing 2 g/l hexamethylenetetramine (HMTA)
the chloride bath for Zn–Mn alloy electrodeposition, a few years as dissolution inhibitor for the steel substrate. Then, Zn and Mn
ago. Meanwhile, no data have been reported on the nucleation and concentration in the obtained solution was determined by atomic
growth mechanisms of Zn–Mn alloys electrodeposition. In return, absorption spectrometry (Analytic Jena ZEENIT 700). For each
several published studies have been only focused on the nucle- deposit, the quantitative analysis was realized in triplicate.
ation and growth mechanisms of either Zn [18] or Zn-alloys such The surface morphologies of the electrodeposited Zn–Mn coat-
as Zn–Ni alloys [19]. In this work, an attempt was made to get an ings were observed using JEOL 5410 Scanning Electron Microscope
insight into the initial stages of nucleation and crystal growth pro- (SEM). Crystalline structures were analyzed by X-ray diffractome-
cess of Zn–Mn electrodeposition from a KCl electrolyte. This study ter (XRD), using X-ray D8 Advance Bruker machine with a CuK∞
was performed using chronoamperometric methods to understand radiation. The measurements were performed in the 2 range of
Zn–Mn electrodeposition under controlled potential conditions. 10–90◦ .
The goal of the present work was to investigate Zn–Mn co-
deposition from a simple additives-free chloride bath on carbon 3. Results and discussion
steel. For this, a detailed investigation must be carried out to bet-
ter understand pure Zn, pure Mn and Zn–Mn alloys deposition at 3.1. Effect of anode material on bath stability
different Mn2+ ion concentrations in the electrolytic bath. The elec-
trochemical process was examined by cyclic sweep voltammetry, In the preliminary experiments, the counter electrode was a
chronopotentiometry and chronoamperometry. platinum Pt placed in a single compartment cell. Whatever the bath
The influence of electrodeposition conditions namely, the Mn2+ used (bath 1 or 2), a transition from pink, due to the color of Mn salt,
ions concentration, the applied current density and the stirring to brown color is observed. A precipitation of a brown solid is then
were investigated with the aim to establish a relation between observed into the working solution, as shown in Fig. 1. A qualitative
the deposition conditions and Mn amount in the Zn–Mn deposits. test confirms that this precipitate is a Mn-based precipitate.
The Zn–Mn alloys obtained galvanostatically were character- The brown precipitate exhibits an amorphous structure as
ized by analyzing their chemical composition, morphology and described in the XRD pattern shown in Fig. 2. In this regard, a
576 N. Loukil, M. Feki / Applied Surface Science 410 (2017) 574–584

Fig. 1. Effect of Pt used as a counter electrode on appearance of electrolyte KCl + H3 BO3 containing both Mn and Zn salts: (a) before deposition (b) after deposition.

study was conducted in a compartmentalized cell with the aim oxidation of Mn2+ to MnO2 [23], as shown in the enlarged zone in
to localize the precipitation and devote the phenomenon. The Fig. 3. Moreover, a more significant increase in the anodic current,
anolyte was separated from the catholyte by fine porosity glass attributed to oxygen evolution from apart of 1.6 V vs. Ag/AgCl, for
frit tube. It is thereby shown that the precipitation is occurring at bath 1 compared to that of bath 0 is observed. As already described
the anodic compartment that has been quickly filled with anode in a previous work [24], the presence of Mn2+ catalyzes the oxy-
sludge. According to literature [20], formation of manganese diox- gen evolution rate. That’s why in this work, the pH of the working
ide and evolution of oxygen gas are the main reactions taking place electrolyte (bath 1) decreases from 4.5 to 2.7 after a deposition
at the anode. Most of studies on Mn [21] and Mn alloys electrode- experiment. The fall in the pH is due to the oxidation of water at
position using a platinum counter electrode signposted that this the anode [25].
brown precipitate is originating from Mn oxides, suggesting that As pointed in literature [21,26], the formation of MnO2 depends
the anodic polarization has been sufficient to enable the Mn2+ oxi- on the anode material. The selection of anode material is based on
dation. As a matter of fact, the Pt anode material exhibits a high the rate of Mn2+ oxidation to MnO2 . It is obviously that this oxide
oxygen overvoltage [21]. would decrease the solution electrical conductivity and increase
Previous works [22] presumed that precipitation can be due the cell voltage, because of its high electrical resistance [21]. To
to the hydrolysis of manganese tri-chloride and manganese tetra- prevent this solution contamination, it is compulsory to use a solu-
chloride, which may be occurred at the anode. ble anode in order to preserve the bath stability. For the subsequent
In order to better understand the anodic processes on Pt, anodic experiments, a pure Zn plate was retained as anode material.
potentiodynamic polarizations were performed in different elec-
trolytic baths: the supporting electrolyte KCl + H3 BO3 (bath 0) and 3.2. Cyclic voltammetry studies
the Zn–Mn electrolyte (bath 1). As shown in Fig. 3, the anodic pro-
cess depends on the salts-containing electrolyte. For the supporting 3.2.1. Mn electrodeposition on glassy carbon
electrolyte (bath 0), the anodic peak, appearing at about 1.2 V vs. According to literature regarding pure Mn electrodeposition
Ag/AgCl, is related to chlorine evolution. It is striking to note that [26–30], there are a discrepancy due to the use of different
chlorine gas is continually evolving at the Pt counter electrode as a electrolyte compositions, electrode materials and plating condi-
secondary product during electrodeposition [6,21]. However, this tions. According to Sylla et al. [15], the potential at which Mn
anodic peak is masked in the curve recorded from the bath 1. A
small anodic peak that appears at almost 1 V corresponds to the

50
45
40
35
intensity (a.u)

30
25
20
15
10
5
0
0 10 20 30 40 50 60 70
2θ (degree)

Fig. 2. XRD pattern of the as-obtained brown precipitate during Zn–Mn electrode- Fig. 3. Anodic voltammetric curves recorded from bath 0 and bath 1 on Pt substrate
position using a Pt anode. (scan rate = 5 mV/s).
N. Loukil, M. Feki / Applied Surface Science 410 (2017) 574–584 577

0.3

0.2

0.1

E (V vs. Ag/AgCl)
0

j (A/cm²)
-2.2 -2 -1.8 -1.6 -1.4 -1.2 -1 -0.8 -0.6 -0.4 -0.2 0

-0.1

[Mn2+]= 0 M -0.2

[Mn2+]=0.3 M

[Mn2+]=1M -0.3
Fig. 4. Cyclic voltammogram of Mn deposition from bath 0 containing 1 M of MnCl2
on a glassy carbon electrode (scan rate 100 mV/s). Fig. 5. Typical Cyclic voltammograms performed on steel from electrolytic bath:
1.25 M KCl + 0.4 M H3 BO3 + 0.3 M ZnCl2 , containing different Mn2+ concentrations
0 M; 0.3 M (bath 2) and 1 M (bath 1).
deposition on steel substrate starts could not be detected due to the
concomitant hydrogen evolution in an acidic chloride bath. More-
over, Genasan et al. [6] demonstrated that Mn electrodeposition The current density then stays steady between −1.3 and −1.5 V
study is complicated as Mn is readily oxidized during the anodic vs. Ag/AgCl, at about 42 mA/cm2 . Therefore, the reduction of Zn2+
sweep. On the basis of these observations reported in literature, a ions is limited by diffusion, i.e. controlled by mass transfer [32]. A
cyclic voltammogram, shown in Fig. 4, is recorded on a glassy car- new increase in current density is due to hydrogen evolution on
bon electrode at 100 mV s−1 in order to detect the exact potential at the freshly formed Zn deposit [32].
which Mn2+ ions reduction occurs. It is well known that the over- When Mn2+ ions are present (at 0.3 and 1 M) in the electrolyte
voltage of hydrogen evolution on the glassy carbon is high. Thus, already containing Zn2+ ions, the two corresponding voltammo-
distinction between the reactions of Mn2+ reduction and hydrogen grams exhibit similar features to that of pure Zn. The potential
evolution at low cathodic potential is possible. This voltammo- at which the cathodic peak arises is similar to that obtained dur-
gram (Fig. 4) confirms that Mn deposition takes place at around ing the reduction of pure Zn. This peak is followed by a current
−1.55 V vs. Ag/AgCl. Such occurrence has been reported for pure hump at approximately −1.5 V vs. Ag/AgCl, when [Mn2+ ] = 0.3 M,
Mn deposition on glassy carbon electrode in previous published corresponding to both Mn2+ ions and H2 O reduction.
results dealing with acidic chloride bath [15,30]. In the backward It can obviously be seen in Fig. 5 that increasing Mn2+ ion con-
scan from −2 V vs. Ag/AgCl, a slight anodic peak, corresponding centration from 0.3 to 1 M depolarizes the Mn deposition potential
to Mn dissolution, is detected at −1.33 V vs. Ag/AgCl. The voltam- to −1.5 V vs. Ag/AgCl, 50 mV more positive than that obtained for a
mogram recorded from an acidic chloride bath, in another work, Mn2+ ion concentration of 0.3 M, −1.55 V vs. Ag/AgCl. Taking in the
exhibits a similar dissolution peak at about −1.34 (± 0.01) V [30]. mind the gap between Zn and Mn deposition potentials, increasing
It has to be noted that the magnitude of this anodic peak is small. Mn salt concentration seems to be beneficial to reduce this gap and
This could be attributed to a spontaneous and rapid chemical disso- subsequently to bring Zn and Mn discharging potentials closer.
lution of the deposited Mn which occurs simultaneously with the Increasing Mn2+ concentration until 1 M generates important
electrochemical one during the anodic scan, as it was suggested by current fluctuations below −1.7 V vs. Ag/AgCl, as shown in Fig. 5.
Ewa Rudnik [30]. It is important to note that the easy dissolution This is attributed to an accelerated reactions of both hydrogen
of Mn deposit is due to its weak stability in acidic media [6,15]. evolution via H2 bubbles and manganese reduction [15]. These
fluctuations were also observed by Boudinar et al. [33] in elec-
trodeposition of Mn–Bi. This suggests that a higher Mn2+ ions
3.2.2. Effect of Mn2+ concentration on Zn–Mn electrodeposition concentration into the electrolyte enables its faster reduction at
The Zn–Mn electrodeposition process in presence of various the cathode. This result was in accord with previous results given
Mn2+ concentrations was investigated. Corresponding voltammo- by Xu et al. [26]. As a matter of fact, Xu et al. [26] investigated
grams were recorded for several Mn2+ concentrations, namely 0, Mn salt concentration effect on pure Mn electrodeposition. Xu and
0.3 M (bath 2) and 1 M (bath 1). The potential interval was scanned his co-workers demonstrated, by investigating the partial current
from − 0.6 V to the cathodic direction up to −2 V and the back scan density of Mn2+ reduction, that Mn electrodeposition reaction was
was ended before anodic steel dissolution. The scan rate was fixed significantly accelerated with increasing Mn2+ concentration.
at 20 mV/s. Fig. 5 shows typical voltammogram for Zn deposition Accordingly, a higher Mn percent in Zn–Mn deposits is expected
recorded from Mn2+ ions-free electrolyte ([Mn2+ ] = 0 M, i.e. the sup- to be incorporated. This fact is obeyed to the normal co-deposition
porting electrolyte (bath 0) containing only Zn2+ ions). As it can be [34]. That is to say, Mn amount into Zn–Mn deposits will increase
seen in Fig. 5, the cathodic current increases sharply from −1.03 V with increasing Mn2+ ion concentration in the electrolyte.
vs. Ag/AgCl and gives rise to a cathodic peak at around −1.2 V vs. Regarding the anodic scan, the anodic peaks seen for the two
Ag/AgCl. This peak is related to Zn2+ reduction during the cathodic Mn2+ concentration (Fig. 5) show two shoulders. This suggests that
scan. Previous studies [31] demonstrated that this reduction pro- the film deposited during the cathodic scan does not contain only
cess involves the species ZnCl4 2− , which is reduced to Zn via the the pure Zn phase. The potential range of the anodic peak is between
following reaction: −1.08 V and −0.1 V vs. Ag/AgCl, that is very close to pure Zn disso-
lution potential range (−1.08 V to −0.2 V vs. Ag/AgCl). This suggests
ZnCl4 2− + 2e → Zn + 4 Cl− (1) that the Zn–Mn alloy deposited during the cathodic scan contains
578 N. Loukil, M. Feki / Applied Surface Science 410 (2017) 574–584

t(s)
0 t(s)
0
0 2 4 6 8 10 0 2 4 6 8 10
-0.05
-0.05
-0.1
j(A/cm²) -0.1 -0.15

j(A/cm²)
-0.2
-0.15
-0.25
E=-1.52 V
E= -1.52 V
-0.2 E=-1.58 V -0.3 E= -1.58 V
E=-1.61 V (b)
(a) -0.35 E= -1.61 V
-0.25 E=-1.75 V E= -1.75 V
-0.4

Fig. 6. Current transients for initial stage of Zn–Mn electrodeposition on steel at cathodic potentials ranging between −1.52 and −1.75 V; (a): bath 1 (b): bath 2.

a high percentage of Zn. Meanwhile, it is striking to note that the [40]. The number of nuclei increases during the process.Thus, the
magnitude of the anodic peak recorded from the bath 1 is the small- age of nuclei will be different [41].
est. This denotes that rising Mn2+ concentration in the electrolytic The experimental chronoamperometric data, shown in Fig. 6,
bath decreases the Zn amount in the Zn-Mn alloy deposited dur- are normalized to a non-dimensional plots (j/jmax )2 vs. t/tmax and
ing the cathodic sweep and consequently favors Mn incorporation compared with the theoretical plots derived for instantaneous
in the Zn–Mn coatings. The decrease of anodic peak magnitude and progressive three-dimensional (3D) nucleation/growth models
is also probably due to additional inhibiting effect of Mn2+ ions [38], whose equations are given for instantaneous (2) and progres-
towards Zn reduction. These observations support the above sug- sive nucleation (3):
gestion according to which Zn–Mn deposits richer in Mn may be  j 2 1.9542
   t 2
obtained as Mn2+ concentration in the electrolytic bath increases. = 1 − exp −1.2564 (2)
jmax t tmax
On switching the potential scan to the anodic direction from tmax
−2 V, the absolute value of the current density is greater than that  j 2    t 2  2
observed during the potential scan in the cathodic direction, lead- 1.2254
= t
1 − exp −2.3367 (3)
ing to two crossover: a nucleation overpotential (E) at the cathodic jmax tmax
tmax
region and a crossover potential at zero current region (Eco), which
is equal to the reversible metal/ion potential [35]. This is character- The theoretical (instantaneous and progressive) and experimen-
istic of three dimensional (3D) nucleation and subsequent crystal tal results for Zn–Mn electrodeposition from bath 2 and bath 1 are
growth process [36]. presented in Figs. 7 and 8, respectively. Fig. 7 shows the nucleation
mechanism of Zn–Mn alloy deposition from bath 2, containing the
lowest Mn2+ ions. The two curves recorded at potentials of −1.52
3.3. Transients analysis and −1.58 V do not confirm to either of the models of diffusion-
controlled 3D nucleation, however, the rising parts of the transients
To better understand the phenomenon of Zn–Mn alloy nucle- are located near to the progressive nucleation. As soon as the
ation mechanism during the early stage of deposition, different potential turns to more negative values, the nucleation process
current transients recorded from bath 1 and 2, in a potential win- confirms well the progressive nucleation process. But at longer
dow between −1.52 and −1.75 V vs. Ag/AgCl, are shown in Fig. 6. time (t > tmax ), the experimental (j/jmax )2 is much larger than that
These transients exhibit similar behavior; the transients can be calculated from the theoretical model. Two facts can explain this
divided into three regions. In the first region, an initial decrease behavior: hydrogen reduction, as reported by other authors, and
in the current is related to the charging of the double layer [37]. manganese incorporation in Zn–Mn deposits [42,43].
The second region corresponds to an increase in the cathodic cur- On the other hand, Fig. 8 related to the nucleation mechanism
rent density until a maximum current value jmax reached at a of Zn–Mn alloy deposition from bath 1, shows that the initial parts
time tmax . This step is due either to the growth of a new phase of the transients recorded at potentials ranging between −1.52
and/or to an increase of the number of nuclei [37]. The third region and −1.61 V vs. Ag/AgCl match the progressive nucleation (t < tmax ).
corresponds to a plateau of current density which is typical to a These results indicate that during the initial stage of Zn–Mn elec-
diffusion-controlled process [37]. trodeposition on steel substrate, the reaction sites on the surface
As the deposition potentials are more negative, jmax increases are activated progressively. The number of formed nuclei is less
in magnitude, denoting arise of the nucleation and growth rate. than the maximum saturation value and the new nuclei begin to
Similarly, jmax increases as Mn salt concentration in the electrolytic grow progressively [43].
bath is higher. As shown also in Fig. 8, an increase in potential deposition until
Scharifker and Hills [38] developed a theoretical model to ana- E = −1.75 V induces a deviation in the 3D nucleation mechanism.
lyze the current-time transient results of chronoamperometric The current time is well fitted by the progressive nucleation model,
experiments. According to this model [39], there are two limiting at the initial stage (t/tmax )=1, and approaches to the instantaneous
nucleation mechanisms: the instantaneous and the progressive. model till around (t/tmax )=2.
Instantaneous nucleation corresponds to a slow growth of nuclei The above data prove that as soon as the applied potential is
on a small number of active sites; all are activated in the same time. shifted to more negative values, the nucleation process is modified,
So, the number of nuclei would be constant during the deposition suggesting that Mn is probably being more incorporated into the
process and all nuclei have similar age and size [38]. In return, pro- deposits. It must be also stressed that the nucleation mechanism
gressive nucleation corresponds to fast growth of nuclei on many for Zn–Mn electrodeposition is not extremely affected by the Mn2+
active sites, all activated during the course of electro-reduction ions concentration in the electrolytic bath.
N. Loukil, M. Feki / Applied Surface Science 410 (2017) 574–584 579

Fig. 7. Non dimensional plots (j/jmax )2 vs (t/tmax ) of the chronoamperometric curves for Zn–Mn electrodeposited from bath 2.

3.4. Galavanostatic deposition and characterization of Zn–Mn Zn–Mn alloys deposited from the bath 1, comprising 1 M of Mn2+
coatings ions concentration, at 40, 60, 80, 140 and 200 mA/cm2 exhibit an
average deposition potential Ed −1.52, −1.58, −1.61, −1.75 and
3.4.1. Effect of Mn2+ concentration in the plating bath on −2 V vs. Ag/AgCl, respectively.
deposition potential Ed , Mn content and current efficiency Similarly, it is clear that rising Mn2+ ions concentration mod-
Electrodeposition can be performed by controlling either the ifies the shape of the chronopotentiometric curves. At the same
potential or current. In industrial coatings preparation, the cur- deposition condition (jimp = cte), reduction potentials Ed are sig-
rent step method, also known as the galvanostatic method, is the nificantly shifted towards more negative values when Mn2+ ions
most practical [44]. The advantage of the galvanostatic method is concentration is higher.
that the thickness of the as-deposited layer can be easily controlled Fig. 10 shows the Mn content into the Zn–Mn alloys electrode-
according to the Faraday’s law [45]. posited from the two baths 1 and 2 at different current densities
According to voltamogramm data described in Fig. 5, it is clear jimp . Namely, the Zn–Mn alloys deposited from bath 1 contain a
that Zn2+ ions reduction is already under diffusion control before greater amount of Mn as compared to that obtained from bath
Mn is being electrodeposited. Therefore, co-deposition of Mn and 2. Comparing Ed at a constant current density jimp = 40 mA/cm2
Zn as well as formation of Zn–Mn alloys with different composi- (Fig. 9), the deposition potential Ed moves from −1.31 to −1.52 V
tions, by changing the current density, seem to be possible. On vs. Ag/AgCl, 210 mV is more negative, when the Mn2+ ions con-
the basis of this assumption, the applied current densities jimp centration increases from 0.3 to 1 M, respectively. As a result, the
should be superior to the limiting current density of Zn deposi- Mn content into the Zn–Mn film is enhanced from 0.3 to 0.98 wt%
tion in attempt to obtain Zn–Mn alloys with high Mn amount. (Fig. 10). Thus, increasing Mn2+ concentration is obviously benefi-
Galvanostatic experiments were carried out in a range of current cial for an easier Mn incorporation. This is in agreement with the
density varying from 40 to 200 mA/cm2 . Fig. 9 presents the evolu- basic principles of normal alloy electrodeposition [46].
tion of the deposition potential Ed during Zn–Mn electrodeposition The composition analysis of these alloys indicates that Mn
from the two working baths: bath 1 ([Mn2+ ] = 1 M) and bath 2 co-deposition is only possible if the current density is equal to
([Mn2+ ] = 0.3 M). 40 mA/cm2 . Further experiments, conducted at current densities
As seen, a further increase in current density jimp into the same lower than 40 mA/cm2 (Results are not shown), revealed that
bath leads to a notable shift of Ed towards lower values. For instance, no Mn has been incorporated in the coatings implying that the
580 N. Loukil, M. Feki / Applied Surface Science 410 (2017) 574–584

Fig. 8. Non dimensional plots (j/jmax )2 vs (t/tmax ) of the chronoamperometric curves for Zn–Mn electrodeposited from bath 1.

current density of the onset manganese deposition is estimated to Zn–Mn alloy. This significant increase in Mn amount is explained
be about 40 mA/cm2 . Besides, it is stressed to note that there is a by the fact that the deposition current density is applied where
critical value [16,47] of the deposition current density where Mn Zn2+ ions are reduced under diffusion limiting current density.
content is rapidly improved. As it can be seen in Fig. 10, regardless Thus, further increasing in current density (i.e. corresponding to
the electrolytic bath, this value is about 60 mA/cm2 . This value is more negative cathodic potentials) enhances merely the Mn depo-
so close to that of 50 mA/cm2 stated in a previous study on Zn–Mn sition and hydrogen evolution reaction [15]. However, there is
deposition from additives-free electrolyte [47]. a small decrease in the Mn amount at higher current densities
As soon as the deposition current density is higher than (200 mA/cm2 ), as shown in Fig. 10. This limitation of Mn co-
40 mA/cm2 , as Ed is lower and thus becomes closer to that of Mn deposition might be attributed to a certain inhibitor effect caused
(Fig. 9). This will facilitate the deposition rate of the less noble by the concomitant hydrogen evolving reaction [47]. Indeed, in
metal, Mn. Whatever the electrolytic bath, an appreciable increase spite of the descending shift of the deposition potential Ed (Fig. 9),
in the Mn content is observed with increasing jimp (Fig. 10). Similar the Mn content is almost stable at very high current densities,
results have been pointed out [2,16] that the increase in deposi- revealing subsequently a significant strengthening of the hydrogen
tion current density is beneficial for enhancing Mn content into evolving kinetic at high current densities. This intensive hydrogen

Fig. 9. Potential-time transients for Zn–Mn alloy electrodeposition from the two electrolytes bath 1 and bath 2 at various current densities (jimp ).
N. Loukil, M. Feki / Applied Surface Science 410 (2017) 574–584 581

14 without stirring

with stirring
12

10

% Mn in alloy
8

0
0 50 100 150 200 250
jimp (mA/cm2)
Fig. 10. Dependence of Mn content in Zn–Mn alloys on the current density from
two acidic chloride baths (bath 1 and bath 2).
Fig. 12. Effect of stirring the solution on Mn content in Zn–Mn alloys electrode-
posited from bath 1.
95 bath 1
bath 2
3.4.3. Effect of the applied current density jimp on the morphology
85
and crystallographic structure of Zn–Mn alloys
current efficiency (%)

The effect of current density jimp on the morphology and


75
structure of Zn–Mn deposits obtained from bath 1 were inves-
tigated using SEM and XRD, respectively. Visual examination of
65 the obtained Zn–Mn deposits indicates that all deposits cover
completely the substrate surface and their appearance depends
55 systematically on jimp .
At same deposition conditions (jimp = cte), it should be stated
that the Zn –Mn coatings electrodeposited from the both baths
45
0 50 100 150 200 250 (bath 1 and bath 2) show a similar morphology. Nevertheless, some
jimp (mA/cm²) differences can be observed as current density jimp increases for the
same electrolytic bath. On the basis of visual observations, Zn–Mn
Fig. 11. Current efficiency for Zn–Mn alloys deposition from (bath 1 and bath 2) at coatings electrodeposited at low current density seem to be gray
different current densities. colored and slightly dendritic, whereas those produced at high val-
ues, containing an enhanced Mn content, are darker, porous and
powdery. The difference in color may be due to Mn content into
evolution causes probably a detachment of Zn–Mn alloys in the
the deposits. These visual observations are in good agreement with
form of fine powders from the cathode causing a loss of final weight
the literature results [6,16].
of the Zn–Mn coatings electrodeposited at high current densities
According to SEM examination of the Zn–Mn deposits, the mor-
[47]. This plausible reason explains the fall in the current efficiency
phology varies from fine grains to coarse grains as jimp increases.
at high current densities, as shown in Fig. 11. Such occurrence has
These observations reveal that a further incorporation of Mn into
been reported in recent investigations on Zn–Mn coatings [6,47].
the Zn matrix alters significantly the surface morphology of the
As shown in Fig. 11, the highest current efficiency is observed at
Zn–Mn alloys.
low current density. This finding is very close to the results reported
Moreover, a large quantity of dendrites is clearly observed, par-
by Bucko et al. [16]. Moreover, it should be underlined that the
ticularly on the edges of Zn–Mn deposits, when the jimp increases.
cathodic current efficiency increases with the raise of Mn2+ con-
As mentioned earlier, the co-evolution of hydrogen is a strong con-
centration in the electrolytic bath [26]. In this regard, Ganesan and
comitant reaction when Mn2+ reduction occurs. In fact, hydrogen
his co-workers [6] demonstrated that hydrogen evolution reaction
bubbles formation leads subsequently to a poorly adherent, den-
decreases with increasing Mn2+ concentration from around 0.4 to
dritic and rough coatings (observed by naked eye) [6,47,49]. This
2 M.
finding is more pronounced for Zn–Mn deposited from the bath 1 at
200 mA/cm2 . As denoted previously in the corresponding transient
3.4.2. Effect of stirring on Mn content curve at this current density (Fig. 9), the descending evolution of
Taking into account all previous considerations, namely the cur- Ed for longer deposition time may be ascribed to an increase of the
rent efficiency and the Mn content, bath 1, containing 1 M of MnCl2 , effective cathodic area. So, this increase seems to be linked to this
was retained to study the effect of stirring and to characterize particular morphology.
Zn–Mn alloys. Fig. 12 illustrates the Mn content in the Zn–Mn plated To illustrate the effect of current density on surface morphology
without and with stirring of the solution, at different applied cur- and structure properties, SEM micrographs of two selected sam-
rent densities jimp . As can be seen in Fig. 12, stirring of the solution ples deposited at 60 and 140 mA/cm2 are presented in Fig. 13 (a, c).
decreases abruptly the Mn amount. Previous studies showed simi- Fig. 13-a presents the Zn–Mn coating produced at low current den-
lar results [15–34]. It may be assumed that stirring of the electrolyte sity (60 mA/cm2 ). The latter exhibits a homogeneous deposit with
enhances electrodeposition of Zn. This could be due either to an hexagonal crystals. A more dendritic morphology is observed in the
acceleration of the diffusion controlled Zn2+ reduction [15] or to sample obtained at 140 mA/cm2 (Fig. 13c), but in this case the den-
the removal of hydrogen bubbles from the metallic surface by the drites show coarse-grained deposit with non-uniform crystal size.
stirring that leads to an increase of its active area and consequently This Zn–Mn deposit is clearly more porous, due to hydrogen evolu-
to a decrease in the real current density [34]. tion during Zn and Mn co-deposition. This agrees well with results
582 N. Loukil, M. Feki / Applied Surface Science 410 (2017) 574–584

Fig. 13. SEM micrographs and X-ray diffraction patterns of two Zn–Mn samples deposited onto steel substrate from bath 1 at different current densities (a–b): 60 mA/cm2 ;
(c–d): 140 mA/cm2 .

previously stated [16,48,50], evidencing that a dendritic shape is the two others detected phases appear with weak reflection lines.
characteristic of metal coatings deposited at very negative deposi- Several works [2,34] reported that the Zn–Mn coatings consisting
tion potentials (i.e. high current densities) as the growth process is of only ␧-Zn–Mn phase provide the best corrosion protection.
under diffusion control. It is striking to note that the ␩-Zn phase [51] has also Hexago-
Fig. 13(b, d) show typical diffractograms for the two selected nal Close Packed (HCP) structure. Thus, the hexagonal morphology
Zn–Mn coatings electrodeposited from bath 1. Only peaks whose observed in SEM photograph is basically attributed to this phase.
relative intensity is above 30% are considered. The X-ray diffraction Such occurrence is in agreement with the results of Ortiz et al. [51].
pattern of the considered Zn–Mn coatings reveal that jimp affects Contrary to the Zn–Mn deposited at low current density
significantly the structure of the obtained deposits. (60 mA/cm2 ), the diffractogram corresponding to the Zn–Mn
The phases were identified after an accurate analysis of the peak deposited at higher current density (140 mA/cm2 ), containing
positions and comparison with structural parameters obtained in 11.4% of Mn, seems to be metallic glass or amorphous [16,21]. It
previous works on Zn–Mn alloys electrodeposited from chloride is clear from Fig. 13d that there are no significant signal attributed
acidic baths [27,34,51]. It can be stated from Fig. 13b that the film to the crystalline structure of Zn, Mn or Zn–Mn alloy. This indicates
electrodeposited at 60 mA/cm2 , containing 5.5%, Mn is tri-phasic a transition of the Zn–Mn deposit from crystalline, at low current
consisting of three crystalline phases. An additional diffraction peak density, to an amorphous structure at high current density. This
corresponding to the Fe substrate is also detected. The XRD diffrac- finding also corroborates that reported by Gong et al. who investi-
togram (Fig. 13b) shows a mixture of ␩ Zn (a solid solution of less gated electrodeposition of Mn [28,52,53] and Mn alloys [49]. These
than 1% of Mn in Zn) and intermetallic Zn–Mn phases with differ- authors noticed that after exceeding a critical deposition current
ent crystallographs namely, monoclinic ␨-MnZn13 [15] and ␧-HCP density, a transition from crystalline to amorphous structure may
(Hexagonal Close Packed) with crystal orientations (101) and (112). occur in the Mn alloy electrodeposits such as (Sn–Mn) [49]. The
Very similar structure properties were reported in a previous work crystalline to amorphous transition in Mn and Mn alloys is due to
[34] dealing with Zn–Mn coatings deposited from an acidic chloride the extremely high nucleation rate induced by the high applied
bath. The two crystalline phases ␨-MnZn13 and ␩-Zn detected in the overvoltage and by the strong hydrogen evolution and potential
XRD diffractogram corresponding to the sample electrodeposited incorporation [49,50].
at 60 mA/cm2 may have their origin in the high Zn content and Visual examinations denote that Zn–Mn coatings obtained
subsequently the low Mn amount (5%) in the coating [34]. Results under high current densities are dark tending to be black deposits,
previously obtained from acidic chloride bath [15] confirm also the particularly at the edges of Zn–Mn samples. According to literature
result of the present work. Sylla et al. [15] demonstrated that ␧- [54], the oxy/hydroxide compounds are responsible for the black
Zn–Mn was the main phase detected. Similarly, in this work, the appearance. Likewise, black Mn deposits obtained at high current
␧-HCP (101) is the major phase in the XRD pattern (Fig. 13b), while densities are well documented [21]. Additionally, recent research
N. Loukil, M. Feki / Applied Surface Science 410 (2017) 574–584 583

[28] demonstrated the formation of black films at the edges of mainly to hydrogen evolution that becomes excessively intense at
Mn samples at current densities above 150 mA/cm2 . This finding high current densities.
confirms then that the oxy/hydroxide inclusions are strongly incor- The deposition current density significantly affects the surface
porated into Zn–Mn deposits. morphology and crystal structure. XRD studies shows that the
The most probable explanation for oxy/hydroxides formation is Zn–Mn alloy electrodeposited at a low current density, more accu-
the intense hydrogen evolution occurring in parallel to Zn–Mn co- rately at 60 mA/cm2 , is tri-phasic and containing mainly a mixture
deposition at high current densities. Because this large amount of of ␩ Zn, ␨-(MnZn13 ) and ␧-HCP. Meanwhile, in the case of a high
hydrogen gas evolved at the cathode surface, hydroxyl ions (OH− ) current density, i.e. 140 mA/cm2 , the Zn–Mn coating is amorphous.
are abundant in the near-cathodic layer. Local pH in the vicinity This transition from crystalline to amorphous structure is certainly
of the cathode surface becomes then much higher than that of the due to the incorporation of non-metallic hydroxide inclusions in
bulk solution, leading to a quick alkalization of the metal/solution the Zn–Mn alloy electrodeposits at high current densities.
interface [17–26]. It is well known that the hydroxides forma-
tion Zn(OH)2 and/or Mn(OH)2 is excessively occurring at high pH
resulting at the cathode surface. Pourbaix [55] stated that pH at Acknowledgement
which insoluble Zn(OH)2 and Mn(OH)2 starts to precipitate are
around 5.5 and 7.5, respectively. Bucko et al. [16,47] studied exten- This work was supported by the Ministry of High Education and
sively the Zn–Mn co-deposition and reported that the buffering Scientific Research, Tunisia.
effect ascribed to boric acid and the complexation with Cl− ions
can prevent hydroxide formation only to a limited extent, but the
formation of Zn(OH)2 and Mn(OH)2 can occur at sufficiently high References
pH.
[1] G.D. Wilcox, D.R. Gabe, Electrodeposited zinc alloy coatings, Corros. Sci. 35
Several authors [28,50,54] demonstrated that hydroxide com-
(1993) 1251–1258.
pounds greatly inhibit crystal growth and subsequently lower [2] N. Boshkov, Galvanic Zn-Mn alloys-electrodeposition, phase composition
appreciably the current efficiency, as demonstrated above in Fig. 11. corrosion behavior and protective ability, Surf. Coat. Technol. 172 (2003)
Hence, Zn–Mn coatings predominantly grow as hydroxides at high 217–226.
[3] N. Boshkov, S. Vitkova, K. Petrov, Corrosion products of zinc-manganese
current densities. coatings: part I investigations using microprobe analysis and X-ray
diffraction, Met. Finish. 99 (2001) 56–60.
[4] N. Boshkov, K. Petrov, S. Vitkova, Corrosion products of zinc-manganese
coatings part III: double-protective action of manganese, Met. Finish. 100
(2002) 98–102.
4. Conclusion [5] B. Bozzini, A. Accardi, P.L. Cavallotti, F. Pavan, Electrodeposition and plastic
behavior of low manganese Zn-Mn alloys coating for automotives
applications automotive, Met. Finish. (1999).
Zn–Mn coatings were successfully electrodeposited on steel [6] S. Ganesan, G. Prabhu, B.N. Popov, Electrodeposition and characterization of
substrate from an additives-free chloride bath at current den- Zn-Mn coatings for corrosion protection, Surf. Coat. Technol. 238 (2014)
sities superior to the limiting current density of Zn deposition. 143–151.
[7] O.D. Morales, J. Mostany, C. Borras, B.R. Scharifker, Current transient study of
An electrochemical investigation of Zn–Mn electroplating process
the kinetics of nucleation and diffusion-controlled growth of bimetallic
with different Mn2+ ions concentration (0.3 and 1 M) has been phases, J. Solid State Electrochem. 17 (2013) 345–351.
conducted. Results revealed that the increase in Mn2+ ions con- [8] Y. Tanaka, The forefront of practical superconducting wires, Physica C 335
(2000) 69–72.
centration enables its faster reduction on the cathode and acts as
[9] A. Sulcius, E. Griskonis, Silvery matt coatings of Zn–Mn alloy, Trans. Inst. Met.
inhibitor towards Zn deposition. Finish. 86 (3) (2008) 153–156.
A dimensionless graph model was used to analyze the effect of [10] B. Bozzini, E. Griskonisb, A. Fanigliulo, A. Sulcius, Electrodeposition of Zn-Mn
both Mn2+ ions concentration and the applied cathodic potential on alloys in the presence of thiocarbamide, Surf. Coat. Technol. 154 (2002)
294–303.
Zn–Mn nucleation process. This study showed that the initial stage [11] D.R. Gabe, G.D. Wilcox, A. Jamani, B.R. Pearson, Zinc-manganese alloy
of Zn–Mn electroplating depends mainly on the applied potential. electrodeposition, Met. Finish. 91 (1993) 34–36.
Several deposition parameters were explored in this study, [12] B. Bozzini, Morphological artefacts in EDX analyses of electrodeposited
Zn-Mn films, Trans. Inst. Met. Finish. (2000) 78–93.
namely Mn2+ ions concentration, current density and stirring con- [13] M.F. de Carvalho, W. Rubin, I.A. Carlos, Study of the influence of the
ditions. Their effects on the Zn–Mn electroplating process, i.e. the polyalcohol mannitol on zinc electrodeposition from an alkaline bath, J. Appl.
resulting Mn content into the coatings, the current efficiency as Electrochem. 40 (2010) 1625–1632.
[14] J.S. Alabaster, P.D. Anderson, D. Calamari, V. Dethlefsen, M. Grande, R. Lloyd,
well as the coating morphology were investigated. To sum up, J.-L. Gaudet, R. Welcome, Water quality criteria for European freshwater fish
results showed that Mn content in Zn–Mn coatings is significantly report on combined effects on freshwater fish and other aquatic life of
enhanced with the increasing of either Mn2+ concentration or the mixtures of toxicants in water, EIFAC 37 (1980) 1–49.
[15] D. Sylla, J. Creus, C. Savall, O. Roggy, M. Gadouleau, Ph. Refait,
applied current density, in particular. Indeed, chronopotentiometry
Electrodeposition of Zn–Mn alloys on steel from acidic Zn–Mn chloride
results display that increasing of each those two deposition condi- solutions, Thin Solid Films 424 (2003) 171–178.
tion, the deposition potential Ed is shifted towards more negative [16] M. Bučko, J. Rogan, B. Jokić, M. Mitrić, U. Lačnjevac, J.B. Bajat,
Electrodeposition of Zn-Mn alloys at high current densities from chloride
values, close to that of Mn. This, in turn, favors Mn incorporation
electrolyte, J. Solid. State Electrochem. 17 (2013) 1409–1419.
into the deposit. [17] M.V. Tomić, M.M. Bučko, M.G. Pavlović, J.B. Bajat, Corrosion stability of
Moreover, it was concluded that the electrolytic bath containing electrochemically deposited Zn-Mn alloy coatings, Contemp. Mater. (2010)
1 M of Mn2+ shows a simultaneous improvement of both cathodic 87–93, I–1.
[18] Y. Yang, S. Liu, X. Yu, C. Huang, S. Chen, G. Chen, Q.-H. Wu, Effect of additive on
current efficiency and Mn content. For this reason, the electrolyte zinc electrodeposition in acidic bath, Surf. Eng. (2014) 1–6.
containing 1 M of MnCl2 was selected as a better solution to conduct [19] S. Basavanna, Arthoba Naik, Electrochemical studies of Zn–Ni alloy coatings
further experiments. from acid chloride bath, J. Appl. Electrochem. 39 (2009) 1975–1982.
[20] C.J. Clarke, G.J. Browning, S.W. Donne, An RDE and RRDE study into the
Increasing the current density far from the limiting current den- electrodepositionof manganese dioxide, Electrochim. Acta 51 (2006)
sity of Zn deposition generates an increase in the Mn amount 5773–5784.
in Zn–Mn alloys electrodeposits. The Mn content is markedly [21] W.H. Bradtand, H.H. Oaks, The electrodeposition of manganese from aqueous
solutions. ii. Sulfate electrolytes, Trans. Electrochem. Soc. 71 (1) (1937)
improved at a critical current density of 60 mA/cm2 . However, a fur- 279–286.
ther increase in the current density until 200 mA/cm2 has a reverse [22] M. Gonsalves, D. Pletcher, The electrodeposition of manganese from aqueous
effect and causes a decrease in Mn content. This finding is due solutions. I. Chloride electrolytes, J. Elecroanal. Chem. 285 (1990) 185–193.
584 N. Loukil, M. Feki / Applied Surface Science 410 (2017) 574–584

[23] The effect of nonionic surfactant Triton X100 during electrochemical [39] G. Oskam, J.G. Long, A. Natarajan, P.C. Searson, Electrochemical deposition of
deposition of MnO2 on its capacitance properties, J. Electrochem. Soc. 154(10) metals onto silicon, J. Phys. D 31 (1998) 1927.
(2007) 901–909. [40] M.P. Pardave, M.T. Ramirez, I. Gonzales, A. Serruya, B.R. Scharifker, Silver
[24] Q.B. Zhang, Y. Hua, Effect of Mn2+ ions on the electrodeposition of zinc from electrocrystallization on vitreous carbon from ammonium hydroxide
acidic sulphate solutions, Hydrometallurgy 99 (2009) 249–254. solutions, J. Electrochem. Soc. 143 (5) (1996) 1551–1558.
[25] J. Lu, D. Dreisinger, T. Gluck, Manganese electrodeposition—a literature [41] K. Raeissi, A. Saatchi, M.A. Golozar, Texture and surface morphology in zinc
review, Hydrometallurgy. 141 (2014) 105–116. electrodeposits, J. Appl. Electrochem. 33 (635) (2004) 1249–1258.
[26] F. Xu, Z. Dan, W. Zhao, G. Han, Z. Sun, K. Xiao, L. Jiang, N. Duan, [42] C. Han, Q. Liu, D.G. Ivey, Nucleation of Sn and Sn–Cu alloys on Pt during
Electrochemical analysis of manganese electrodeposition and hydrogen electrodeposition from Sn?citrate and Sn–Cu–citrate solutions, Electrochim.
evolution from pure aqueous sulfate electrolytes with addition of SeO2 , J. Acta 54 (2009) 3419–3427.
Electroanal.Chem. 741 (2015) 149–156. [43] S. Boudinar, N. Benbrahim, B. Benfedda, A. Kadri, a E. Chainet, L. Hamadoua,
[27] P. Díaz-Arista, G. Trejo, Electrodeposition and characterization of manganese Electrodeposition of heterogeneous Mn-Bi thin films from a sulfate-nitrate
coatings obtained from an acidic chloride bath containing ammonium bath: nucleation mechanism and morphology, J. Electrochem. Soc. 161 (5)
thiocyanate as an additive, Surf. Coat. Technol. 201 (2006) 3359–3367. (2014) 227–234.
[28] J. Gong, Giovanni Zangaria, Electrodeposition and characterization of [44] S. Wen, J.A. Szpunar, Nucleation and growth of tin on low carbon steel,
manganese coatings, J. Electrochem. Soc. 149 (2002) 209–217. Electrochim. Acta 50 (2005) 2393–2399.
[29] P.D. Arista, R. Antano Lopez, Y. Measa, R. Ortega, E. Chainet, P. Ozil, G. Trejo, [45] C. Han, Qi Liu, D.G. Ivey, Nucleation of Sn and Sn–Cu alloys on Pt during
EQCM study of the electrodeposition of manganese in the presence of electrodeposition from Sn–citrate and Sn–Cu–citrate solutions, Electrochim.
ammonium thiocyanate in chloride-based acidic solutions, Electrochimi. Acta Acta 54 (12) (2009) 3419–3427.
51 (2006) 4393–4404. [46] A. Brenner, Electrodeposition of Alloys, vol. I, Academic Press, New York, 1963.
[30] E. Rudnik, Effect of gluconate ions on electroreduction phenomena during [47] M. Bučko, J. Rogan, S.I. Stevanović, S. Stanković, J.B. Bajat, The influence of
manganese deposition on glassy carbon in acidic chloride and sulfate anion type in electrolyte on the properties of electrodeposited Zn-Mn alloy
solutions, J. Electroanal. Chem. 741 (2015) 20–31. coatings, Surf. Coat. Technol. 228 (201) (2016) 221–228.
[31] G. Trejo, R. Ortega B, Y. Meas, P. Ozil, E. Chainet, B. Nguyen, Nucleation and [48] P.S.D. Brito, Sandra Patrício, Luiz F. Rodrigues, César A.C. Sequeira,
growth of zinc from chloride concentrated solutions, J. Electrochem. Soc. 145 Electrodeposition of Zn-Mn alloys from recycling Zn-MnO2 batteries
(1998) 4090–4097. solutions, Surf. Coat. Technol. 206 (2012) 3036–3047.
[32] T. Casanova, F. Soto, M. Eyraud, J. Crousier, Hydrogen absorption during zinc [49] J. Gong, G. Zangari, Electrodeposition of sacrificial tin-manganese alloy
plating on steel, Corros. Sci. 39 (1997) 529–537. coatings, Mater. Sci. Eng. A 344 (2003) 268–278.
[33] S. Boudinar, N. Benbrahim, B. Benfedda, A. Kadri, E. Chainet, L. Hamadoua, [50] V.D. Jović, B.M. Jović, M.G. Pavlović, Electrodeposition of Ni, Co and Ni–Co
Electrodeposition of heterogeneous Mn-Bi thin films from a sulfate-nitrate alloy powders, Electrochim. Acta 51 (2006) 5468–5477.
bath: nucleation mechanism and morphology, J. Electrochem. Soc. 161 (2014) [51] Z.I. Ortiz, P. Díaz-Arista, Y. Meas, R. Ortega-Borges, G. Trejo, Characterization
227–234. of the corrosion products of electrodeposited Zn, Zn–Co and Zn–Mn alloys
[34] C. Savall, C. Rebere, D. Sylla, M. Gadouleau, P. Refait, J. Creus, Morphological coatings, Corros. Sci. 51 (2009) 2703–2715.
and structural characterisation of electrodeposited Zn–Mn alloys from acidic [52] J. Gong, I. Zana, G. Zangari, Electrochemical synthesis of crystalline and
chloride bath, Mater. Sci. Eng. A 430 (2006) 165–171. amorphous manganese coatings, J. Mat. Sci. Lett. 20 (2001) 1921–1923.
[35] S. Fletcher, C.S. Halliday, D. Gates, M. Westcott, T. Lwin, G. Nelson, The [53] J. Gongand, G. Zangari, Increased metallic character of electrodeposited mn
response of some nucleation/growth processes to triangular scans of coatings using metal ion additives, Electrochem. Solid-State Lett. 7 (9) (2004)
potential, J. Electroanal. Chem. 159 (1983) 267–285. 91–94.
[36] G. Gunawardena, G. Hills, I. Montenegro, Electrochemical nucleation: part II. [54] P. Wei, O.E. Hileman Jr., M. Reza Bateni, X. Deng, A. Petric, Manganese
The electrodeposition of silver on vitreous carbon, J. Electroanal. Chem. 138 deposition without additives, Surf. Coat. Technol. 201 (2007) 7739–7745.
(1982) 241–254. [55] M. Pourbaix, Atlas of Electrochemical Equilibria in Aqueous Solutions, 2nd
[37] S. Basavanna, Arthoba Naik, Electrochemical studies of Zn-Ni alloy coatings edn., Pergamon Press Ltd., Oxford, 1966.
from acid chloride bath, J. Appl. Electrochem. 39 (2009) 1975–1982.
[38] B. Scharifker, G. Hills, Theoretical and experimental studies of multiple
nucleation, Electrochim. Acta 28 (1983) 879–889.

You might also like