You are on page 1of 9

Surface & Coatings Technology 240 (2014) 311–319

Contents lists available at ScienceDirect

Surface & Coatings Technology


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

Tin–zinc alloy electrodeposition from aqueous citrate baths


Honorata Kazimierczak a,⁎, Piotr Ozga a, Aldona Jałowiec b, Remigiusz Kowalik b
a
Institute of Metallurgy and Material Science, Polish Academy of Sciences, 30-059 Krakow, Reymonta 25, Poland
b
AGH University of Science Technology, Faculty of Non-Ferrous Metals, 30-059 Krakow, al. Mickiewicza 30, Poland

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

Article history: The process of tin–zinc alloy electrodeposition from aqueous citrate electrolytes was studied. The influence of ap-
Received 20 September 2013 plied potential, current density, hydrodynamic conditions, electrolyte composition and charge transfer on the
Accepted in revised form 19 December 2013 electrodeposition of Sn–Zn alloy was determined. Depending on these parameters, coatings with different com-
Available online 29 December 2013
positions and appearances can be obtained. The surface composition of deposits was ascertained by chemical
analysis (WDXRF). The morphology of coatings was studied by SEM. The electrodeposition of bright, shiny Sn–
Keywords:
Sn–Zn alloys
Zn coatings on steel is possible from the citrate baths studied. The presence of zinc ions in the electrolyte results
Coatings in a significant inhibition of tin reduction rate, which enables the electrodeposition of tin–zinc alloy layers con-
Electrodeposition taining from about 0.5 to 75 wt.% of Zn, with high current efficiency, within the range from 80% to almost
Citrate baths 100%. Two phases are formed in deposits: a hexagonal zinc phase and a tetragonal β-tin phase.
© 2013 Elsevier B.V. All rights reserved.

1. Introduction difference in the standard redox potentials of tin and zinc (620 mV
[9]) means that it is essential to add an appropriate complexing agent
Tin–zinc alloy layers are widely used in the metal finishing industry which enables the electrodeposition of Zn–Sn alloy.
because they have a number of attractive properties. They have good The first patents describing the electrochemical method of tin–zinc
corrosion resistance, good friction properties, high wear resistance and alloy deposition date from the beginning of the 20th century [10–14].
excellent solderability [1–3]. They are especially interesting as a re- These alloys were already proposed as a replacement for cadmium coat-
placement for toxic tin–lead solders, because both of these alloys have ings from an economic point of view, because the price of cadmium was
similar melting temperatures [4,5]. A further advantage is the substan- 2–3 times greater than the price of tin in that period. Then the alloys
tially lower cost of such alloys in comparison to the alternative alloys were obtained from cyanide baths [15] so the main problem was the
based on tin and silver. In this case, alloys containing about 8–9 wt.% toxicity of the electrolytes used.
of zinc are applied. However, due to their very good corrosion resis- Moreover, electrodeposition of Sn–Zn alloy with a high content of
tance, Sn–Zn alloy layers are often used as protective coatings. In this zinc (N50%) from these baths was not possible.
case, the optimum zinc content in the alloy is about 20–30 wt.% [6–8]. In recent years the number of papers describing the electrodeposi-
There is currently a great interest in this type of coating, in relation to tion of Sn–Zn alloy from various electrolytes has increased, because it
the need to eliminate of anticorrosive cadmium coatings, which were is currently considered important to find an environmentally friendly
widely used but proved to be carcinogenic and highly toxic, and hence way of preparing alternative, non-toxic coatings [16–25].
their use is restricted by the European Union's Restriction of Hazardous Vitkova et al. [18] describe a gluconate–citrate bath for the prepara-
Substances directive. By comparison, the carcinogenicity of tin and its tion of Sn–Zn alloy with low tin content. Guaus et al. [23,24] examine
compounds was not observed in animal studies. Moreover the Sn–Zn sulphate–gluconate and sulphate–tartate baths. Gluconate baths have
alloy coatings on steel combine the anti-corrosion barrier properties of also been used for the preparation of Sn–Zn alloy solder by Hu et al.
tin and zinc (barrier protection) with sacrificial properties of zinc (ca- [25], while studies on alkaline baths were conducted by Dubent et al.
thodic protection). [6].
Tin–zinc alloys can be obtained by electrodeposition, which is a rel- In this work, citrate baths are proposed for the electrodeposition of
atively straightforward and low cost technique. However, a significant Sn–Zn alloys because citrates are non-toxic and form strong complexes
with Zn(II) and Sn(II). The previous investigation by Ozga and
Kazimierczak [26,27] proved the possibility of Sn–Zn electrodeposition
from citrate baths.
⁎ Corresponding author. Tel.: +48 12 2952812; fax: +48 12 2952804. The purpose of this work was to study the kinetics of co-reduction of
E-mail address: h.kazimierczak@imim.pl (H. Kazimierczak). tin and zinc from non-toxic citrate electrolytes, and determine the

0257-8972/$ – see front matter © 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.surfcoat.2013.12.046
312 H. Kazimierczak et al. / Surface & Coatings Technology 240 (2014) 311–319

optimal ranges of electrodeposition parameters enabling the 3. Results and discussion


preparation of good quality coatings with varying contents of tin and
zinc. The influence of deposition potential, current density, concentra- 3.1. Development of stable Sn(II)–Zn(II)-cit baths
tion of bath constituents and variable hydrodynamic conditions on
the composition and morphology of deposited alloy coatings was The preparation of stable solutions is one of the basic problems in the
investigated. development of the electrochemical method of alloy deposition. The
electrolytic bath should be homogeneous, with tin(II) and zinc(II) in
forms of electroactive aqueous species. Type of complexes formed in
2. Experimental aqueous citrate solutions depends on pH, concentration of components
and temperature. Various monomeric, dimeric, and heteronuclear com-
The chemical compositions of the electrolytes studied are given in plexes, and sparingly soluble polymeric complexes can also be formed.
Table 1. The solution pH was adjusted to 5.0 by the addition of sulphuric The thermodynamic model was built on the basis of stability constants
acid. All chemicals used in this work were of analytical grade. All from the appropriate databases [28–32] and conditional stability con-
electrochemical measurements were carried out in a 100 cm3 cell, in a stants determined in previous research [26,33]. The primary task of
system with a rotating disc electrode (RDE) to ensure constant and con- the model is to indicate the range of bath parameters for further precise
trolled hydrodynamic conditions. The measurements were performed experimental studies.
potentiostatically in a three-electrode cell or galvanostatically in a two- The study of the Sn(II)–Zn(II)–H4cit system did not indicate the for-
electrode cell, by means of potentiostat/galvanostat PGSTAT302N. In mation of heteronuclear complexes containing both metals in the com-
the voltammetric studies the working electrode was copper disc mon ion form [32]. Fig. 1a,b presents the diagram of phases and
(surface area 2.83 cm2) and a Sn rod was used as the counter electrode
(surface area 5 cm2). The working electrode potentials were referred to
the saturated calomel electrode (SCE) and were corrected for ohmic
drop (CI method). During the potentiostatic and galvanostatic deposi-
tion a low-carbon steel disc was the cathode (surface area 2.83 cm2).
A 5 cm2 Sn rod was used as an anode. Steel electrodes were chemically
polished using a solution of oxalic acid and a 30% solution of hydrogen
peroxide (mixture of 14 ml oxalic acid 100 g/dm3 with 2 cm3 30%
H2O2 and 40 cm3 H2O in 40 °C). The chemical composition of
electrodeposits was characterised by the wavelength dispersive X-ray
fluorescence method (WDXRF). The morphology of the resulting
coatings was studied by a scanning electron microscopy (SEM). X-ray
diffractometry (XRD) was used to identify the phase composition
of the deposits. The SEM analysis of surface of coatings was
performed using the Hitachi SU-70 scanning electron microscope. The
SEM analysis of cross section of coating was performed using a JEOL
JSM-7500F Field Emission Scanning Electron Microscope equipped
with INCA PentaFETx3 Oxford Instruments EDS system. The WDXRF
analysis was determined via a Rigaku Primini spectrofluorimeter
using scintillation counters (LiF1 crystal). XRD measurements
were made on a sing the Rigaku Miniflex II diffractometer with Cu K
radiation.

Table 1
Chemical composition of electrolytic baths used.

No. Electrolyte composition + 0.5 g/dm3 PEGa + 0.05 g/dm3 SDSb,


pH = 5

C6H5Na3O7 ∙ 2H2O, Mc ZnSO4 ∙ 7H2O, M SnSO4, M

1 0.25 0.16 –
2 0.25 – 0.08
3 0.25 0.16 0.08
4 0.35 0.16 0.08
5 0.45 0.16 0.08
6 0.55 0.16 0.08
7 0.65 0.16 0.08
8 0.25 0.16 0.02
9 0.25 0.16 0.04
10 0.25 0.16 0.06
11 0.25 0.16 0.10
12 0.25 0.16 0.12
13 0.25 0.08 0.08
14 0.25 0.12 0.08
Fig. 1. The dependence of phase and dominant species in electrolyte on citrate ion concen-
15 0.25 0.14 0.08
tration and pH (for solutions containing 0.16 M ZnSO4 and 0.08 M SnSO4): (a) species of
16 0.25 0.20 0.08
zinc(II) and (b) species of tin(II). Dark grey areas — non-electroactive citrate complexes,
a
PEG-3000 — polyethylene glycol 3000 (molecular mass between 2700 and 3300). orange areas — solid phases built of hydroxides (precipitates immediately), yellow
b
SDS — sodium dodecyl sulphate salt. areas — solid phases built of polymers (precipitates after approximately a few weeks),
c
Denoted in text as Na3Hcit (where cit = C6H4O7). white areas — stable homogeneous solutions.
H. Kazimierczak et al. / Surface & Coatings Technology 240 (2014) 311–319 313

dominant species in analysed electrolytes, depending on the sodium cit- at a potential of about −0.8 V vs. SCE, and the plateau of the cathodic
rate concentration and pH. current density can be observed in the range of potentials from −1.1.
The area of dominance of the Zn2cit complex is marked as sedi- to − 1.4 V vs. SCE. Next, a rapid increase of current density starts,
ment because this species builds polymers in which the ligand neu- which is due to the hydrogen coevolution process. During the reverse
tralises the metal charge by binding to the metal simultaneously scan an anodic peak (a 1 ), connected with tin oxidation, starts at
through the various groups. The resultant polymers form insoluble − 0.7 V vs. SCE in these conditions. In the case of voltammograms
solid phases. This is one of the reasons for citrate baths' instability, obtained in citrate electrolyte containing only zinc ions, a peak (c2) con-
in certain ranges of their compositions. Besides the formation of nected with the underpotential deposition of zinc on the copper sub-
polymers, another reason of instability of citrate solutions is the pos- strate occurs at a potential of − 1.1 V vs. SCE, with a maximum at
sibility of the formation of sparingly soluble hydroxides and oxides, about − 1.2 V vs. SCE. Immediately after this region a slight decrease
as in the case of simple salt-based electrolytes. However, the pre- of current density can be observed and the electrodeposition of zinc
sented thermodynamic model considers only the presence of hy- starts at a potential of about −1.5 V vs. SCE. The cathodic current den-
droxides, because the formation of oxides in this system is sity increases significantly from about 0.3 A/dm2 at − 1.5 V vs. SCE to
relatively slow enough to be neglected. Areas of the formation of hy- 4.4. A/dm2 at −2.0 V vs. SCE. This indicates a high rate of zinc electro-
droxides are sufficient to determine the parameters useful in deposition at highly negative potentials, which occurs simultaneously
electroplating bath preparation. with hydrogen coevolution. Next, when scanning towards the positive
A homogeneous area with predominant forms of tin citrate com- range, the anodic peak (a2) indicating the oxidation of zinc occurs
plexes is in the weakly acidic range of the bath's pH (Fig. 1b). Areas starting from the potential of − 1.1 V vs. SCE. The simultaneous pres-
dominated by high-negative zinc citrate complexes (dark grey ence of both zinc and tin ions in the citrate electrolyte do not cause sig-
areas in Fig. 1a) are homogeneous. However, electrochemical studies nificant changes in the voltammetric response. First, the peak indicating
[30] have shown that these complexes are useless in the direct prep- the underpotential deposition of tin occurs on the cathodic part of the
aration of the electrolytic bath because they are electrochemically voltammogram. It is present in the same range of potential as that in
inactive and do not undergo the reduction processes at the elec- the case of electrolyte without zinc ions (c1). Then the plateau of ca-
trodes, in studied conditions. Nonetheless, the area of electrolyte pa- thodic current, connected with tin deposition, is apparent. It can be
rameters in the slightly acidic range (pH of about 5) indicates noted that the value of current density in this instance is smaller than
simultaneous occurrence of the dominant forms of stable, during the separate reduction of tin. Next, a rapid increase of current
electroactive citrate complexes of tin and zinc. Furthermore, such density occurs, starting from the potential of about − 1.5 V vs. SCE.
an area of bath homogeneity is favourable in electroplating because Such a course for this part of voltammogram is similar for all examined
of the low chemical and corrosion aggressiveness of the bath in this systems, containing tin and zinc ions both separately and together.
pH range. However, the value of the current density observed in the Sn–Zn-cit
system is very close to that observed in Sn-cit system. During the
3.2. Voltammetric study reverse scan, two separated peaks are apparent: a3, which is analogous
to the peaks of a2 (zinc oxidation), and a4, which is analogous to a1 (tin
Fig. 2 presents cyclic voltammograms measured on copper electrode oxidation).
in citrate solutions containing zinc and tin ions separately and together. Fig. 3a shows a series of voltammograms recorded at different rota-
The same polarisation rate, equal to 20 mV/s, was used in all measure- tion speeds, in citrate electrolyte containing tin and zinc ions together.
ments. The potential scan was initiated in the negative direction from In general, it can be noted that the values of cathodic and anodic current
potential −0.5 V vs. SCE. The wave of tin reduction appears first at a po- densities increase together with the increase in the rotation rate of the
tential of about −0.7 V vs. SCE when only tin ions are present in citrate working electrode. The linear dependence of the cathodic current on
electrolyte. It forms a small peak (c1) with a maximum at a potential of the rotation rate of the working electrode, at potential −2.0 V vs. SCE,
− 0.74 V vs. SCE. However, under stationary conditions, continuous indicates that a diffusional controlled process takes place during the for-
deposition of tin does not occur in this range of potential. Hence it is mation of alloy (Fig. 3b). Also, the height of the tin oxidation peak (a2)
probably connected with the process of underpotential deposition of grows linearly with the increase of RDE speed. However, the intercept
tin on the copper substrate in such conditions, which leads to the forma- of the straight line does not pass through the origin of the plot, which
tion of only a monolayer of tin. Then the principal reduction of tin starts suggests a more complicated mechanism of the deposition process.

a2
6
c1 peak
5 8 7
|i| at E=-2.0 V
a2
modulus current density

6 a1
a4 5 a2
4 6
Current density (A/dm2)

|i| [A/dm ]
2
Current density (A/dm2)

4
3
3 4
a1 2

2 1
2 0
a3 0 1 2 3 4 5 6 7 8 9 a1
1 (rotation rate)1/2 [ (rad/s) 1/2]
0
0 c1
c1
-1 -2
c2
-2 -4
sol. no 1 "Na3Hcit + Zn" =16 rad/s
-3 -6
sol. no 2 "Na3Hcit + Sn" =40 rad/s
-4 =68 rad/s
sol. no 3 "Na3Hcit + Zn + Sn" -8
-5 -2,0 -1,8 -1,6 -1,4 -1,2 -1,0 -0,8 -0,6 -0,4
-2,0 -1,8 -1,6 -1,4 -1,2 -1,0 -0,8 -0,6 -0,4 Potential vs.SCE (V)
Potential vs.SCE (V)
Fig. 3. Cyclic voltammograms measured on a Cu substrate in solution no. 3, at various ro-
Fig. 2. Cyclic voltammograms measured on a Cu substrate in citrate solutions containing tation rates of working electrode. Scan rate = 20 mV/s, T = 20 °C. Arrows indicate scan
zinc and tin ions together, at ω = 16 rad/s, scan rate = 20 mV/s, T = 20 °C. Arrows indi- direction. The dependence of current density at selected points of voltammograms on
cate scan direction. the rotation rate of RDE is shown as an inset.
314 H. Kazimierczak et al. / Surface & Coatings Technology 240 (2014) 311–319

Moreover, the anodic part of voltammograms shows that the height of a 0


the peak a1, assigned to zinc oxidation, does not change with the in-
-1
crease of rotation rate. Hence the zinc deposition process is independent

i[A/dm ] H2
-2
of the speed at which the working electrode is rotated, which indicates "Zn+ Na3Hcit"

2
kinetic control of this process, in the applied range of potentials. -3
-4 "Sn+ Na3Hcit"
3.3. Steady state behaviour -5 "Sn+Zn+ Na3Hcit"
-6
For further study of tin–zinc alloy electrodeposition process, steady- 0
state partial polarisation curves for Sn(II), Zn(II) and hydrogen for their
-1
separate and co-discharge were made (Fig. 4a). They were compared

i[A/dm ] Sn
-2
with the dependence of the current efficiency (Fig. 4b) and the zinc

2
contents in deposits (Fig. 4c) on the electrode potential. The current -3
efficiency and partial polarisation curves were determined from the -4
chemical analysis of deposits electroplated under selected conditions, -5
the charge passed and the mass of deposits, using Faraday's law. The -6
separate reduction of zinc starts at potential −1.6 V vs. SCE. The partial
0
current of Zn reduction (modulus) increases consistently, while shifting
to more negative potentials, up to E = −1.9 V vs. SCE, where it reaches -1

i[A/dm ] Zn
a maximum. The current efficiency in this range of potential is 80%, -2

2
while the partial current of hydrogen reduction is relatively low at -3
potentials from − 1.6 V vs. SCE to − 1.8 V vs. SCE. Then it starts to -4
grow sharply at higher polarisations, which causes a significant decline -5
of current efficiency. Furthermore, a decrease in the zinc partial current
-6
density, starting from potential − 1.9 V vs. SCE, is strictly related to -2,0 -1,8 -1,6 -1,4 -1,2 -1,0
hydrogen co-evolution process which causes alkalisation of the area Potential vs.SCE (V)
close to the cathode surface. Subsequently, this causes the increase in
the concentration of inactive zinc citrate complex near the cathode. b 100
The partial polarisation curves for separate tin reduction can be divided
90
into three areas. The reduction of tin starts at potential −1.1 V vs. SCE,
Current efficiency (%)

80
but the rate of tin deposition in this case is very low, and partial current
density is only 0.03 A/dm2. Next, the value of Sn partial current in- 70
creases very slightly from 0.32 A/dm2 at −1.2 V vs. SCE to 0.63 A/dm2 60
at −1.5 V. With further potential shift towards more negative values, 50
the partial current of tin reduction increases significantly, reaching a 40
value of 4.75 A/dm2 at potential −2.1 V vs. SCE. Furthermore, the par- 30
"Zn+ Na3Hcit"
tial current of hydrogen co-evolution is relatively low over the whole "Sn+ Na3Hcit"
20
range of applied potentials in the case of separate tin deposition. Thus "Sn+Zn+ Na3Hcit"
10
it results in high current efficiency of the electrodeposition process, of
an average value of 95%. The partial polarisation curves of tin and zinc 0
-2,0 -1,8 -1,6 -1,4 -1,2 -1,0
codeposition show that tin deposition is significantly inhibited, com-
Potential vs.SCE (V)
pared to its separate reduction. The partial current of tin codeposition
increases gradually from 0.01 A/dm2 at −1 V vs. SCE to 1.17 A/dm2 at c 80
− 2.1 V vs. SCE, while zinc deposition is noticeably enhanced at the
range of potentials from − 1.0 V to − 1.8 V vs. SCE. It is interesting 70
that zinc codeposition is already underway at the potential of − 1 V 60
Zn content (%)

vs. SCE, although the zinc partial current is about 5 · 10− 5 A/dm2
50
then, and zinc content in the deposit reaches a value of only 0.3 wt.%.
The course of partial current of zinc codeposition can be divided into 40
three parts. At low polarisations (i.e. E ≥ − 1.3 V vs. SCE) it grows 30
very slightly up to 0.05 A/dm2 at −1.3 V vs. SCE. Next, at the intermedi-
ate polarisation range (with a potential of between −1.4 and −1.8 V vs. 20
SCE) the partial current of Zn reduction increases consistently when po- 10
tential is shifted to more negative values, up to E = − 1.8 V vs. SCE, "Sn+Zn+Na3Hcit"
0
where it reaches a maximum. Such polarisation curve characteristics
-2,0 -1,8 -1,6 -1,4 -1,2 -1,0
are very similar to those of zinc separate reduction. However, in the
Potential vs.SCE (V)
case of zinc codeposition, it starts and has a maximum at less negative
potentials, in comparison to the case of separate zinc deposition. The Fig. 4. (a) Steady state partial polarisation curves for Zn–Sn, Zn-separate and Sn-separate
effect of zinc codeposition enhancement results in the high content of deposition. (b) The dependence of the current efficiency on the potential of the working
zinc in the deposit. The content of zinc deposited at −1.5 V vs. SCE is electrode. (c) Zn content in tin–zinc alloy deposits as a function of potential of the working
70.4 wt.%, whereas the deposit obtained at − 1.8 V vs. SCE contains electrode. Electrolytes: 1–3, Q = 20C, ω = 16 rad/s, and T = 20 °C.
74.8 wt.%. The partial current of hydrogen coevolution during the
alloy formation is very low at potentials up to −1.7 V vs. SCE, where
it does not exceed the value of 0.3 A/dm2. Then, when shifting towards SCE. Consequently, the current efficiency of the tin–zinc alloy electrode-
more negative potentials, the partial current of hydrogen coevolution position process declines sharply from 94% at −1.7 V vs. SCE to 29% at
starts to increase sharply, reaching a value of 5.6 A/dm2 at −2.1 V vs. −2.1 V vs. SCE.
H. Kazimierczak et al. / Surface & Coatings Technology 240 (2014) 311–319 315

a 60 b 100
90
50

Current efficiency (%)


80

Zn content (wt.%)
40 70
60
30 50
40
20
30
10 20
16 rad/s 16 rad/s
10
68 rad/s 68 rad/s
0 0
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
Current density [A/dm2] Current density [A/dm2]

Fig. 5. The dependence of current density on the composition of deposited layers (a) and on the current efficiency of the electrodeposition process (b). Layers obtained from solution no. 3,
Q = 20C, T = 20 °C.

a 60
b 100
90
50

Current efficiency (%)


80
Zn content (wt.%)

40 70
60
30
50
20 40
2
i = 0.5 A/dm
30 2
10 i = 1 A/dm
20 2
i = 3 A/dm
0 10
2 3 4 5 6 7 8 9 2 3 4 5 6 7 8 9
1/2 1/2 1/2 1/2
(rotation rate) [ (rad/s) ] (rotation rate) [ (rad/s) ]

Fig. 6. Effect of rotating disc electrode speed on the composition of deposited layers (a) and the current efficiency of the electrodeposition process (b). Layers obtained from solution no. 3,
at various current densities which are indicated on the graph, Q = 20C, T = 20 °C.

3.4. Galvanostatic deposition decreases — first sharply, from 99% at current density 0.25 A/dm2 to
80% at 2 A/dm2. Then it remains at the constant level in the range of cur-
Fig. 5a,b shows the dependence of current density on the composi- rent densities in which obtained deposits contain from 40% to 50% Zn.
tion of deposited layers (a) and current efficiency of the electrodeposi- Next, when the current density increases up to 10 A/dm2, the current
tion process (b) for deposits obtained from bath no. 4 in different efficiency decreases constantly down to the value of 24%. In the case
hydrodynamic conditions. In the case of low rotation rate (16 rad/s), of high rotation rate (68 rad/s), the dependence of current density on
the content of zinc in Sn–Zn layers increases with the increase of current the composition of deposits can be divided into two parts. In the range
density from 0.5 to 2 A/dm2. Then it starts to decrease as current density of lower current densities, the content of zinc in layers rises from
increases to 6 A/dm2. Next, in the range of current densities from about 0.1 wt.% at 0.25 A/dm2 to 54.8 wt.% at 5 A/dm2. In the case of high rota-
7 to 10, the content of Zn in the deposit remains at almost the same tion rate, zinc content in deposits is considerably lower at the range of
level, while the current efficiency of the electrodeposition process current densities from 0.25 A/dm2 to 3 A/dm2, than when the low

70 100 80 100

60
80 80
current efficiency (%)

current efficiency (%)

60
Zn content (wt.%)
Zn content (wt.%)

50

40 60 60
40
30 40 40
20
20
20 20
10 Zn content Zn content
current efficiency current efficiency
0 0 0 0
0,25 0,35 0,45 0,55 0,65 0,02 0,04 0,06 0,08 0,10 0,12
Na3Hcit concentration (M) SnSO4 concentration (M)

Fig. 7. Effect of sodium citrate concentration in electrolyte on the composition of deposited Fig. 8. Effect of tin sulphate concentration in the electrolyte on the composition of depos-
layers and the current efficiency of the electrodeposition process. Layers obtained from ited layers and the current efficiency of the electrodeposition process. Layers obtained
electrolyte nos. 3–7, i = 3 A/dm2, ω = 16 rad/s, Q = 20C, T = 20 °C. from electrolytes: 3 and 8–12; i = 3 A/dm2, ω = 16 rad/s, Q = 20C, T = 20 °C.
316 H. Kazimierczak et al. / Surface & Coatings Technology 240 (2014) 311–319

70 100 the electrodeposition process. The effect of the hydrodynamic condi-


60 tions on the electrodeposition process has been studied in more detail.
80

current efficiency (%)


Zn content (wt.%)

50 3.4.1. Influence of hydrodynamic conditions


60 Fig. 6 shows the dependence of the speed of the rotating disc elec-
40
trode on the composition of deposits (a) and current efficiency of the
30 electrodeposition process (b) at three selected current densities. In the
40
case of the lowest current density under consideration, i = 0.5 A/dm2,
20 increasing the RDE speed causes a decrease of zinc content in deposits,
20 while the current efficiency is constant over the whole range of applied
10 Zn content
rotation rates. In the case of i = 1 A/dm2, increasing the RDE speed is
current efficiency
0 0 connected with a decrease of zinc content in deposits, with a simul-
0,08 0,10 0,12 0,14 0,16 0,18 0,20 taneous, slight increase in current efficiency. However the current
ZnSO4 concentration (M) efficiency is maintained at a relatively high level (from about 80 to
95%) in the whole range of rotation rates. When current density
Fig. 9. Effect of zinc sulphate concentration in the electrolyte on the composition of depos- i = 3 A/dm2 is applied, the zinc increases significantly from 33 wt.%
ited layers and the current efficiency of the electrodeposition process. Layers obtained
at 6 rad/s to 55 wt.% at 27 rad/s and then it starts to decrease very
from electrolytes: 3 and 13–16; i = 3 A/dm2, ω = 16 rad/s, Q = 20C, T = 20 °C.
slightly with the further increase of RDE speed, reaching a value of
rotation rate was applied. Then it reaches almost the same value, for 50% at 68 rad/s. In this instance the current efficiency of the electrode-
both rotation rates, at 3 A/dm2. Next the content of zinc in deposits position process does not exceed 20% at the lowest rotation rate used.
starts to decrease constantly with the increase of current density At 16 rad/s it increases rapidly to the value of about 80% and then at
from 6 A/dm2 to 10 A/dm2. The value of current efficiency of the elec- higher RDE speed it stays at this constant and relatively high level.
trodeposition process at high rotation rate does not differ significantly This sharp decrease of current efficiency at low RDE speed is associated
at the current densities from 0.25 to 3 A/dm2. It decreases sharply with the differences in control of the mechanism of hydrogen and metal
from 99% at 0.25 A/dm2 to 85% at 2 A/dm2, and then it stays almost in- reduction. At high current densities, the hydrogen evolution process is
variable with the further increase of current density applied. The above- kinetic-controlled, while deposition processes of metals are diffusion-
noted effects are the result of limitations of mass transport of tin–citrate controlled. Therefore, decreasing the RDE speed causes a decrease in
complexes to the surface of the electrode (diffusion control), at low cur- the deposition rate of metals. This leads to an increase in the hydrogen
rent densities (i.e. i ≤ 2 A/dm2). Then, the zinc deposition process is evolution rate in relation to the deposition of alloy rate, and conse-
controlled by the rate of the reduction process of zinc-citrate complexes quently to a decline in current efficiency of the electrodeposition pro-
(kinetic control). Therefore, in this range of current densities, an in- cess. These effects may also be due to the fact that, at current densities
crease of rotation rate results in an increase of tin content in the alloy. equal or greater than 3 A/dm2, hydrogen evolution takes place during
At higher current densities (i.e. 2 b i b 4 A/dm2) the zinc deposition the formation of the alloy. Then, at low RDE speed (ω b 16 rad/s),
process is characterised by mixed kinetic-diffusion control, which hydrogen bubbles are blocking the cathode surface. Therefore the actual
next becomes diffusion control (i.e. i ≥ 4 A/dm2). This results in the in- surface area on which further electrodeposition of the alloy takes place
version of the dependence of RDE speed on the electrodeposition of the is smaller. Hence the real current density is higher, which results in an
alloy. Thus, the higher the RDE speed, the higher the zinc content in the increased hydrogen evolution rate and a decline in current efficiency.
alloy, whereas the current efficiency of electrodeposition is associated
with the hydrogen co-evolution process. At low current densities, the 3.4.2. Influence of bath composition
hydrogen evolution rate is low, hence current efficiency is high. When The effect of sodium citrate concentration in the electrolyte is
the current density increases, the process of hydrogen evolution starts presented in Fig. 7. The concentration of Na3Hcit was varied from
to occur as a result of the reduction of protonated forms of citrate ion. 0.25 M to 0.65 M. Deposits were obtained at the same current density
This process is diffusion-controlled, hence hydrogen evolution depends (i = 3 A/dm2) and the same rotation rate (ω = 16 rad/s). Generally,
on the hydrodynamic conditions at this range of current densities. At the Zn content in coatings decreases significantly with the increase of
high current densities, the dominant process of hydrogen evolution oc- sodium citrate concentration from 50 wt.% at 0.25 M Na3Hcit to
curs due to the reduction of the solvent (water). This process is inde- 2.4 wt.% at 0.65 M NH3Hcit. These changes in the composition of
pendent of the hydrodynamic conditions. Hence the increase of RDE deposits are directly related to the decrease in current efficiency of elec-
speed favours deposition of alloy and increase of current efficiency of trodeposition process, which decreases from 82% at 0.25 M NH3Hcit to

a 60 b 100
50
Current efficiency (%)

80
Zn content (wt.%)

40
60
30
40
20
2
i = 0.5 A/dm
10 20 i = 1 A/dm
2

2
i = 3 A/dm
0 0
0 10 20 30 40 50 0 10 20 30 40 50
Charge (C) Charge (C)

Fig. 10. Effect of charge passed on the composition of deposited layers (a) and the current efficiency of the electrodeposition process (b). Layers obtained from electrolyte no. 3,
i = 2 A/dm2 ω = 16 rad/s, T = 20 °C.
H. Kazimierczak et al. / Surface & Coatings Technology 240 (2014) 311–319 317

22% at 0.65 M NH3Hcit. This can be explained by the fact that the a) i= 1 A/dm2, 12.4 wt% Zn
increase of the sodium citrate concentration in the electrolyte causes
the formation of non-electroactive citrate complexes of zinc [26]. It
does not change the type of tin-citrate complexes formed in the electro-
lyte, which are electroactive in the whole range of sodium citrate
concentration used. The effect of the tin sulphate concentration on the
electrodeposition process is presented in Fig. 8. The concentration of
tin sulphate was varied in the range of 0.02 M to 0.12 M. Sodium citrate
concentration was 0.25 M, and zinc sulphate concentration was 0.16 M.
Deposits were obtained under the same conditions (i = 3 A/dm2,
ω = 16 rad/s). It is shown that zinc content in deposits decreases
with the increase of tin sulphate concentration in the electrolyte. Also,
the current efficiency increases slightly with the increase of sodium
citrate concentration in the electrolytic bath. Therefore it has been
shown that a coating containing 42.6 wt.% of Zn and 57.4 wt.% of Sn
can be obtained in a process with a very high current efficiency of b) i= 1.5 A/dm2, 24.3 wt% Zn
about 91%, from the bath containing 0.12 M SnSO4 (bath no. 12). The
effect of zinc sulphate concentration on the electrodeposition is similar
to that of tin sulphate concentration variations, i.e. the increase of
concentration of certain metal ions in the electrolyte causes the increase
of this metal content in the alloy electrodeposited (Fig. 9). The content
of zinc in the deposit increases consistently with the increase of zinc
ions in the electrolyte, and the increase of current efficiency is observed
simultaneously.

3.4.3. Effect of total charge passed


The effect of the value of the total charge passed during electrodepo-
sition of Sn–Zn alloy layers was studied (Fig. 10). Deposits were obtain-
ed from bath no. 3, at various current densities and the same rotation
rate (ω = 16 rad/s). The value of the charge passed was varied from
5C to 50C. It can be noticed that the composition of the alloy is stable
c) i= 2 A/dm2, 36.9 wt% Zn
in the range of applied charge from 5 to 20C (which is equivalent to
the layer thickness from 1 to 3 μm). However, a further increase of
applied charge results in a very slight decrease of zinc content in
the alloys. The biggest variations in the composition of the alloy are
observed in the case of the highest current efficiency under consider-
ation (i = 3 A/dm2). Accordingly, the layer obtained at Q = 20C con-
tains 48.3% of Zn, while the alloy layer containing 42.2 wt.% of Zn is
formed when applying Q = 50C (then the layer thickness is about
10 μm). When lower current densities are applied, the differences in
the alloy composition are slightly smaller, maintained in the range of
about 3%.

3.5. Characterisation of coatings


d) i= 4 A/dm2, 52.7 wt.% Zn
The SEM images of Sn–Zn coatings are presented in Fig. 11. It shows
the morphological characteristics of layers deposited on steel substrate,
from bath no. 3, under the same hydrodynamic conditions but different
current densities. The studied layers were bright and shiny. The varia-
tions in current densities cause adequate changes in the composition
of alloys, and also influence the deposit's morphology. Each layer
analysed is characterised by a granular structure. However, the surface
structure is more fine-grained in the case of deposits obtained at current
densities up to 1.5 A/dm2 (Fig. 11a,b). The layer deposited at 2 A/dm2
consists of agglomerates of bigger granules, in comparison to the previ-
ously mentioned samples, whereas the deposit obtained at 4 A/dm2
shows a more irregular structure in which relatively large single gran-
ules are on the surface of a layer built of more fine-grained nodules.
Next, Fig. 12 shows the cross section of chosen Sn–Zn coating on steel,
with marked points where the chemical analysis was made (Table 2). Fig. 11. SEM images (SE) of Sn–Zn coatings with various contents of zinc, deposited
galvanostatically on steel substrate. Electrolyte no. 3, Q = 20C, ω = 68 rad/s, T = 20 °C.
It can be noted that the tin–zinc coating consists of some small dark par-
ticles embedded in a compact and bright matrix. Due to the very fine
grain size these particles have not been identified precisely. The on the cross section of coating are observed. Certainly, it can be noted
presence of such small particles (significantly less than 1 μm) and a that the bright matrix phase is enriched in tin, and zinc content in it
limited resolution of EDS microanalysis method affect the results. varies from 3.5 wt.% to about 8 wt.%. Nevertheless, the chemical analy-
Therefore, considerable differences in the composition at various points sis of the dark particles, which are relatively bigger than the majority of
318 H. Kazimierczak et al. / Surface & Coatings Technology 240 (2014) 311–319

it can be concluded that the Sn–Zn coating studied consists of fine grains
of Zn-rich phase uniformly arranged within the phase enriched in Sn.
The XRD analysis of Sn–Zn coatings, of various compositions, is
presented in Fig. 13. These are the same coatings that were analysed
by SEM. Each diffraction pattern shows peaks corresponding to the hex-
agonal Zn phase and tetragonal β-Sn phase, thus revealing that electro-
deposited layers consist of a mixture of these phases. Also, the peaks
corresponding to iron are apparent on the diffractograms. They origi-
nate from the steel substrate and, because of varied thickness of coat-
ings, they appear with varying intensities. Furthermore, changes in the
intensity of certain reflexes are disproportionate in relation to the vari-
ations in the composition of the alloy. This implies that significant
changes in texture of deposits occur while increasing the current densi-
ty applied. However, both of the abovementioned problems require
Fig. 12. SEM image (BSE) of cross section of Sn–Zn12% coating. Marked points indicate
places where EDS microanalysis was made. further detailed studies.

4. Conclusions
Table 2
Results of EDS microanalysis of chosen points of Zn–Sn coating cross section (marked at
Fig. 11).
• Thermodynamic analysis of Sn(II)–Zn(II)-citrate baths models indi-
Point number Fe [wt.%] Zn [wt.%] Sn [wt.%] cates that the optimal conditions enabling the preparation of stable
1 1.1 6.2 92.7 electrolytic baths are in the relatively narrow pH range of about 5.
2 0.3 3.5 96.2 • The composition of the alloy layers can be controlled easily by varying
3 0.8 8.4 90.8 the metal ion concentration in the electrolyte.
4 0.7 4.2 95.1
• The presence of zinc ions in the electrolyte results in a significant inhi-
5 0.7 51.2 48
6 11.4 4 84.6 bition of tin reduction rate, which enables the introduction of relative-
7 99.3 0.4 0.3 ly large amounts of zinc in the tin–zinc alloy.
8 0.8 15.9 83.3 • The alloy layers containing from about 0.5 to 75 wt.% Zn can be elec-
9 0 4.0 96.2 trodeposited with high current efficiency, within the range from 80%
to almost 100%.
• The morphology of the surfaces of the electrodeposited layers varies
noticeably with the increase of applied current density. The morphol-
it (i.e. points 5 and 8) indicates that this phase contains much higher ogy of surface layers seems to be more fine-grained when low current
amounts of zinc than the bright phase. In the case of point 5 it is more densities are applied (up to about 1.5 A/dm2).
than 50% of zinc. The analysis of point 8 indicates that there is about • Coatings obtained at current densities in the range of 0.5 to 4 A/dm2
15.9 wt.% of zinc. However in that instance the dark particle was are bright and shiny, whereas those obtained when higher current
smaller than that represented by point 5, hence the analysed area was density is applied appear dull grey with pores on the surface, which
surrounded by the bright phase, which may affect the result. Generally, is related to hydrogen coevolution.
Sn (200)

Sn (112)
Sn (220) Zn (101)

Zn (103)
Zn (110)

Zn (112)
Sn (400)

Sn (420)
Zn (102)

Zn (200)
Zn (100)
Zn (002)

Sn(321)

Zn (201)
Sn (211)
Sn (101)

Sn (501)
Sn(312)
Sn(411)
Intensity (counts)

Sn (301)

Fe (211)
Fe (200)

(d)

(c)

(b)

(a)

36.261 40 44,899 55.340 64.580 73.149 86.496


32.010 43.868 54.319 60 63.782 72.395 80 83.736
70.627 82.043 89.338
30.641 38.990 43.216 62.512
70.034 79.455
2 theta (degrees)

Fig. 13. X-ray diffraction patterns of Sn–Zn alloys electrodeposited at current density: (a) 1 A/dm2, (b) 1.5 A/dm2, (c) 2 A/dm 2 and (d) 4 A/dm2. Electrolyte no. 3, Q = 20C, ω = 68 rad/s,
T = 20 °C.
H. Kazimierczak et al. / Surface & Coatings Technology 240 (2014) 311–319 319

• Two phases are formed in deposits: a hexagonal zinc phase and a [12] M.M. Thompson, J.C. Patten, U.S. Patent 1876156 (1932).
[13] P. Marino, British Patent, 405433 (1932).
tetragonal β-tin phase. [14] Mead Research Engineering Co., W. Sailer, British Patent, 407670 (1934).
[15] J.W. Cuthbertson, R.M. Angles, J. Electrochem. Soc. 94 (2) (1948) 73.
Acknowledgments [16] O.A. Ashiru, J. Shirokoff, Appl. Surf. Sci. 103 (1996) 159–169.
[17] V.S. Vasantha, M. Pushpavanam, P. Kamaraj, V.S. Muralidharan, Trans. Inst. Met.
Finish. 74 (1) (1996) 28–32.
The authors would like to thank for the SEM analysis Dr Tomasz [18] St. Vitkova, V. Ivanova, G. Raichevsky, Surf. Coat. Technol. 82 (1996) 226–231.
Tokarski from AGH University of Science and Technology, Faculty of [19] M. An, Y. Zhang, J. Zhang, Z. Yang, Z. Tu, Plat. Surf. Finish. (May 1999) 130–132.
[20] L. Sziraki, A. Cziraki, Z. Vertesy, L. Kiss, V. Ivanova, G. Raichevski, S. Vitkova, S.
Non-Ferrous Metals, Krakow, Poland and Dr. Elzbieta Bielanska from Marinova, T. Marinova, J. Appl. Electrochem. 29 (1999) 927.
Jerzy Haber Institute of Catalysis and Surface Chemistry, Polish Academy [21] S.M. Abdel-Wahab, E. Mohamed, S.M. Rashwan, METALL 54 (2000) 268.
of Sciences, Krakow, Poland. [22] K. Wang, H.W. Pickering, K.G. Weil, Electrochim. Acta 46 (2001) 3835–3840.
[23] E. Guaus, J. Torrent-Burgues, J. Electroanal. Chem. 549 (2003) 25–36.
This work was supported by projects: POKL-04.01.01-00-004/10 and
[24] E. Guaus, J. Torrent-Burgues, J. Electroanal. Chem. 575 (2005) 301–309.
POIG-01.01.02-00-015/09-00, [25] C.C. Hu, C.K. Wang, G.L. Lee, Electrochim. Acta 51 (2006) 3692–3698.
[26] P. Ozga, Polska Metalurgia w latach 2006–2010, Akapi, Kraków, ISBN:
978-83-60958-64-3, 2010, pp. 138–147, (in Polish).
References [27] H. Kazimierczak, P. Ozga, Surf. Sci. 607 (2013) 33.
[28] Open-File Report 91–183, Database WATEQ4F, U.S. GEOLOGICAL SURVEY, 2001.
[1] S.J. Blunden, Adv. Mater. Processes (December 1991) 37–39. [29] A.E. Martell, R.M. Smith, Critical Stability Constants, vol. 1–6Plenum Press, New
[2] E. Budman, M. McCoy, Met. Finish. 9 (1995) 10. York, 1974–189.
[3] E. Budman, D. Stevens, Trans. Inst. Met. Finish. 76 (3) (1998) B34. [30] P. Ozga, The Role of Complexation on the Electrolytic Deposition of Metals and
[4] S. Vaynman, G. Ghosh, M.E. Fine, Mater. Trans. 45 (2004) 630–636. Alloys from Citrate Solutions IMIM PAN, Kraków, , 2006, ISBN 83-921845-8-0.
[5] J. Villain, W. Jillick, E. Schmitt, T. Qasim, Proc. of the 2004 Int. IEEE Confer. on the [31] L.D. Pettit, K.L. Powell, SC-Database for Windows, Academic Software, Version 4.79,
Asian Green Electronics, AGEC, 2004, pp. 38–41. Sourby Old Farm, Timble, Otley, Yorks, UK.
[6] S. Dubent, M. De Petris-Wery, M. Saurat, H.F. Ayedi, Mater. Chem. Phys. 104 (2007) [32] SCDatabase, IUPAC Stability Constants Database, Release 5.8 for Windows, Academic
146–152. Software, 2010.
[7] V.C. Opaskar, L.D. Capper, US Patent 6436269 (2002). [33] P. Ozga, Z. Świątek, A. Dębski, J. Bonarski, L. Tarkowski, E. Bielańska, P. Handzlik, B.
[8] E.W. Brooman, Met. Finish. 4 (2000) 42–50. Onderka, M. Michalec, The Layers and the Protective Coatings on the Basis of Zinc
[9] M. Antelman, Chemical Electrode Potentials, Plenum Press, New York, 1982. Alloys with Ferrous Metals and Manganese for Replacing Cadmium Coatings
[10] P. Marino, British Patent, 10133 (1915). Obtained by Electrodepositing from Complex Solutions, IMN Gliwice, 2010, ISBN
[11] B.R. Haueisen, J.C. Patten, U.S. Patent 1904732 (1930). 978-83-925546-6-0. 295–306.

You might also like