Sie sind auf Seite 1von 12

Electrochimica Acta 141 (2014) 149–160

Contents lists available at ScienceDirect

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

Electrodeposition of Sn-Zn-Cu alloys from citrate solutions


M. Slupska, P. Ozga ∗,1
Institute of Metallurgy and Material Science, Polish Academy of Sciences, 30-059 Krakow, Reymonta 25, Poland

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

Article history: The first stage of this study involves the development of stable baths for electrodepositing Sn-Zn-Cu
Received 17 April 2014 alloys; these developments are based on thermodynamic models of citrate baths and experiments explor-
Received in revised form 1 July 2014 ing the stability of the baths. The effects of the sodium citrate (complexation agent) concentration and
Accepted 2 July 2014
the pH of the solution on the stability of the baths were examined experimentally. The stability of the
Available online 17 July 2014
baths was determined through spectrophotometric analysis. Stable baths designed for the electrodepo-
sition of Sn-Zn-Cu alloys were produced in the range of dominant citrate complexes with highly negative
Keywords:
charge (the reduction process is strongly inhibited by high activation energy). Voltammetric studies and
electrodeposition
Sn-Zn-Cu alloys potentiostatic deposition were conducted to analyse the co-deposition of tin, zinc, and copper, and the
lead-free solders co-evolution of hydrogen. The effect of the solution pH, the concentration of sodium citrate and the hydro-
dynamic conditions on the electrodeposition process and the composition of the deposits were examined.
The deposits were analysed using chemical (␮XRF) as also grain size and phase analysis (X-ray diffrac-
tion). The possibility of electrodepositing Sn-Zn-Cu alloys from citrate solutions was confirmed. The tin
content of these coatings varied from 60 to 96 wt.%, the copper content varied from 3.5 to 20.5 wt.%, and
the zinc content reached up to approximately 37.5 wt.%.
© 2014 Elsevier Ltd. All rights reserved.

1. Introduction as protective layers against corrosion, specifically to replace cad-


mium, which was more expensive at the time [4–9]. Copper, silver,
Sn-Zn-Cu alloys are of great interest because they may be indium, bismuth and several other elements can be used as the
used industrially as lead-free solders, in kesterite-based solar cells, third and fourth component in these alloys to improve the solde-
because Cu2 ZnSn(S,Se)4 is an intermediate product during the syn- ring and mechanical properties of the Sn-Zn alloys [2,10]. Adding
thesis of kesterites, as negative electrodes in lithium-ion batteries, a small amount of Cu can improve the flexural strength, corro-
and as corrosion resistant layers (combining the properties of brass sion resistance and reduce the dezincification of these alloys during
and bronze). The composition of the alloys varies between appli- corrosion processes [3,10–12]. Copper also improves the electrical
cations. SAC (tin-silver-copper alloys) and eutectic Sn-Ag alloys properties [2].
have been used to replace solders containing lead for many years. Sn-Zn-Cu alloys combined with S or Se are also promising can-
These materials have good physicochemical properties, but they are didates as layers in thin films (CZTS - Cu2 ZnSnS4 and CZTSe -
difficult to obtain through electrodeposition or electroless deposi- Cu2 ZnSnSe4 ) due to their excellent mechanical properties, large
tion methods; therefore, tin is currently the primary material used absorption coefficients, suitable band gap energies and efficient
for the electrodeposition of solder layers [1]. Sn-Zn eutectic alloys production [13–21]. Therefore, these materials may replace thin
are among the most attractive lead-free solder alloys [2,3]. These film solar cells containing indium and gallium, which are more
next-generation materials can easily be obtained through elec- expensive.
trodeposition [4–9]. Sn-Zn eutectic alloys have good mechanical Electrodeposition may be the best way to obtain Sn-Zn-Cu alloys
properties, low melting points (198.5 ◦ C), and relatively low costs. because this inexpensive and environmentally friendly method can
Sn-Zn alloys were developed at the beginning of the 20th century be carried out at room temperature. Knowledge regarding the pro-
duction of Sn-Zn-Cu ternary alloys as lead-free solders is limited.
Few authors have generated Sn-Zn-Cu layers through other means,
such as electroless plating [22] and combinatorial, two stage meth-
∗ Corresponding author. Tel.: +48122952818; fax: +48122952804.
ods. These stages are (i) the electrodeposition of a compositionally
E-mail address: p.ozga@imim.pl (P. Ozga).
varied, binary alloy film (Sn-Zn), and (ii) dripped immersion
1
ISE member. plating (Sn-Zn-Cu) [23]. A Sn-Zn-Cu alloy was obtained by

http://dx.doi.org/10.1016/j.electacta.2014.07.039
0013-4686/© 2014 Elsevier Ltd. All rights reserved.
150 M. Slupska, P. Ozga / Electrochimica Acta 141 (2014) 149–160

Table 1 (SCE). The chemical composition of the electrodeposits on the steel


Chemical composition of electrolytic baths used.
substrate was characterised via ␮-XRF analysis (Bruker Micro X-ray
Electrolyte composition + 0.5 g dm−3 PEG1 +0.05 g dm−3 SDS2 Fluorescence). The partial polarisation curves for the deposition of
the individual metals were obtained using partial currents; these
SnSO4 ZnSO4 CuSO4 Na3 HCit3 pH
mol dm−3 mol dm−3 mol dm−3 mol dm−3 partial currents were calculated from Faraday’s law, based on the
passed electrical charge, the mass of the deposit generated at a con-
1 0.65 5.5
2 0.25 0.65 5.5 stant potential, and chemical analysis. The partial current for the
3 0.1 0.65 5.5 evolution of hydrogen was calculated using the difference between
4 0.01 0.65 5.5 the total charge and the quantity of charges consumed during metal
5 0.1 0.25 0.65 5.5 deposition. Increases or decreases in the current and partial cur-
6 0.1 0.25 0.01 0.65 5.5
7 0.01 0.65 5.5
rent density values, which are discussed in the present study, refer
8 0.1 0.25 0.01 0.4 5.5 to the modulus. The mechanism of the electrodeposition was also
9 0.1 0.25 0.01 0.65 5.5 subjected to voltammetric analysis. The working electrodes in the
10 0.1 0.25 0.01 0.65 5 voltammetric study were copper discs, chemically polished using a
11 0.1 0.25 0.01 0.4 5
mixture of concentrated nitric, acetic and phosphoric acids (1:1:1)
12 0.1 0.25 0.01 0.8 5.5
13 0.1 0.25 0.01 0.65 5.75 at ambient temperatures. X-ray structural studies were performed
14 0.1 0.25 0.01 0.8 5.75 using a polycrystalline diffractometer with CuK␣ radiation (Philips
15 0.1 0.25 0.01 0.65 5.25 X’Pert Pro). The phase composition of the precipitates received
1
PEG - polyethylene glycol 3000, 2 SDS - dodecyl sulphate sodium salt, 3 Cit - C6 H4 O7 from the baths, their stability, and the phase composition of the
electrodeposits, were analysed using the PDF4 crystallographic
databases [24]. X-ray line broadening (B), after correcting for
electrodeposition (single step [13,16] or two step electrodeposition instrumental broadening, was used to investigate the crystal grain
processes [21]) to produce solar cells. size using the Scherrer formula: D = (k•␭)/(B•cos␪), where D is the
The major problem regarding the electrodeposition of Sn-Zn- crystallite size, k is the shape factor (assumed spherical = 0.94), ␭–is
Cu alloys is the development of stable baths. The simultaneous the wavelength of the x-rays (1.54060 Å) and ␪ is the Bragg angle.
presence of Sn(II) and Cu(II) species in a simple electrolytic bath
facilitates the reduction of Cu (II) to metallic copper by Sn(II).
Choosing a proper complexation agent may stabilise the elec- 3. Results and discussion
trolytic baths by changing the dominant tin(II) and copper(II)
species. Therefore, (i) the reduction of Cu (II) can be inhibited 3.1. Development of stable Sn (II)-Zn(II)-Cu(II)-Cit baths
(kinetic effect) or (ii) the electrochemical potential of the dominant
species can be changed (thermodynamic effect) [1]. The preparation of stable electrolyte baths is a major prob-
The main purpose of this work was to develop stable baths for lem during the electrodeposition of Sn-Zn-Cu ternary alloys. These
the electrodeposition of Sn-Zn-Cu alloys; these developments were baths must contain a mixture of tin(II), zinc(II) and copper(II)
based on thermodynamic models of the Sn(II)-Zn(II)-Cu(II)-citrate species without precipitates. The dominant aqueous metal species
baths and experimental investigations describing their stability. should be electroactive and reducible on the electrode surface.
Subsequently, the mechanism and kinetics of the electrodeposition Citrate electrolytes are very attractive because citrate ions form
of Sn-Zn-Cu alloy were analysed. Electrodeposition experiments various electroactive aqueous complexes with Sn(II), Zn(II) and
were performed using several stable Sn(II)-Zn(II)-Cu(II)-citrate Cu(II), and because citrate baths are environmentally friendly. Ther-
baths to confirm that the desired Sn-Zn-Cu deposits could be modynamic models based on stability constants are very useful
obtained. when preparing stable baths. The major sources of stability con-
stant data are databases, including those by Sillen and Martel
2. Experimental [25] or the IUPAC stability Constants Database (SC-Database) [26].
These data often require corrections for the conditional stability
The chemical compositions of the solutions used while studying constants, based on pH-metric studies in systems with similar com-
the stability of the baths and taking electrochemical measure- positions, concentrations and ionic strengths to analysed systems
ments are given in Table 1. The optimal ranges for the pH and [27–30]. Predominance area diagrams for Sn(II), Zn(II) and Cu(II)
concentration of the complexation agent were chosen based on species in a mixture of 0.01 M CuSO4 , 0.1 M SnSO4 and 0.25 M
the analysis presented in sub-Section 3.1. All chemicals used in ZnSO4 , depending on the concentration of the complexation agent
this work were analytical grade. Deionised water (ultrapure, resis- (sodium citrate) and the pH of the solution, are shown in Fig. 1, 2a
tivity 18.2 M•cm, produced by a Milllipore Direct-Q 3 system) and 3. The results reveal the optimal range of sodium citrate con-
was used to prepare the solutions and to rinse samples containing centrations and pH values. The pH for copper(II) as the dominant
electrodeposits. The solution pH was adjusted by adding sulphuric electroactive complex species ranges from pH = 4 to 6; for tin(II),
acid or sodium hydroxide. Spectrophotometric analysis was carried this value ranges from pH = 4 to 7, while for zinc(II), it ranges from
out using a Jasco V-630 UV-VIS spectrophotometer with 10 mm pH = 3.5 to 5.5. The rest of the areas contain the following species:
quartz cuvettes (Spectrosil). All of the electrochemical measure- (i) non-electroactive citrate complexes (orange areas on the fig-
ments were carried out in a 200 cm3 cell while using a system with a ures), (ii) neutral citrate complexes, although these areas can also
rotating disc electrode to ensure that the hydrodynamic conditions be non-homogeneous because polymeric species from the neutral
were constant and controlled. The surface area of the work- citrate complex may form over several days, several weeks or even
ing electrode was 2.83 cm2 . The measurements were performed several months (light yellow areas on the figures), (iii) precipitated
in a three-electrode cell with a potentiostat (Princeton Applied hydroxyoxides and hydroxysulfates that are formed immediately,
Research PARSTAT 2273). The working electrodes were composed and these areas are also non-homogeneous (yellow areas on the
of low-carbon steel and were chemically polished using a solution figures). The optimal concentration of sodium citrate in the solu-
containing oxalic acid and 30% hydrogen peroxide (perhydrol). A tion ranges from 0.3 to 1 mol/dm3 . The concentration of tin(II) in
tin counter electrode (5 cm2 ) was also used. The working elec- the bath is limited by the precipitation of tin(II) hydroxide (Fig. 2b)
trode potentials were referenced to the saturated calomel electrode within the optimal range of pH values for Zn(II) and Cu(II) (4 to
M. Slupska, P. Ozga / Electrochimica Acta 141 (2014) 149–160 151

Fig. 1. Predominance area diagram for Cu(II) species and precipitates in citrate solu- Fig. 3. Predominance area diagram for Zn(II) species and precipitates in citrate solu-
tions (solution composition: 0.1 M SnSO4 , 0.25 M ZnSO4 , 0.01 M CuSO4 ; T = 20 ◦ C). tions (solution composition: 0.1 M SnSO4 , 0.25 M ZnSO4 , 0.01 M CuSO4 ; T = 20 ◦ C).

6). The dominant electroactive complex forms in the selected pH


range are as follows: Cu2 HCit2 3− and Cu2 Cit2 4− for Cu(II) (Fig. 1),
SnCit2− for Sn(II) (Fig. 2) and ZnHCit− for Zn(II) (Fig. 3). The relation-
ship between the potential and pH for a solution containing 0.01 M
CuSO4 , 0.1 M SnSO4 , 0.25 M ZnSO4 and 0.65 M Na3 HCit is shown in
Fig. 4. The area in which Cu(II) reduction may begin is marked in red.
Cu(II) can be reduced by Sn(II) in this section; therefore, the bath can
only be stabilised by kinetic inhibition of the reduction process (the
area denoted by Cu(s) in Fig. 4). The areas denoted by Cu(s) + Sn(s) ,
and Cu(s) + Sn(s) +Zn(s) are where Cu(II) and Sn(II), as well as Cu(II),
Sn(II) and Zn(II), can be reduced to form Cu-Sn or Cu-Zn-Sn alloys.
This last area is a theoretically optimal range of parameters for the
electrodeposition of Sn-Zn-Cu ternary alloys. Due to the overpoten-
tial, the real potential range for the electrodeposition of Sn-Zn-Cu
alloys should be moved toward lower potentials. The equilibrium

Fig. 4. Potential (E) -pH diagram for solution 0.1 M SnSO4 , 0.25 M ZnSO4 , 0.01 M
CuSO4 , 0.65 M Na3 HCit, (for simplicity, zinc(II) aqueous species and compounds are
Fig. 2. Predominance area diagram for Sn(II) species and precipitates in citrate
omitted, dotted lines (short marks) divide Sn(II) species areas, solid lines divide
solutions: 2a) solution composition: 0.1 M SnSO4 , 0.25 M ZnSO4 , 0.01 M CuSO4 ; 2b)
Cu(II) species areas, and dotted lines (long marks) divide the upper/lower stability
solution composition: 0.25 M ZnSO4 , 0.01 M CuSO4 , 0.65 M Na3 HCit; T = 20 ◦ C.
limit of water, T = 20 ◦ C).
152 M. Slupska, P. Ozga / Electrochimica Acta 141 (2014) 149–160

lines in Fig. 4 for the reduction processes with metals(II) (Cu(II)/Cu, reduction of a dominate protonated citrate ion, according to
Sn(II)/Sn, Zn(II)/Zn) are calculated under the assumption that the reaction R1 when the current is below -1.5 V vs SCE:
species are reduced to form pure metallic phases (Cu, Sn and Zn).
This assumption is accurate for the Zn-Sn system, but the Cu-Sn- 2H2 Cit 2− + 2e− → H2 + 2HCit 3− (R1)
Zn system also contains intermetallic phases and solid solutions,
Further, a very sharp increase of cathodic current density (below
slightly altering its equilibrium lines.
-1.75 V vs SCE) occurs due to the evolution of hydrogen and is
The compositions of the solutions given in Table 1 were assigned
related to the decomposition of water:
based on an analysis of the thermodynamic models. The baths were
clear and blue after preparation; no deposits were observed in the 2H2 O + 2e− → H2 + 2OH − (R2)
solution. The stability of the baths was examined spectrophoto-
metrically at specific intervals. The absorbance of the baths was The reduction of zinc(II) begins at approximately -1.4 V vs SCE
measured from 200 to 1100 nm. The most interesting results are (Fig. 6, curve 2). Peak (a) indicates that the oxidation of zinc begins
shown in Fig. 5. The intense peak between 1000 and 550 nm is at approximately -1.1 V vs SCE. The beginning of the tin (II) reduc-
attributed to copper (II) species (blue colour of solutions). Fig. 5 tion occurs at approximately -0.8 V vs SCE (Fig. 6, curve 3), while
a, b and c show the relationship between the concentration of the oxidation of tin starts at a potential of about -0.7 V vs SCE (peak b).
complexation agent and the stability of the bath in solutions 8, 9 The reduction of copper(II) starts at a potential of about -0.4 V vs
and 6 (Table 1). The baths become more stable when the sodium SCE (Fig. 6, curve 4). Oxidation of copper is invisible in the voltam-
citrate content is increased. The solution is decolourised when the mogram because this process begins at high potentials, which
concentration of the complexation agent is low. The intensity of may oxidise the copper electrode (substrate in the voltammetric
the peak decreased every day (Fig. 5a), and a red precipitate (cop- study). All of the reduction processes for the metal(II) species are
per powder) is observed in this solution. X-ray diffraction analysis shifted toward lower potentials relative to the equilibrium poten-
confirmed the presence of only copper in the precipitate (JCPDS tials (Fig. 4), which enables the reduction of complex ions with high
851326 [24]). Fig. 5c shows results for the solutions that con- cathodic overpotentials (0.3 to 0.5 V).
tain the highest sodium citrate concentration (0.8 mol/dm3 ). The Voltammetric studies show electrochemical instability for the
bath is stable; no changes are observed in the intensity of the cathodic peaks in negative potentials, which is connected with
peaks, leaving the solution clear and blue after 15 days. The baths hydrogen evolution. The hydrogen (I) reduction of protonated
are completely unstable when the pH is decreased (solution 11, citrate ions is inhibited by a tin layer, and by an underpotential-
Fig. 5d). The solution becomes colourless, rapidly generating red deposited zinc monolayer on the copper electrode, agreeing with
copper deposits. Both of these relationships are attributed to an the higher cathodic overpotential observed for the hydrogen evo-
Sn(II)-mediated Cu(II) reduction to metallic copper. Increasing the lution on the tin and zinc surfaces than on the copper surface (Tafel
pH beyond 5.75 destabilises the solutions through a slow, sys- “a” parameter is equal -1.24 V vs NHE for tin and zinc and -0.77 to -
tematic precipitation of a white deposit over several weeks. The 0.82 V vs NHE for Cu [31]). High concentrations of tin(II) and zinc(II)
X-ray diffraction analysis reveals that the major component of also decrease the concentration of the non-complexed, protonated
this deposit is tin(II) oxy-hydroxide (Sn6 O4 (OH)4 ) (JCPDS 46-1486 citrate ions in the solution.
[24]). Beyond pH 6.5 to 7.0, the precipitation of tin(II) hydroxide is The voltammograms for the separate tin(II) and zinc(II) reduc-
instantaneous. tions and their co-reduction are presented in Fig. 7a. The reduction
Therefore, the Cu(II) citrate complexes that have highly neg- of tin(II) starts at approximately -1 V vs SCE (Fig. 7a, curve 1) in a
ative charges (Cu2 HCit2 3− and Cu2 Cit2 4− ) contribute the most to solution with zinc(II). The reduction of hydrogen due to the decom-
the stability of the baths. The reduction of copper(II) is strongly position of water is favoured in solutions containing both zinc(II)
inhibited due to the high overpotential of the reduction from and tin(II), compared to solutions containing either zinc(II) or tin(II)
complexes with such high negative charges. Similarly, when the separately. Peak a1 , which is related to the oxidation of zinc, is
copper citrate complexes are reduced by the tin(II) complex, the smaller than peak a (Fig. 7a, curve 3). The anodic portion also con-
charge transfer process requires complexes with negative charges tains a slightly larger peak (b1 ), and is related to the oxidation of
to approach one another. Therefore, the activation energy for the tin. The reduction of zinc(II) is inhibited by the reduction of tin (II).
charge transfer process is greater when complexes with a high Fig. 7b presents the voltammograms obtained in citrate solu-
negative charge are present, inhibiting the reduction of Cu(II) tions containing copper(II) and tin(II) together. The beginning of
species. Significant amounts of Cu (II) species occur as sulphate the copper(II) reduction is shifted toward more negative potentials
complexes ([Cu(SO4 )n ]2−2n ) in solutions with low complexation (about -0.7 V vs SCE, Fig. 7b, curve 2); therefore, the limiting cur-
agent concentrations and pH values. The reduction of these com- rent of copper(II) reduction is also shifted toward more negative
plexes proceeds easily, particularly for neutral [CuSO4(aq) ]0 species. potentials. Peak b3 (oxidation of tin) is distinctly larger, indicating
Consuming these complexes via reduction causes their regenera- that more tin(II) has been reduced.
tion from the citrate complexes with high negative charges. The The zinc(II) reduction is faster in solutions containing copper(II)
Cu(II) reduction is slow, but not inhibited. When the sulphate and zinc(II) together, and peak a3 , which is related to the oxidation
complex concentration decreases due to increases in the concen- of zinc, is clearly larger (Fig. 7c, curve 2). The small peak (d) shown
tration of the complexation agent and pH of a solution, the highly on the anodic portion of the voltammogram is related to the oxida-
charged citrate complexes are only present in solution, inhibiting tion of the intermetallic CuZn phase (Cu5 Zn8 ); the presence of this
the reduction of copper (II) and stabilising the baths over long species in the deposits was confirmed by X-ray structural studies
periods. (Table 2). In addition, an increase in the reduction of hydrogen (I)
relative to the Zn(II)-citrate solution can be observed. Peak (c) is
3.2. Voltammetric studies related to the oxidation of porous copper residue from the CuZn
phase due to the considerable surface development.
Fig. 6 shows the cyclic voltammograms in citrate solution Fig. 8 shows the cyclic voltammograms for a solution containing
(0.65 M) for zinc(II), tin(II) and copper(II) separately. Curve (1) tin (II), zinc (II), and copper (II) together (curve 1). The reduction
depicts the voltammograms received from a solution containing process begins at approximately -1 V vs SCE. Peak a, which is related
only sodium citrate. The sharp increase in the cathodic current den- to the oxidation of zinc in the presence of tin(II), is lower; however,
sity for potentials below -1 V vs SCE is attributed to hydrogen(I) in the presence of tin(II) and copper(II) together, this peak nearly
M. Slupska, P. Ozga / Electrochimica Acta 141 (2014) 149–160 153

Fig. 5. UV-Vis spectra measured after specified intervals in solutions: a) 8 b) 9 c) 6 d) 11, T = 20 ◦ C.

disappears (peak a4 ). Likewise, the peak related to the oxidation of The voltammetric studies reveal that the electrodeposition
the CuZn phase decreases (d1 ). In addition, two anodic peaks (e1 and of an Sn-Zn-Cu alloy is possible when using citrate baths. The
e2 ) appear, related to the oxidation to Sn(II) and Sn(IV) respectively deposition processes for tin, zinc and copper begin in accor-
[4]. dance with the order of their equilibrium reduction potentials
(Cu(II)/Cu, Sn(II)/Sn, Zn(II)/Zn, Fig. 4); therefore, this type of co-
deposition is normal, according to the Brenner classifications
[32].

3.3. Influence of the hydrodynamic conditions on


electrodeposition

The effect of the RDE speed on the electrodeposition of tin from


stable citrate solutions containing Cu(II), Sn(II), Zn(II) species is
shown in Fig. 9a. These investigations utilised rotation rates of 15
and 68 rad/s for the cathode while using a number 9 electrolyte. The
partial current density for Sn electrodeposition remains indepen-
dent of the rotation rate at potentials from -1.0 to approximately
-1.3 V vs SCE. The limiting current density can be observed in this
range of the potential. The electrodeposition process was also car-
ried out at a constant -1.3 vs SCE, when the rotation rate of the
cathode ranged from 10 to 68 rad/s and electrolyte number 9 was
used to determine the limiting current (Fig. 10a). The partial current
Fig. 6. Cyclic voltammograms on a Cu substrate measured in solutions from 1 to 4. densities obtained at -1.3 V vs SCE are independent of the rotation
Scan rate 20 mV s−1 , ␻ = 16 rad s−1 , T = 20 ◦ C rate of the cathode. Therefore, the rate of the tin(II) reduction in
154 M. Slupska, P. Ozga / Electrochimica Acta 141 (2014) 149–160

Table 2
Phase structure and grain size of deposits.

No Bath E vs SCE/V j1 /mA cm−2 /rad s−1 Composition of deposits (wt. %) Phases2 Grain size3 /nm

Sn Zn Cu

1 No 13 -1.0 0.37 15 79.4 0.1 20.5 ␤-Sn, Cu6 Sn5 58.4


2 No 13 -1.3 1.29 15 89.3 0.3 10.4 ␤-Sn, Cu6 Sn5 55.8
3 No 13 -1.5 3.8 15 80.9 5.6 13.5 ␤-Sn, Cu6 Sn5 49.8
4 No 14 -1.9 23.8 68 77.2 13.5 9.3 ␤-Sn, Cu5 Zn8 49.0
5 No 8 -2.0 30.4 68 60.2 35.9 3.9 ␤-Sn, Zn, CuZn5 65.4
1
- average current density (potentiostatic deposition).
2
- ␤-Sn: tetragonal, space group I41/amd (141), PDF# 4-673; Cu5 Zn8 : regular, space group
I-43 m (217), PDF# 4-007-1117; Zn: hexagonal, space group P63/mmc, PDF# 4-0831;
Cu6 Sn5 : hexagonal, space group P63/mmc (194), PDF# 47-1575; CuZn5: hexagonal, space group P6/mmc, PDF# 35-1152
3
- for ␤-tin phase, on the basis of (101) peak broadening after correcting for instrumental broadening.

Fig. 8. Cyclic voltammograms on a Cu substrate measured in solutions from 5, 6 and


9. Scan rate 20 mV s−1 , ␻ = 16 rad s−1 , T = 20 ◦ C.

vs SCE when the rotation rate of the cathode ranged from 10 to


68 rad/s at pH = 5.5 (number 9) to determine the limiting partial
current density of tin (Fig. 10a). The tin partial current densities are
linearly related to the square root of the rotation rate at -1.9 V vs
SCE, confirming that the limiting current has a diffusional nature
and is connected to the direct reduction of the dominant tin (II)
citrate complex (SnCit2− ) according to reaction R5:

SnCit 2− + 2e− + 2H + → Sn0 + H2 Cit 2− (R5)

The diffusion coefficient of the Sn(II) species was calculated


based on Levich’s equation, generating a value of 6.3 10−7 cm2 s−1 .
This value is about one order of magnitude lower than for stannous
ions (6.5·10−7 cm2 s−1 [34]), but this diffusion coefficient is deter-
mined mainly through the dominant citrate complex (monomeric
SnCit−2 complex). The value determined for the diffusion coeffi-
cient is closer to values for dimeric citrate complexes (point E’ on
Fig. 7. a) Cyclic voltammograms on a Cu substrate measured in solutions 2, 3 and 5. the Fig. 11) than for monomeric citrate complexes (point E on the
Scan rate 20 mV s−1 , ␻ = 16 rad s−1 , T = 20 ◦ C. b) Cyclic voltammograms on a Cu sub- Fig. 11); therefore, the formula of this complex should be Sn2 Cit2 4− .
strate measured in solutions 3, 4 and 6. Scan rate 20 mV s−1 , ␻ = 16 rad s−1 , T = 20 ◦ C. This conclusion requires additional investigation. The Sn content
c) Cyclic voltammograms on a Cu substrate measured in solutions 2, 4 and 7. Scan
rate 20 mV s−1, ␻ = 16 rad s−1 , T = 20 ◦ C.
in the deposits increased slightly when the speed of the RDE was
increased (Fig. 12).
The relationship between the partial current density and the
this potential range is limited by the dissociation of the dominant
RDE rate for zinc deposition is presented in Fig. 9b. The partial cur-
tin(II) citrate complex (SnCit2− ) according to R3 and R4:
rent densities are independent of the rotation rate of the cathode,
SnCit 2− + H + ↔ SnHCit − (R3) even at -1.9 V vs SCE (Fig. 10b). Throughout the investigated range
of potentials and disc rotation rates, the zinc electrodeposition pro-
SnHCit − + 2e− + H + → Sn0 + H2 Cit 2− (R4) cess is controlled by the activation of the zinc(II) reduction from the
dominate citrate complex (ZnHCit− ), according to reaction R6:
The next limiting current density can be observed below -1.4 V
vs SCE. Electrodeposition was also carried out at a constant -1.9 V ZnHCit − + 2e− + H + → Zn0 + H2 Cit 2− (R6)
M. Slupska, P. Ozga / Electrochimica Acta 141 (2014) 149–160 155

Fig. 11. The dependence of diffusion coefficient values on the inverse radius of
Cu(II) complex species according to A. Norkus [33]. Values of Cu(II) complex species
determined by P. Ozga [27] were added in the figure (A- sulphate complexes B-
citrate monomers C- citrate dimers) and values of Cu (II) and Sn (II) complex species
determined in this work (D- value of Cu(II) complex species, E–value of Sn(II) com-
plex species for radius of citrate Sn(II) complex according to monomeric complexes,
E’ = –value of Sn(II) complex species for radius of citrate Sn(II) complex according to
dimeric complexes).

The amount of Zn in the deposits decreases for electrodeposition


at higher rotation rates and at more strongly negative potentials
(Fig. 12) when the electrodeposition of Sn and Cu is under diffusion
or mixed control.
Fig. 9c demonstrates the relationship between the partial
current density and the rate of the RDE during copper electrodepo-
sition. Two limiting current densities can be observed, similar to the
reduction of tin(II). The first occurs at potentials from -1 to -1.3 V vs
SCE. To determine the nature of the process, the electrodeposition
was carried out at a constant 1.3 V vs SCE, when the rotation rate
of the cathode ranged from 10 to 68 rad/s and a pH = 5.5 electrolyte
(number 9) was used. The partial current densities are independent
of the rotation rate of the cathode. The rate of copper reduction
is limited by the dissociation of the dominate copper(II) citrate
complex (Cu2 Cit2 4− ), according to R7 and R8:

Cu2 Cit24− + H + ↔ Cu2 HCit23− (R7)

Fig. 9. Steady state partial polarisation curves for: a) tin, deposition as a function
of pH of solution and RDE speed in solutions10, 15, 9, 13, ␻ = 16 rad s−1 , T = 20 ◦ C, b) Cu2 HCit23− + 4e− + 3H + → 2Cu0 + 2H2 Cit 2− (R8)
zinc deposition as a function of pH of solution and RDE speed in solutions10, 15, 9,
13, ␻ = 16 rad s−1 ,T = 20 ◦ C, c) copper deposition as a function of pH of solution and The relationship between the partial current densities for cop-
RDE speed in solutions10, 15, 9, 13, ␻ = 16 rad s−1 ,T = 20 ◦ C, per and the square root of the rotation rate was linear at -1.9 V
vs SCE (Fig. 10c); therefore, the limiting current is diffusional and

Fig. 10. Effect of RDE speed on electrodeposition of tin, zinc and copper. a) The partial current densities of tin deposition as a function of RDE speed, T = 20◦ C, b) The partial
current densities of zinc deposition as a function of RDE speed, T = 20◦ C, c) The partial current densities of copper deposition as a function of RDE speed, T = 20◦ C.
156 M. Slupska, P. Ozga / Electrochimica Acta 141 (2014) 149–160

Fig. 12. Content of elements in obtained deposit: a) content of elements in obtained deposit depending on pH of the solution, (␻ = 15 rad s−1 , 0.65 M Na3 HCit) b) content of
elements in obtained deposit depending on pH of the solution, (␻ = 68 rad s−1 , 0.65 M Na3 HCit) c) content of elements in obtained deposit depending on concentration of
sodium citrate, (␻ = 15 rad s−1 , pH = 5.5), d) content of elements in obtained deposit depending on concentration of sodium citrate, (␻ = 68 rad s−1 , pH = 5.5).

arises from the direct reduction of the dominant copper(II) citrate 3.4. Influence of the concentration of sodium citrate on
complex (Cu2 Cit2 4− ) (R9): electrodeposition

Fig. 13 depicts the steady-state partial polarisation curves for


Cu2 Cit22− + 4e− + 4H + → 2Cu0 + 2H2 Cit 2− (R9)
the co-deposition of Sn, Zn, Cu and the co-evolution of hydrogen
for electrolytes 8, 9 and 12 (Table 1). The deposits were obtained
The diffusion coefficient of the Cu2 Cit2 4− species is 8.9 at constant potentials ranging from -1 to -2.0 V vs SCE while the
·10−7 cm2 s−1 , similar to other dimeric citrate complexes (point D rotation rate was 15 rad/s. The sodium citrate concentration ranged
on the Fig. 11). When increasing the rate of the RDE, the Cu content from 0.4 M to 0.8 mol/dm3 . In this range of concentrations, dramatic
in the deposits increases (Fig. 12). changes in speciation are observed, particularly for zinc(II) (Fig. 3).
The current efficiency of the electrodeposition process changes The concentration of Na3 HCit is also strongly dependent on the cur-
slightly when increasing the rotation rate of the cathode (Fig. 12) rent density of the electrodeposition process. The partial current
because the reduction of hydrogen(I) (according to reaction R2) is density decreases when increasing the sodium citrate concentra-
under activation control at all of the investigated potentials. R1 is tion, and the reduction of Zn(II) decreases. The most significant
inhibited by the presence of tin and zinc on the surface of the disc decrease occurs between 0.4 M and 0.65 M. When increasing the
electrode (Fig. 6). concentration of sodium citrate, the dominant citrate complex
M. Slupska, P. Ozga / Electrochimica Acta 141 (2014) 149–160 157

Fig. 13. Steady state partial polarisation curves for tin, zinc, copper and hydrogen as a function of concentration of sodium citrate. a) Steady state partial polarisation curves
for tin as a function of sodium citrate concentration in solutions 8, 9 and 12, ␻ = 16 rad s−1 , T = 20 ◦ C b) Steady state partial polarisation curves for zinc as a function of sodium
citrate concentration in solutions 8, 9 and 12 ␻ = 16 rad s−1 , T = 20 ◦ C c) Steady state partial polarisation curves for copper as a function of sodium citrate concentration in
solutions 8, 9 and 12 ␻ = 16 rad s−1 , T = 20 ◦ C d) Steady state partial polarisation curves for hydrogen as a function of sodium citrate concentration in solutions 8, 9 and 12,
␻=16 rad s−1 , T = 20 ◦ C.

changes (Fig. 3). Less ZnHCit− complex is present in solution, allow- The partial current densities of tin increase when increasing the
ing the ZnH2 Cit2 4− complex to dominate. The ZnH2 Cit2 4− complex sodium citrate concentration at potentials below -1.4 V vs SCE;
has a highly negative charge; because this complex is reduced at however, further increases in the concentration (0.65 M to 0.8 M)
higher overpotentials than the ZnHCit− and ZnH2 Cit2 4− complexes, do not affect the electrodeposition of tin. The Sn content decreases
it is almost electrochemically inactive under these experimental when decreasing the potential (from 92 to 62 wt.%). The highest
conditions, which are limited by significant hydrogen evolution. tin content in the deposit is reached at higher potentials (from -1
Therefore, increasing the concentration of this complex decreases to -1.4 V vs SCE). The lowest tin content is achieved at low con-
the partial current density of zinc, which is determined only by centrations of sodium citrate, at lower potentials (below -1.7 V vs
the reduction of the ZnHCit− complex. Consequently, the zinc con- SCE).
tent in the deposits increases when the potential decreases, and The partial current densities during the electrodeposition of
increasing the sodium citrate concentration in a solution decreases copper are low relative to the current densities for tin and zinc
the zinc content of the deposit (Fig. 12 c and Fig. 12 d). The highest (Fig. 13c). Changes in the sodium citrate concentration do not affect
zinc content occurs at low Na3 HCit concentrations (approximately the curves significantly. This effect is most apparent between -1.3
32 wt.%). and -1.9 V vs SCE (between 0.4 M and 0.65 M Na3 HCit). The current
The electrodeposition potential strongly influences the partial density increases when increasing the Na3 HCit concentration. Sim-
current density of tin (Fig. 13a) for all investigated concentrations ilar to tin, the dependence for copper between 0.65 M and 0.8 M
of sodium citrate. The limiting current densities are slightly visible nearly disappears. The copper content in the received deposit is
on the polarisation curves at low citrate concentrations (0.4 M); shown in Fig. 12. At low potentials, the Cu content remains indepen-
they are more visible when the solution has a high sodium citrate dent of the sodium citrate concentration. The relationship between
concentration (0.65 M and 0.8 M). The significant influence exerted the concentrations of the complexation agent and the Cu is appar-
by the complexation agent is apparent between 0.4 M and 0.65 M. ent at higher potentials (from -1 to -1.6 V vs SCE). The Cu content
158 M. Slupska, P. Ozga / Electrochimica Acta 141 (2014) 149–160

Fig. 14. Steady state partial polarisation curves for tin, zinc, copper and hydrogen as a function pH of solution. a) Steady state partial polarisation curves for tin as a function
of pH of solution in solutions 10, 15, 9, and 13, ␻ = 16 rad s−1 , T = 20 ◦ C b) Steady state partial polarisation curves for zinc as a function of pH of solution in solutions 10, 15,
9, and 13 ␻ = 16 rad s−1 , T = 20 ◦ C c) Steady state partial polarisation curves for copper as a function of pH of solution in solutions 10, 15, 9, and 13, ␻ = 16 rad s−1 , T = 20 ◦ C d)
Steady state partial polarisation curves for hydrogen as a function of pH of solution in solutions 10, 15, 9, and 13, ␻ = 16 rad s−1 , T = 20 ◦ C.

increases when the concentration of the complexation agent is to the surface of the electrode due to the reduction of the tin(II)
increased. The Cu content is also related to the increasing current complexes and the release of H2 Cit2− ions in reactions R4 and R5.
density in this range. The maximum Cu content is 8.5 wt.%. Consequently, inhibiting the Zn(II) reduction is related to decreases
Fig. 13d shows the steady state partial polarisation curves in the concentration of the ZnHCit− complex, due to the formation
for hydrogen. In general, the current density increases when of electrochemically non-active complex ZnH2 Cit2 4− species in the
decreasing the electrodeposition potential, according to reaction R1 layers of the solution close to the surface of the electrode, according
and R2. When decreasing the concentration of sodium citrate, the to chemical reaction R10:
partial polarisation curves for hydrogen decrease, while the current
efficiency increases. These changes occur due to the lower concen- ZnHcit − + H2 Cit 2− → ZnH2 Cit24− + H + (R10)
tration of protonated citrate ion (H2 Cit2− ) in baths with lower ratios
of Cit/Me(II) (the hydrogen(I) from the free protonated species of
citrate ion is reduced according to reaction R1). Maximum effi- 3.5. Influence of solution pH on electrodeposition
ciency occurs at high potentials; the efficiency can reach 100%. A
small quantity of copper can also be deposited through an electro- Fig. 14 shows the steady-state partial polarisation curves for
less process. This parallel deposition process is very slow and is only Sn(II), Zn(II), Cu (II) and hydrogen during their codeposition with
significant during electrodeposition at very small cathodic current electrolytes 9, 10, 13, and 15 (Table 1). These curves reveal the rela-
on steel electrodes. tionship between the solution pH and the partial current density.
Reactions R3 to R7 do not explain the observed connection The solution pH varies from 5 to 5.75. The deposits were obtained
between the reduction of tin(II) and zinc(II) (inhibition of zinc(II) from -1 to -2.0 V vs SCE at 15 rad/s.
reduction by the tin(II) reduction, Fig. 7a). This effect can account for Fig. 14a presents the steady-state partial polarisation curves
the increased citrate ion concentration in the solution layers close for tin. The first limiting current density observed for the higher
M. Slupska, P. Ozga / Electrochimica Acta 141 (2014) 149–160 159

potential decreases when increasing the pH. This limiting current 4. Conclusions
is connected to reduction of the SnHCit− citrate complex described
in reaction R4; therefore, when the concentration of this com- • Thermodynamic analysis of the model of citrate baths reveals
plex decreases with increases in the pH, the limiting current is that the optimal conditions for the electrodeposition of Sn-Zn-
decreased. The second limiting current density is observed at more Cu alloys from citrate solutions require a pH between 4 and 6,
negative potentials, but is not fully attained; therefore, there is a but the citrate baths generated under these conditions may be
minor relationship between the pH of solution and the first wave unstable due to the reduction of Cu(II) by Sn(II) species.
of reduction (presence of a reduced SnHCit− citrate complex). The • The development of a stable electrolyte citrate bath for the elec-
Sn content in the deposits, which is related to the pH and depo- trodeposition of SnZnCu alloys is possible at pH 5 to 5.75 when the
sition potential, remains independent of the solution pH (Fig. 12a citrate ion concentration exceeds 0.4 M. The reduction of copper
and Fig. 12b) and only changes when decreasing the potential. The (II) species by negatively charged tin(II) complexes is inhibited
maximum Sn content at low pH values is observed from -1 V to because the activation energy of the charge transfer process is
-1.4 V vs SCE (94-96 wt.%). high. High overpotentials are also observed for the reduction of
The steady-state partial polarisation curves for Zn (II) are shown citrate copper(II) complexes during voltammetric investigations.
in Fig. 14b. The curves remain independent of the solution pH. • The voltammetric studies reveal the possibility of electrodeposit-
The Zn reduction begins below -1.2 V. Between -1.2 V and -2 V vs ing Sn-Zn-Cu alloys from citrate baths. The deposition processes
SCE, the Zn content remains independent of the solution pH and for tin(II), zinc(II) and copper(II) agree with their standard poten-
increases when increasing the solution pH over a narrow range at tials, which is normal co-deposition according to the Brenner
potentials below -2 V vs SCE. classification. The zinc(II) reduction was strongly inhibited during
The steady-state partial polarisation curves for copper are its co-reduction with tin(II).
similar to those of tin (Fig. 14c). The solution pH is weakly • A mechanism for the co-reduction of the tin(II), zinc(II) and cop-
related to the limiting current density. When increasing the solu- per(II) citrate complexes, as well as the protonated citrate species,
tion pH, the current density decreases. This result is related to was proposed to explain the influence of the potential, hydro-
the decreased Cu2 HCit23− concentration at increased pH; there- dynamic conditions, pH and concentration of the complexation
fore, the limiting current related to R8 decreases. The Cu agent on the composition of the Sn-Zn-Cu deposit and current
content in the deposit remains independent of the solution efficiency.
pH (Fig. 12) and decreases when the electrodeposition poten- • Unlike the electrodeposition of tin and copper, the electrodeposi-
tial decreases. The Cu content peaks at the highest potentials tion of zinc is under activation control due to the reduction of an
(8 wt.%). electroactive citrate zinc(II) complex (ZnHCit− ) with the inves-
Fig. 14d presents the steady-state partial polarisation curves for tigated range of electrode potentials; therefore, changes in the
hydrogen, which agree with the results obtained previously. Below hydrodynamic conditions (increase of RDE rate) decrease the zinc
-1.6 V vs SCE, the reduction of hydrogen increases significantly, content.
according to reactions R1 and R2. The solution pH appears inde- • Increasing the citrate concentration in the bath induces the
pendent of the current density. The current efficiency decreases release of citrate ions during the co-reduction of tin(II) citrate
when decreasing the electrodeposition potential. This process is complex conduct to (i) decrease the concentration of the
the most efficient at high potentials, while remaining independent electroactive zinc(II) citrate complex (ZnHCit− ) near the elec-
of the solution pH. trode surface and to (ii) increase the concentration of the
non-electroactive zinc(II) citrate complexes with high negative
charges (Zn2 Cit2 4− and ZnH2 Cit2 4− , which have a high overpo-
3.6. Phase structure and grain size tential for Zn(II) reduction).
• The changes in pH between 5.0 and 5.75 have little influence on
The initial investigations were performed for phase structure the concentration of the electroactive zinc(II) citrate complex.
and grain size determination by X-ray diffraction (Table 2). The ␤- However, these changes in pH decrease the concentration of com-
tin phase (tetragonal, space group I41/amd (141), PDF# 4-673) is plexes with lower overpotentials. Consequently, the first plateau
the main phase in all deposits. The second phase Cu6 Sn5 (hexag- in the limiting current for Cu(II) and Sn(II) decreases, and the zinc
onal, space group P63/mmc (194), PDF# 47-1575) is present in content increases slightly when the pH is increased.
deposits with low zinc content. The Cu5 Zn8 phase (regular, space • The possibility of electrodepositing Sn-Zn-Cu alloys from citrate
group I-43 m (217), PDF# 4-007-1117) replaces Cu6 Sn5 in deposits solutions was confirmed via potentiostatic deposition. The tin
containing more zinc (13.5 wt.% Zn, 9.3 wt.% Cu). The deposit with content of the coatings varied from 60 to 96 wt.%, the copper con-
the highest content of Zn (35.9 wt.% Zn, 3.9 wt.% Cu) contains CuZn5 tent varied from 3.5 to 20.5 wt.%, and the zinc content approached
phase (hexagonal, space group P6/mmc, PDF# 35-1152) as the 37.5 wt.%.
second phase, and zinc (hexagonal, space group P63/mmc, PDF#
4-0831) as the third phase. The determined phases are in accor-
Acknowledgements
dance with the phase diagram for the Cu-Sn-Zn system [10]. All
investigated deposits are nanocrystalline with a grain size of about
This work was supported by IMMS PAS (Z1: Environment
50 to 65 nm. The grain size of deposits obtained from the same
friendly technologies and materials. Lead-free soldering).
bath decreases with an increase of cathodic overpotential and
current density (samples No 1 to 3, Table 2) according to a funda-
mental relationship, that an increase of the cathode overpotential References
leads to an increase of the probability of crystalline nucleus for-
mation. However, the comparison of the deposits obtained from [1] P. Ozga, Electrodeposition of Sn-Ag and Sn-Ag-Cu alloys from the thiourea
solutions, Archives of Metallurgy and Materials 3 (2006) 413.
the different baths (samples No 4 and 5, Table 2) also indicates [2] J.E. Lee, K.S. Kimb, M. Inoue, J. Jiang, K. Suganumab, Effects of Ag and Cu addition
the important role of the parameters of bath (citrate concentration, on microstructural properties and oxidation resistance of Sn–Zn eutectic alloy,
pH) or co-reduction of tin(II) with zinc(II) species on the grain size. Journal of Alloys and Compounds 454 (2008) 310.
[3] M. Grobelny, N. Sobczak, Effect of pH of sulfate solution on electrochemical
Determination of the nature of this phenomenon requires further behavior of Pb-Free solder candidates of SnZn and SnZnCu systems, Journal of
investigation. Materials Engineering and Performance (2012) 614.
160 M. Slupska, P. Ozga / Electrochimica Acta 141 (2014) 149–160

[4] H. Kazimierczak, P. Ozga, Electrodeposition of Sn–Zn and Sn–Zn–Mo layers [20] H. Katagiri, Cu2 ZnSnS4 thin film solar cells, Thin Solid Films 480–481 (2005)
from citrate solution, Surface Science 607 (2013) 33. 426.
[5] S. Dubent, M. De Petris-Wery, M. Saurat, H.F. Ayedi, Composition control of [21] R. Juskenas, S. Kanapeckaite, V. Karpaviciene, Z. Mockus, V. Pakstas, A. Selskiene,
tin-zinc electrodposit through means of experimental strategies, Materials R. Giraitis, G. Niaura, A two-step approach for electrochemical deposition of Cu-
Chemistry and Physics 104 (2007) 146. Zn-Sn and Se precursors for CZTSe solar cells, Solar Energy Materials & Solar
[6] Chi-Chang Hu, Chun Kou Wang, Gen Lan Lee, Composition control of tin zinc Cells 101 (2012) 277.
deposits using experimental strategies, Electrochim. Acta 51 (2006) 3692. [22] Xe Xin-kuai, Chen Bai-zhen, Hu Geng-sheng, Deng Ling-feng, Zhou Ning-bo,
[7] K. Wang, H.W. Pickering, K.G. Wei, EQCM studies of the eletrodeposition and Tian Wen-zeng, Process of electroless plating Cu-Sn-Zn ternary alloy, Trans.
corrosion of tin-zinc coatings, Electrochim. Acta 46 (2001) 3835. Nonferrous Met. Soc. China 16 (2006) 223.
[8] E. Guaus, J. Torrent–Burgues, Tin-zinc electrodposition from sulphate-tartrate [23] S.D. Beattiea, J.R. Dahn, Combinatorial electrodeposition of Ternary Cu–Sn–Zn
bath, Journal of Electroanalytical Chemistry 575 (2005) 301. alloys, Journal of Electrochemical Society 152 (2005) C542.
[9] E. Guaus, J. Torrent–Burgues, Tin-zinc electrodposition from sulphate- [24] Joint Committee on Powder Diffraction Standards, JCPDS in: International Cen-
gluconate baths, Journal of Electroanalytical Chemistry 549 (2003) 25. tre for Diffraction Data. Powder Diffraction File-PDF-4, ICDD, Pennsylvania
[10] Chin-yi Chou, Sinn-Wen Chen, Phase equilibria of the Sn-Zn-Cu ternary system, 2007 (CDROM).
Acta Materialia 54 (2006) 2393. [25] L.G. Sillen, A.E. Martel, Stability Constants of Metal–Ion Complexes, The Chem-
[11] Yu-chih Huanga, Sinn-wen Chena, Chin-yi Choub, W. Gierlotka, Liquidus pro- ical Society, London, 1964, Supplement No. 1 (1971).
jection and thermodynamic modeling of Sn–Zn–Cu ternary system, Journal of [26] L.D. Pettit, K.J. Powell, The IUPAC Stability Constants Database, SC-Database
Alloys and Compounds 477 (2009) 283. for Windows, Academic Software, Release 5, Sourby Old Farm, Timble, Otley,
[12] Byeong–Joo Lee, Nong Moon Hwang, Hyuck Mo Lee, Prediction of interface Yorks, UK (2012).
reaction products between Cu and various solder alloys by thermodynamic [27] P. Ozga, The Influence of Complexation on the Electrodeposition of Metals and
calculation, Acta Materialia Vol.45 No. 5 (1997) 1867. Alloys from Citrate Solutions, (in Polish), IMIM PAS, Kraków, ISBN-83-921845-
[13] S.M. Pawar, B.S. Pawar, A.V. Moholkar, D.S. Choi, J.H. Yun, J.H. Moon, Single 8-0 (2006).
step electrosynthesis of Cu2 ZnSnS4 (CZTS) thin films for solar cell application, [28] H. Kazimierczak, P. Ozga, R. Socha, Investigation of electrochemical co-
Electrochim. Acta 55 (2010) 4057. deposition of zinc and molybdenum from citrate solutions, Electrochim. Acta
[14] T. Maeda, S. Nakamura, T. Wada, First-principles calculations of vacancy forma- 104 (2013) 378–390.
tion in In-free photovoltaic semiconductor Cu2 ZnSnSe4 , Thin Solid Films 519 [29] E. Beltowska-Lehman, P. Ozga, Effect of complex formation on the diffusionco-
(2011) 7513. efficient of Cu(II) in citrate solution containing Ni(II) and Mo(VI), Electrochim.
[15] T. Tanaka, A. Yoshida, D. Saiki, K. Saito, Q. Guo, M. Nishio, T. Yamaguchi, Influ- Acta 43 (1998) 617.
ence of composition ratio on properties of Cu2 ZnSnS4 thin films fabricated by ˛
[30] H. Kazimierczak, P. Ozga, M. Słupska, Z. Światek, K. Berent, Electrodeposition of
co-evaporation, Thin Solid Films 518 (2010) S29. Sn-Mn Layers from Aqueous Citrate Electrolytes, Journal of The Electrochemical
[16] B.S. Pawar, S.M. Pawar, S.W. Shin, D.S. Choi, C.J. Park, S.S. Kolekar, J.H. Kim, Society 161 (6) (2014) D309–D320.
Effect of complexing agent on the properties of electrochemically deposited [31] L.I. Kristhalik, in: P. Delahay, C. Tobias (Eds.), Advances in Electrochem-
Cu2 ZnSnS4 (CZTS) thin film, Applied Surface Science 257 (2010) 1786. istry and Electrochemical Engineering, vol. 7, Interscience, New York, 1970,
[17] K. Tanaka, Y. Fukui, N. Moritake, H. Uchiki, Chemical composition dependence p. 283.
of morphological and optical properties of Cu2 ZnSnS4 thin films deposited by [32] A. Brenner, Electrodeposition of Alloys, Academic Press, New York, 1963, vols.
sol-gel sulfurization and Cu2 ZnSnS4 thin film solar efficiency, Solar Energy 1 and 2.
Materials & Solar Cells 95 (2011) 838. [33] E. Norkus, Diffusion coeffcients of Cu(II) complexes with ligands used in
[18] G. Suresh Babu, Y.B. Kishore Kumar, P. Uday Bhaskar, Sundara Raja Vanjari, alkaline electroless copper plating solutions, J. App.Electrochem. 30 (2000)
Effect of Cu/(Zn+Sn) ratio on the properties of co-evaporated Cu2 ZnSnSe4 , Solar 1163.
Energy Materials & Solar Cells 94 (2010) 221. [34] C.T.J. Low, F.C. Walsh, The stability of an acidic tin methanesulfonate electrolyte
[19] H. Araki, A. Mikaduki, Y. Kubo, T. Sato, K. Jimbo, W.S. Maw, H. Katagiri, M. in the presence of a hydroquinone antioxidant, Electrochim. Acta 33 (2008)
Yamazaki, K. Oishi, A. Takeuchi, Preparation of Cu2 ZnSnS4 thin films by sulfu- 5280.
rization of stacked metallic layers, Thin Solid Films 517 (2008) 1457.

Das könnte Ihnen auch gefallen