You are on page 1of 66

0

0 0

e e e
8

a
8 Q 8
0

PROCESSES AND APPLICATIONS FOR TIN AND TIN-BASED ALLOY SURFACE COATING TECHNOLOGIES

A Technical Review and Assessment of Recent Developments Compiled for Tin Technology

e
8

STAGE 1 ELECTROLYTIC DEPOSITION

PART 1 ELECTRODEPOSITION AND ELECTROPLATING OF TIN

e
@

e a
0

L. M. Baugh, BSc, MSc, PhD, Chem, FRSC Consultant

4 D

e
0

a
0

January 2005

0
>

CONTENTS

0
0

1.

Introduction.. ............ ... .. .....

....................................................................4

0 0 0
0 0

2.

Fundamental Aspects. of the Electrodeposition of Tin.. ... ................

............................5

2.1 2.2 2.3 2.4


3.

Determination of the Mechanism of Nucleation and Growth of Tin.. . .. . .. . . ... . . . .. . . . . .. .. ...5 Electrodeposits Microstructure of Tin Electrodeposits and the Role of the Substrate.. . . . . .. . . .. . ..... ... . . . . . ..9 Preparation of Dendritic Tin Nanoaggregates.. ... .. .. . . . . . . . . ... .... ... . . . . . . . . . ...... .... . ... . ...11 Underpotential Deposition of Tin.. .. . . . . . ... . ... . .. . . ..... . . ... . .. . .. . . . . .. .. .. . . . . . ... ... . . .... . .,.13

0
0 0 0
0

Role of Additives in the Electrodeposition of Tin.. ... ......... ..

. ......... .. ..,................14 .

3.1 3.2 3.3 3.4 3.5 3.6


4.

Adsorption Characteristics... . . . . . . . . . . . . . . . ... . . . . . . . . . . . . . .. . ... . . . . . .. .. . . ... . . . . . . . . .. . . . .... . ....14 Sodium Gluconate.. . .. . . . . . . . . . . . . ... . . . . . .. .. .. . . . . . .... . . .. . . . . .. . . . . . . . . . . . . . . . . . .. .. . . . . .. . . . . . . ..16 Cetyltrimethyl-ammonium Chloride and Cetylpyridinium Bromide.. .. . ... . . .. . . . . .. . ... .. . . ..20 S-dodecylmercaptobenzimidazole.. . . .. . . . . . . . . . .. . . . ... . . . .. . . .. . .. . . .. .. . . . . . . . ... ..... .. .. . . .... 2 1 . Miscellaneous Additives.. . .. . .. . . ... . . . .. . .. .. . . . . . ..... . .. . . . . .. . ... . . ... . . . . . . .. .... . . . . .. . . . ......23 . Synergistic Effects of Additives.. .. . . ... ... . . .. . . . . . .... . . . . . ... . ...... .. . .. . . . .. ... . : . . . . .. . . .. ... 23
24

0 0
0

Development of Plating Baths Based on Sulphonic Acids ............................................

4.1 4.2 4.3 4.4 4.5 4.6

Comparison of Halogen and Methanesulphonic Acid Electrotinning Processes .... .. . .. .. . ... . ..... . .. .. . . ... ... . . . ... .. ..... . . .. ... . . . .. .. . . . ... . . . .:. . .... . ... . .. . .. . .. . . . . . .25 Effect of Process Variables on Electrotinning in Methanesulphonic Acid.. . . . . . . . . . . . . . . .. . . . . ... . . . .. . . .. . . . . . . .. . . . .... . . .. . . . . . . . . . . .. . . . . .. .. . . . ... . . . . .. .. .. .. .. .. . . . ....3 1 Methanesulphonic Acid Electrolyte in the Presence of Polyethylene Glycols. . .. . .. . . . .... . . . . . . ... . . .... . . . . .. .. .. .. . . . . . .. . .. . . . . . .... .. . . . . . . .. . .. . . . .. . . . . . . ... . . . .. . . . ... 38 Napthosulphonic and Phenolsulphonic Acid Electrolytes.. .. . .. . . .. . .. . .. .. . . .. . .. . .. . .. . . .... 4 1 .Elimination of Tin Whiskers by Reducing Stress Levels in Electrodeposits. . . . . . ... . . . . . . .. . ... . . . . . . . . . ... . .. . . . .. . . . . .... . . .. . ... . . . .. . . . . . . . . .. . . .. . . . ..... .. ...46 Production of Tin Methanesulphonate. . .. . .. . . . . . .. .. .... .. . .. . . . . . .. . . . . . . . ... . . . . . . . . . . . . . . . . . . ...46

0
0 0

e
0 0 0 0 0 0

5.

Electrolytes and Additives in High Speed Plating..

....................................................47

5.1 5.2 5.3 5.4


5.5 5.6

Sulphonic Acid + Alkylene Oxide Condensation Product.. ... .. . . . .. . . .. . . . . . .. . ... ..... . ... . . .48 Stannous Alkyl Sulphonate + Alkyl Sulphonic Acid + Reaction Product of Polyalkylene Glycol and Phenolphthalein or Derivatives of Phenolphthalein.. . . . .. .. .. . 48 Sulphonic Acid + Substituted Phenol. .. .. .. . . . . . . . . . . . . .. . .... . . . .. . . . .. . . . . . . . . . . . . . . . .. . . . . .. ... 49 Alkali Metal, Alkaline Earth Metal, Ammonium or Substituted Ammonium Salts of Alkyl or Alkanol Sulphonic Acid.. . . .. ... . . . . .. . .. . .... . .. . .. . . . . .. . .. .. . . . ... . . . . . . . . . . .49 Sulphonic Acid + Carboxylated Polyalkeneimine Compounds.. . . . . . . . .. . . . . . ...... . . .. .. . . . ... 50 Toluene Sulphonic Acid + Ammonium Salts and/or Magnesium Salts.. . . . . . . . . . . . . .. . .. . ... 50

0
0
0 0

0
2

0
0

0
0
0
6.

Electroplating and Electrotinning Quality.. ...............'. . . 6.1 6.2

........................................... 50

0 0 0
0
7=
8.

Testing of Tin Plating Using a Rotating Cone Electrode.. . ...... ... . . . .. . .. .. . . .. .. .. . .. .. . . .... 50 . Surface Morphology and Appearance of Electrodeposited Tin Films.. . . . . . . .......... . . .... .. . ...... . . . .... . . ... ... . . . ..... . . . . ... ... ..... .. .. .. . ....... ... . . . . ... ..53

Conclusions and Recommendations ............... .............. .... ............ ...................... 58 References.. ..

. .

0 0
0

.................................................................................................. 62 .
............................................66

9.

.Acknowledgements................................................

0
0 0

e
0 0

0 0
0

0
0

0
0

0 0
0

0 0 0
0
3

1. INTRODUCTION

One of the largest applications for tin is in the production of tin plate. By 1975 over 98% of total world production of tinplate was produced by an electrolytic deposition process. Although this still remains the largest application for electrolytically deposited tin and therefore continues to be very important, the Coatings Technology Division within Tin Technology is attempting to develop a strategy and identify various research opportunities in related areas. However, the first step in this process is an assessment of existing technologies and product applications. This and subsequent reviews are concerned with detailed analyses of the current situation.
1

It was initially envisaged that the subject as a whole could be dealt with in three separate stages covering:r

Stage 1 Stage 2 Stage 3

Electrolytic deposition of tin and tin alloys Electroless deposition of tin and tin alloys Surface engineering processes involving tin and tin alloys

However, preliminary searches revealed that in view of the vast number of publications and patents in the field of the electrodeposition and electroplating of tin and tin alloys, separate reviews would be necessary to do justice to this work. It was therefore proposed that the initial stage 1 review would be sub-structured as follows:Stage 1 Electrolytic Deposition Electrodeposition and Electroplating of Tin Electrodeposition and Electroplating of Tin Alloys Solders and Applications in Electronics Electrodeposition and Electroplating of Tin Alloys Corrosion Resistant Coatings and Applications in Electrochemical Power Sources

Part 1 Part 2

Part 3

The current review (Part 1) covers the very important topic of electrodeposition and electroplating of tin. The visual appearance and properties of electrodeposited tin coatings are determined by their surface morphology, grain structure, grain size, grain orientation and general surface roughness, etc. Coatings have often been classified as burnt, matte, satin-bright and bright, among other descriptions. Clearly, it is essential to specify the type of coating required for particular applications. In order to achieve any degree of control over the type of deposit produced, it is necessary to understand as much as possible about the mechanism of electrodeposition and the effect of the experimental variables. These include the electrochemical method applied, the hydrodynamic conditions and, most importantly, the composition of the plating bath. This review sets out to assess recent progress in the pursuit of these aims. The topics covered include: a discussion of fundamental aspects of the electrodeposition of tin;
4

0
4D

a
4 D

e
8

the role o additives; the development of plating baths based on sulphonic acids; f electrolytes and additives in high speed plating; and electroplating/electrotinning quality. This and the subsequent reviews will help to define the current state-of-the-art and identify areas of research that are still lacking, thereby facilitating the production of future research proposals.

e
4 D

2.
2.1

Fundamental Aspects of the Electrodeposition of Tin Determination of the Mechanism of Nucleation and Growth of Tin Electrodeposits

e
6
0

Little infomation is available relating to the early stages of tin electrodeposition. Recently, Gomez et al. [ 11 have examined this process on carbon substrates, without additives, from acid solutions containing Sn (11) ions. Although these conditions are not optimal for the deposits required for specific applications, in this study the authors attempted to understand the mechanisms of formation and evolution of the initial nuclei having previously studied the relation between electrodeposition conditions and the morphology of the deposits [ 11. The electrodes used in the study were composed of either vitreous carbon or highly oriented pyrolytic graphite (HOPG) and the tin was deposited fiom 0.01 mol dm-3 HZS04 solutions. The vitreous carbon electrode was polished to a SnS04 + 1 mol mirror finish before each experiment using alumina of different grades and then cleaned ultrasonically in water. HOPG electrodes were prepared by cleavage and exposure of the basal plane. Electrochemical techniques included cyclic voltammetry and potential step methods. In order to analyse the mechanisms of nucleation and growth during electrodeposition, scanning electron microscopy (SEM) has been used extensively as a Complementary technique to conventional electrochemical methods applied to study the mechanisms and kinetics of metal electrocrystallisation. However, scanning tunneling (STM) and atomic force microscopies (AFM) complement the usual techniques of surface observation and provide information about the formation of the initial nuclei and the topography of the crystallites. These complementary techniques were also applied. Fig. 1(a) shows scanning electron micrographs of the tin electrodeposits obtained on the vitreous carbon electrode held at -530 mV (vs. Ag/AgCl) for 100s and at -700 mV for 60s. In the former case, only a few crystallites of definedparallelepipedal shape with an average size of 9 pm were observed. Higher magnification did not reveal new tin particles. When the deposits were obtained at more negative potentials (-700mV) the number of microcrystals increased considerably (Fig. 1(b). In contrast, no well defined crystals were observed under these conditions; the crystallites showing multiple shapes and overlap. The greater growth rate here may have been responsible for the differences in crystallite shape.

e e
8

a e
e a
8

e e
0 0

e
8

e
Q
5

lD

0
3

3
3

0
0
0

Fig. 1. Scanning electron micrographs of tin electrodeposits obtained on vitreous carbon electrodes (a) -530 mV and 100s of deposition time and (b) -700 mV and 60s of deposition time [ 11.

0 0 0 0 0 0 0 0 0 0 0

On the basis of the initial results on vitreous carbon, the AFM technique was chosen to study the earliest formation of crystallites on an ordered substrate (HOPG). The HOPG basal plane electrode is more inert than the vitreous carbon substrate and this implies that the HOPG substrate is appropriate for the analysis of the very initial formation of the tin electrodeposit. The optimized potential range for the slow deposition of tin was established between -460 mV and -530 mV. Under these conditions the current-time transients showed a monotonic increase (Fig. 2). The evolution of the tin electrodeposition was studied at a fixed potential and different deposition times. As can be seen in Fig. 3, the deposition was favoured at step edges, whereas the terraces showed a more homogeneous distribution of particles.

0 0 0

0
0
0 0

0 0 0 0
I

a
0
0 0
.

20

40

60

80

100

120

t/s

Fig. 2. Current-time transient for tin deposition on the highly oriented pyrolytic graphite electrode following a potential step from -1 00 mV to -500 mV [ 13.

5.00

a
0
0
2.50
5 0 . 0 nn

a
0
25.0 nn

0
0
0

2.50

,O 5.00
vu

0 . 0 nu

0
0 0

Fig. 3. Atomic force microscopy image of tin deposit obtained during 130s at -500 mV. HOPG electrode [ 11.

a e
0
0
0

The initial stages of deposition were represented by means of two ex-situ images of deposits obtained on flat terraces at different times at -500 mV (Fig. 4). Fig. 4a, taken at 20s, shows a random distribution of different sized aggregates with a density of around 73 aggregates mrf2. A longer deposition time, 353, produced around 140 aggregates mm-*(Fig. 4b). Rather unstructured three dimensional deposits were obtained. The growth of the aggregates was favoured in the x-y directions rather than in the z-direction. A detailed inspection of images revealed twinning between aggregates.

a
0
7

X
0

2
W

i
I
I

0 0 0 0 0 0 0 0 0 0 0

Fig. 4. Atomic force microscopy images of tin deposit obtained at -500mV and (a) 20s and (b) 35s (points a and b in Fig. 2 [l].

0 0 0 0 0 0 0 0 0 0 0

The following general conclusions were drawn by the authors from this work:Analysis of the electrochemical results indicated that the tin electrodeposition process occurs by an instantaneous nucleation and three dimensional growth process limited by diffusion. Tin electrodeposition becomes easier than other electrodeposition processes, since low overpotentials and short times are enough to form abundant aggregates with an average diameter of a few nanometers. The deposit does not grow by a two dimensional layer by layer mechanism. On the contrary, deposition occurs preferentially at step edges, dislocations and defects in the surface. The AFM images of HOPG indicate that the crystallites have as predecessors three-dimensional poorlystructured overlapped aggregates formed as a result of surface diffusion of the aggregates formed previously. Incipient parallepipedical crystallites grow preferentially in the z-direction. The growth is controlled by the diffusion of Sn(I1) species in solution and facilitated by the easier deposition of tin on tin. The well defined tetragonal shapes are achieved by the low energetic barrier for the reorganisation of tin atoms on the crystalline surface.

0 0

0 0 0
0
2.2 Microstructure of Tin Electrodeposits and the Role of the Substrate

a
0

8
9
0 0

Teshigawara et al. [2] have investigated the relationship between the electrodeposition conditions and the surface morphology, grain size and orientation of tin deposits. An amorphous Ni-P alloy foil was used as substrate material, based on the idea that the amorphous substrate does not affect the crystal orientation direction of electrodeposits and therefore films can be generated which are independent of the nature of the substrate. Results using this substrate were compared and contrasted with those obtained using various singZe crystal copper surfaces. The crystallographic structure of the electrodeposited films was determined using X-ray diffraction; the orientation index was calculated from Wilsons formula; and the crystal size was calculated from Sherrers formula. The solution used for the pure tin plating was 0.15M SnCli + I(4P207 (complexing agent). The solution pH was adjusted to 6 , 8 or 10 using HC1 and ammonia. The solution was stirred. The electroplating procedures were performed within a range of current densities from 5-50 mA cm-2with deposit thicknesses ranging from 0.84-62.4 pm. The amorphous Ni-P film was formed on the surface of pure polycrystalline foil. Copper single crystal electrodes having surface Orientations of {loo}, { 1lo} and { 111} were also studied. Figs 5 , 6 and 7 show SEM images and the effect of current density on the orientation and grain size of the Sn electrodeposited films at pH 4, 6 and 8, respectively, on the Ni-P amorphous substrate.

a
8

e
a

e
6
0 0

0
0 0

( ] F
1 80 5
60
40
20
0

0
0 0
X
Q)

U t :
.I

6
5

0 c
.I

: .

Q)

0
0

2 c .2
c1

4 3
1

t .-:

6 2
0

a
a
0
0 0

30 50 Current density / mA/cm2


10 20

Fig. 5 . SEM images (upper) and relationship of orientation index and grain size to current density (lower) [2].

5mA/cm2

20mA/cm2

50mA/cm2

0
X e,
'0
.e

7.e 0

- 60

:e, .

5
~

. . .-----*

10 20 30 50 Current density / mA/cmZ


-.-- --- -- -. -T--

Fig. 6. SEM images (upper) and relationship of orientation index and grain size to current density (lower) [2].

0 0 0 0 0

5 0 m A/cm
100
X
Q)

C .-

' p

Grain A-, -. size


n

x, 220
0\
\

-po
0 0
0

-4
- 60

p
: .
e,

c
0 .-, c .
c ,

td C

\.

.e

a ,

'\,
A

0"

i i
211

- 40

2oo
20

rs
.e

0 0 0 0 0 0 0 0 0 0 0
0 0

10 20 30 50 Current density / mA/cm2

Fig. 7 . SEM images (upper) and relationship of orientation index and grain size to current density (lower) [2].

10

a
0
4 B
0
The results in Figs 5-7 show that the grain size of the deposits remained approximately constant in the range 4-5 nm and therefore independent of either current density or pH. Furthermore, the current density did not influence the orientation of the deposits appreciably, except at pH 10 where the (220) plane was clearly the preferred one at current densities below 20 mA cm-2.However, the morphology of the deposits was clearly affected by current density at all pH values, becoming smoother with increasing current. Fig. 8 shows SEM images of the Sn electrodeposited films formed on the copper single crystal substrates. Here it can be seen that the surface morphology was dependent upon the crystal orientation.

e e

e
0

0
0

e e
0

e e e
0

a
0

0 . 4 ~

1 . 7 ~

3.4p

5.lpn

Fig. 8. SEM images of deposition on copper single crystal electrode surfaces [2].

0
0
2.3 Preparation of Dendritic Tin Nanoagqegates

a
0

e e
a
0 0

Tin coating on insulating substrates plays a very important role in electronics and often requires techniques such as chemical vapour deposition, laser ablation, or screen printing. A new technique recently suggested by Fleury [31 allows the formation of metallic coatings on insulators using electrodeposition. Devers et al. [4] have recently described the use of this technique to obtain tin nanoaggregates on an insulating surface. The nanoaggregates exhibit a dendritic structure composed of polycrystalline grains with grain size down to 20 nm. The microstructural characterisation was achieved by scanning electron and atomic force microscopy. The measurement of potential-time curves allowed the onset of nucleation to be determined. The electrochemical cell is composed of two glass plates (microscope slides) separated by foils of copper and tin acting respectively as anode and cathode (fig. 9).

a
9

11

0
0

0
Glass plates

0000
20 A

0 0 0

I
Copper foil (cathode)

I
Tin foil (anode)

I
(b)

0
Fig. 9. (a) schematic view of the cell. (b) Glass plates with a gold layer (20 and 1000 A) [4]. The deposition conditions facilitate the growth of a very thin tin film starting from the cathode and proceeding horizontally towards the anode. The 12.5 pm thickness of the metal foils also determines the electrolyte volume. The thickness of the cell is much greater than the final thickness of the deposit. The lower glass plate is covered with a 1000 A gold coating at both its ends and covered with a thin gold coating (20 A) in its centre. The 20A gold coating is in the shape of small islands which do not percolate (Fig. 10).

0 0 0

0 0
0
0

a
0

0
0

a
0 0 0

0
Fig. 10. STM image of 20 A thick gold layer on glass slide showing that the layer is discontinuous [4]. The experimental conditions require the electrodeposition to proceed along the surface. The 1000 A coating is necessary to initiate the nucleation and growth of the tin layer. It is possible, however, for growth to proceed on the surface even without this coating. For example, growth may, although rarely, proceed along the opposite uncoated glass plate. The non-conducting 20 A gold coating, which is present all across the slide to be coated, is used as an activation layer to enable the electrolytic deposit to nucleate on the surface as the deposit progressively invades the glass slide. It also improves the adherence, which allows the recovery of the deposit after opening-thecell and rinsing the electrolyte. The gold coatings are obtained by evaporation in a custom evaporator jar. In different operating configurations, the lower plate can be one of two types. The
12

0
0

a
0
0 0
0

0
0

20.A gold layer is either uniformly deposited or it contains four 1000 A thick contacts in the shape of circular dots located at approximately 1 cm from the cathode for further electrochemical or SEM characterisations. The electrolyte used can be SnC12 and lo-' mol 1 or SnS04 at concentrations between

-'.

Experiments were aimed at forming dendritic deposits covering the glass slide. moVl SnC12 at 10 mA cm-2.These Fig. 11 shows the tree-like deposits formed in were composed of small grains in the 50-100 m range, themselves deposited on a sub-layer of grains in the range 5-50 m.

Fig. 11. SEM image of tree-like tin deposit formed in

mol/l SnC12 at 10 mA cm-*[4].

The authors concluded that their experiments pave the way to possible applications of tin thin films obtained directly on insulating substrates by electrodeposition, ie. galvanic rather than electroless routes. More specifically, the work was claimed to help the understanding of the activation properties of gold particles used as nucleation sites. Similar particles of Pd are commonly used in electroless deposition for metal plating. 2.4 Underpotential Deposition of Tin

Metal underpotential deposition (UPD) is a process that takes place at potentials positive of the respective reversible (or Nemst) potential. Depending on the interaction and the crystallographic misfit between the depositing metal and substrate, the miscibility of the two metals in the bulk, and the electrode potential of interest, UPD may take place with different mechanisms. These include: sub-monolayer adsorption at high potential shift; surface reconstruction (or surface confined alloyng) after prolonged polarisation; and alloy formation at low potential shift. With the rapid development and application of scanning tunneling microscopy (STM) and other structurally sensitive techniques such as grazing-angle X-ray diffraction, considerable progress has been achieved in understanding the physical and chemical nature of UPD in a microscopic perspective. Yan et al. [5] have recently found that Sn UPD on Au {hkl} differs from most of the previously studied UPD systems in that all three classes of UPD mechanism can be observed depending on the specific conditions applied. These authors have used
13

cyclic voltammetry coupled with in situ STM to study Sn electrodeposition on Au {loo} surfaces at underpotentials. The UPD proceeded in three stages. Firstly, the reversible monolayer UPD; secondly, surface alloying; and finally, alloying that extends into the bulk. The formation of the reversible sub-monolayer UPD proceeds in the form of clusters of 2 nm at potentials of 0.40 V vs. SCE. The surface alloying takes place at -0.40 V, evidenced by a sudden immersion of the Sn clusters as well significant changes in steps on the surface. The bulk alloying sets in at -0.30 V, which leads to a rapid change of surface morphology. The facilitated surface and bulk alloying processes were attributed to enhanced interdiffusion of Sn and Au atoms resulting from changes in potential. The authors concluded that Sn UPD on Au {loo} is a much more complex process than that observed previously for other metal UPD systems because a variety of different mechanisms can be pertain.

3.

Role of Additives in the Electrodeposition of Tin

Plating baths used for the electrodeposition of tin and its alloys usually contain various additives favourable for producing coatings of high quality. Optimal conditions for deposition can be realised by adding certain compounds that exert a selective influence on partial processes. The role of additives in the electrodeposition of tin has been discussed recently by Barry and Cunnane [ 171. These authors provide a short bibliography describing the chemical nature of some of the additives that have been investigated to date. Tin is electrodeposited with little or no activationkharge transfer overpotential (an overpotential required to initiate deposition of the metal from acid solution in the absence of additives). The deposits obtained from these electrolytes are porous, coarse and non-adherent. It has long been known that addition of organic compounds to the electrolyte results in an increase in the overpotential for tin reduction. Organic additives produce a compact, finely crystalline adherent electrodeposit. The development of bright electrodeposits commenced over 70 years ago. Until recently, the bulk of knowledge on brightener formulation has been proprietary. A variety of organic additions can be identified from the patent literature, all of which operate as brighteners and many can modify the properties of electrodeposited tin. Developments in understanding the mechanism of action of these substances and their effectiveness are discussed below.

3.1

Adsorption Characteristics

The adsorption of certain additives on the metal is usually regarded as an essential prerequisite for influencing the morphology of electrodeposits. Recently, Survila et al. [6] have published a detailed account of the adsorption behaviour of Laprol2402 C on a tin electrode in strongly acid SnS04 solutions. Laprol2402 C is a copolymer of ethene and propene oxides and has found application as an effective component for producing bright coatings. For example, such polymers, containing -(CH&,-Ochains, are in common use in bronze plating. The electrochemical techniques of voltammetry and a.c.impedance spectroscopy were employed to investigate the deposition processes.

14

0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

8
0 0
0
Addition of Laprol did not affect the open circuit potential but gave rise to a decrease in current density throughout the entire region of cathodic polarisations (Fig. 12).

0
6

0
0

5
E

:
2

k----l.;
cbp= 0 5 mg dm'3
100

e e
Q

-0,251

-0.30

-0,35
E/V

-0,40

-0,45

. -

Fig. 12. Effect of laprol concentration on the voltammetric characteristics obtained on a rotating tin electrode (440 rpm) [61. These results showed that the rate of Sn(I1) reduction was significantly affected as a result of the adsorption of Laprol on the tin surface. This fact was demonstrated in another way in Fig. 13, which illustrates the reduced influence of forced convection as the concentration of Laprol was increased.

e
0

e
0

10

0
Q
N

e
0
0

<
E

S 6

: 4

'
0

790
I

qap 10 mg dm-3 =
I
I I I

e
0 0
0 0

-0.3

-0,4

-0,5
E/V

-0.6

-0.7

Fig. 13. Voltammograms obtained for solutions containing 0.5 and 10 mg dm-3Laprol. Rotation speeds of the rotating disc electrode are indicated on the curves [ 6 ] .

An analysis of impedance data together with the voltammetric data led to the
conclusion that at least two states of adsorption occurred depending upon the cathodic polarisation. Adsorption of Laprol observed at low or sufficiently high cathodic polarisations resulted in a moderate decrease in the double layer capacity from 100 to 18 pF and was accompanied by a 100 fold reduction in the exchange current. At -0.30V VS.Ag/AgCl/sat KC1, a hrther reduction in the double layer capacity occurred to 7pF ~ r nindicative of a more compact surface layer. -~

e
0 '0 0

e
m
15

0
3.2 Sodium Gluconate

3.2.1 Long-term Deposition Abd El Rehim et al. have developed plating baths from which tin can be electrodeposited from gluconate solutions [71. The advantage of these baths is not only their cheapness, but also their non-polluting effect on the environment. The metal was electroplated onto steel substrates from baths containing SnS04, &SO4 and CbH1107Na under different conditions of bath composition, pH, current density and temperature. A detailed study was made of the influence of these parameters on potentiodynamic cathodic polarisation, cathodic current efficiency, microstructure and the morphology of the deposits. Fig. 14 shows the surface of some tin deposits obtained from a gluconate-free bath and from a gluconate-containing bath. In the absence of gluconate ions, shapeless and dendritic-grained deposits were obtained. The deposits were rough, not compact and the surface of the substrate was visible. The presence of the sodium gluconate in the bath substantially improved the quality of the deposits. In this case, smooth and compact deposits were produced covering the whole surface of the substrate and consisted of 'block-like' crystals. It was postulated that this improvement in the quality of the deposits was related to the complexation of tin by gluconate ions that inhibit the outward growth of the grains but allow lateral growth.

0 0 0 0 0 0 0 0 0 0 0 0 0

0 0 0
0 0 0 0 0 0 0 0

Fig. 14. Scanning electron micrographs of tin deposited from: (a) bath composition consisting of 40 g/1 SnS04 + 10 g/l K2S04, pH 1.9, I = 5.6 mA cm-2, T = 16 "C and t = 15 mins, (b) as previous composition + 70 g/l sodium gluconate [7].

The cathodic current efficiency increased with increase in tin content in the bath but decreased with increase in pH, temperature and current density. X-ray diffraction analysis and anodic stripping voltammetry proved that the tin deposit consisted of one phase (P phase) with a tetragonal structure. The optimum plating conditions were
16

found to be 50 g/1 SnS04 + 10 g/1 K2S04, pH 1.6, I = 5.6 rnA cmm2, = 22-40 "C, with T the addition of 70 g/1 sodium gluconate.

3.2.2

Short-term Deposition in Sodium Gluconate Solutions

The initial stages of tin electrodeposition from sulphate baths onto a vitreous carbon electrode in the presence of gluconate has been investigated by Torrent-Burgues et al. [81 using a combination of electrochemical and microscopic techniques. The gluconate sulphate bath contained 0.01 M or 0.016 M SnS04, 0.2-1 M Na2S04 as supporting electrolyte and 0.06-0.2 M sodium gluconate as chelating agent. The pH was adjusted to 4. In addition, a highly acidic sulphate/gluconate bath was studied which contained 0.01 M SnS04, 1M H2S04 and 0.06 M sodium gluconate. The working electode was polished mechanically before each experiment with 3.75 and 1.87 pm alumina powder followed by a short electrochemical conditioning. The use of the sulphate/gluconate at pH 4 shifted the tin electrodeposition process to more negative potentials. The deposits were uniform, in contrast to those obtained in the strongly acid bath, thus demonstrating the potential of sulphate /gluconate baths at pH 4 as useful candidates for metal finishing applications. Fig. 15 shows the effect of pH in baths containing gluconate and also the preference to form cubic crystals in the presence of bath agitation. This situation contrasted with that in the absence of agitation where tetragonal crystals were preferred. When comparing similar overpotentials in baths with and without gluconate, the compound was found to favour the nucleation of tin but to decreases the growth rate of the tin crystallites. Experiments perfomed under varying conditions of tin and gluconate concentrations, revealed the influence of these factors in the overall nucleation and growth mechanisms. The electrodeposition process at the initial stages of the process occurred via instantaneous nucleation with 2D growth control, but the presence of higher gluconate concentrations favoured a change in the mechanism towards progressive nucleation with 2 D growth control. Fig. 16 shows potentiostatic i-t transients for the pH 4 solution in the presence of gluconate and Fig. 17 illustrates the way the data was analysed to reveal the two dimensional mechanism at low to medium gluconate concentrations. It is interesting to note that Gomez et al. [ 11 reported instantaneous nucleation and three dimensionaZ growth controlled by difhsion for the electrodeposition of tin in highly acidic non-gluconate containing sulphate baths (Section 2.1). Consequently, gluconate has an important effect during the initial stages of the deposition process.

17

..

. ~. ..

..-,.....,

.-

I_

..

0 0

0
0 0

0
0 0 0
0

0 0
0

0
0 0 0

0 0
I

0 0 0
0

0
0 0 0
Fig. 15. SEM micrographs of tin electrodeposits obtained with agitation in the presence of 0.06 M gluconate: (a) high pH solution at -530 mV VS. AgIAgC1 (3M KC1) for 100s; (b) and (c), low pH solution at -680 mV for 300s [SI.

0 0 0
0

18

0 0
0 0
0.25

0.2

0
0 0
a
.\

0.15
r

E
1

0.1

a a
0
0

0.05

a
0

6
tI s

10

12

a
0

Fig. 16. Potentiostatic transients for solution 0.01M SnS04,0.8M Na2S04,0.12M gluconate, pH 4. The potential transients were made from Eo= -200 mV to: (a) -675, (9b) -700, (c) -750, (d) -775, (e) -800 and (0 -900 mV vs. Ag/AgCl (3M KCl) [8].

0 0
4

0
0

0
0

.-4.8

Y 1

0
0 0 0

rn 2

a
0 0
-25

-5.6

-2

-1.5

-1 -0.5 log ((t-to) /s)

0.5

e e
0

Fig. 17. Plot of log i vs. log(t-to) for the curves of Fig. 16 [8].

19

3.3

Cetyltrimethyl-ammonium Chloride and Cetylpyridinium Bromide

The influence of the above quaternary ammonium salts on the electrodeposition of tin onto low carbon steel in acidic electrolyte baths has been investigated by da Silva et al. [91. Cyclic voltammetry, galvanostatic polarisation and scanning electron microscopy were employed and the electrolyte composition was 1.4 x lo-' M SnS04, 9.2 x lo-' M p-phenol sulphonic acid, 0.1 M &So4 and 8.0 x lo4 M additive. At low current densities in the range 10-20 mA cm-2,dendritic growth occurred in solutions containing cetyltrimethylammonium bromide (CTAB), but at higher currents in the compact and smooth deposits were formed, cf. Figs. 18 and range 50-100 rnA cmm2 19. Cetylpyndinium bromide was, however, found to be less effective in producing this type of deposit (Fig. 20).

Fig. 18. Micrograph of tin electrodeposit obtained at 20 mA cm-2in the presence of additive CTAB. qdep = 455 mC cm-2 [9].

Fig. 19. Micrograph of tin electrodeposit obtained at 50mA cm-2in the presence of additive CTAB. qdep = 455 mC cm-2[93.

Fig. 20. Micrograph of tin electrodeposit obtained at 50mA cm-2in the presence of additive [9]. CpyB. Q d e p = 455 mC
I

20

0 0

0
0 0
3.4 S-dodecyl-mercaptobenzimidazol

0 0 0 0
0

Bakkali et al. [ 101 have very recently investigated the effect of the above additive (code name M12) on the electrochemical and morphological charcteristics of tin electrodeposition in acid sulphate solutions. It was shown that in the presence of M12, the overvoltage for hydrogen evolution was markedly increased. Tin discharge was controlled by mass transfer and the difhsion coefficient of Sn(n) species was smaller than in pure acid sulphate solutions. The deposit morphology was much smoother and brighter and no dendritic growth was observed over a wide potential range. This was associated with a change in the preferred orientation of the tin films. The new organic additive, which may also be used as a corrosion inhibitor or fungicide, has the formula C ~ H ~ N ~ H C - S - ( C H ~ ) Iand is a double ring compound 1-CH3 (one aromatic and one heterocyclic) with a side chain. It is insoluble in water and must be dissolved in ethanol before adding to aqueous electrolytes. The basic electrolyte used in this study consisted of 0.14 mol dm-3SnS04 and 0.56 x mol dm-3H2S04 (electrolyte A). The additive was present at a concentration of 0.156 x 10-3mol dm-3and in the presence of the additive the electrolyte also contained 20 % ethanol (electrolyte C). In the absence of the additive, but in the presence of ethanol, the electrolyte was designated electrolyte B. To characterise the deposition of tin a voltammetric investigation was camed out using an electrochemical quartz crystal microbalance (Fig. 2 1). The response of the EQCM experiments showed that the deposition commenced at a potential close to -450 mV (vs. SCE) for all the solutions. For electrolyte A, the mass increased rapidly with polarisation up to -700 mV (curve 2A). For more negative potentials the mass increased no further because of the formation of non-adherent dendrites, which were stripped off the electrode. By contrast, in electrolytes B and C, the mass increased regularly (curves 2-B and 2-C) and adherent deposits were formed.

0
0
0

0
0 0

a
0

a
0 0
0

0 0 0
0 0 0
0
-900

-800

-700 -600 Potencial, mV,,,

-500
- - -. .

-400
._

Fig. 21. Cathodic polarisation results with EQCM (scan rate 10 mV/s); Curves 1-A, B, C: current density responses for electrolytes A, B and C, respectively.; Curves 2-A, B, c: mass change responses for electrolytes A, B and C, respectively [ 101.

0 0 0 0 0

The cathodi'c current efficiency was calculated from the results of the EQCM measurements using Faraday's Laws. In the absence of the additive, the efficiency was about 80% at a potential of -600 mV, but decreased at larger overpotentials (Fig, 22 curve 1). For electrolytes B and C, the efficiency remained constant at more

21

negative potentials, especially for electrolyte C, containing both alcohol and M12 additive (Fig. 22 curves 2 and 3).
100

0 0 0 0 0 0 0 0 n v 0 0 0 0 0 0

t.

1I

'

-1 000

-900

-800
Pote ncia I, mV,,,

-700

-600

Fig. 22. Faradaic efficiencies as a function of polarisation; Curves 1,2,3 recorded for electrolytes A, B and C, respectively [ 101. The partial currents for the deposition of tin and for hydrogen evolution were calculated from the results of the EQCM experiments. Fig. 23 shows that in baths B and C at -600 mV the partial current for tin deposition was approximately half that measured in the basic electrolyte. At -700 mV the partial current was one third of the basic electrolyte value. This demonstrated that alcohol and M 12 had an inhibiting effect on tin deposition. Fig. 23(b) clearly demonstrated that in the presence of the additives, the overvoltage for hydrogen evolution was markedly increased.

-15 -900
I

-800

'

'

-700

'

"

-600

'

-500

-400

Potencial, m,, V,

0 0 0 0 0
0 0
0 0 0 0 0 0

-151 " -900

"

-800

"

"

-700

"

"

'

-600

"

"

-500

"

"

-400

Potencial, m,, V,

Fig. 23. (a) Partial current density for tin deposition; curves 1,2,3 recorded for electrolytes A, B and C respectively. (b) Partial current density for hydrogen recorded for electrolytes A, B and C, respectively [ 101.
22

0 0
0

0
0

e
0
0

Morphological studies were also undertaken. Tin layers l O p n in thickness were deposited at 4.OA dm-2on an iron electrode disc rotating at 2000 rpm. They were then examined by X-ray diffraction and electron microscopy about a week after preparation. The least textured sample came from bath B containing alcohol. Bath A without additives showed a strong (200) orientation whereas bath C containing M12 exhibited a strong { 101} orientation. The grain sizes were 100, 70 and 55 nrn for baths A, B and C, respectively. Fig 24 shows the morphologies of the films. Those for electrolytes A and B were rather irregular, whereas that for electrolyte C containing M12 was-more regular and smooth. The films were quite stable and no dendrite formation was observed even after long storage periods at ambient temperature in the laboratory atmosphere. 3.5 Miscellaneous Additives

a
0

e a
0 0
0

a a a
0

In a series of publications, Medvedev et al. [ 11-16] have explored the use of a wide variety of organic additives with the primary objective of locating compounds capable of imparting lustrous qualities to tin electrodeposits. These are described as either lustre forming or simply as inhibitors. For example, the electrodeposition process of tin from electrolytes containing SnS04, sulphuric acid and organic additives such as synthanol,formalin and benzyl alcohol, was investigated on copper electrodes [ 1I]. The Zeveling capability of the electrolytes and the degree of brightness were determined. An influence of the organic additives on the microdistribution of electrodeposited tin was also shown. Lustrous coatings with leveled surfaces can be produced using an electrolyte containing 5-50 g/1 SnS04, 90-100 g/1 H2S04,2-3 g/1 synthanol, 6-8 g/1 formalin and 6-8 g/l benzyl alcohol. Variations on this theme include the use of propargyl alcohol [ 13 ] and coumarine [ 16 ] in place of benzyl alcoho 1. 3.6 Synernistic Effects of Additives

a
0

a
0 0

Using a combination of electrochemical techniques, including cyclic voltammetry, potential step and a.c.impedance methods; Barry and Cunnane [ 171 have recently presented what they claim to be the first quantitative effort at determining the synergistic effects of organic compounds on the discharge, nucleatTon and growth mechanisms of electrodeposited tin phases at galvanostatically prepared polycrystalline copper electrodes. These authors have commented that in previous investigations on tin electrodeposition in the presence of organic additives that confer brightness, there is an absence of kinetic data, thermodynamic data and information on the discharge, nucleation and growth mechanisms. Conversely, in those papers containing kinetic data, the additives have not necessarily been shown to confer brightness, but are reported to have a grain refining effect. Therefore, relevant information regarding additive structures that confer brightness is lacking. Barry and Cunnane have determined cathodic and anodic transfer coefficients, exchange current densities and the rate constant for the discharge of Sn(II) in HzS04 solutions. A variety of additives were studied including t-octylphenoxy ethanol, formaldehyde and phenyl-2-butenalimine, either separately or in all combinations. The discharge mechanism of Sn(I1) remained a two step process with the transfer of thefirst of the two electrons being the rate determining step. The organic additives
23

e
0

a
0
0 0

had no effect on the electrochemical mechanism. However, it was shown that the additives altered the nucleation mechanism of tin electrodeposition. In all cases, hemispherical nuclei were formed on the electrode surface. This is predicted to occur when the interfacial tension between the substrate and the deposit and that between the substrate and the solution, are the same. Most often, growth of spherical clusters is accounted for theoretically by slow ion transfer across the cluster/electrolyte interface. The mechanism by which these growing centres developed was found to change from instantaneous 3D growth to progressive 3D growth when t-octylphenoxyethoxy ethanol was removed from the electrolyte formulation. It was postulated that this resulted from a steric effect of the former additive. Although no microscopical data was given in the paper, it was stated that the presence of t-octylphenoxyethoxy ethanol facilitated the production of bright deposits and therefore it could be concluded that the observation of an instantaneous nucleation mechanism is a prerequisite for obtaining such deposits. However, a truly bright deposit was apparently obtained only in the presence of all three additives and hence a basic understanding of the system is still somewhat lacking.

4.

Development of Plating Baths Based on Sulphonic Acids

The development of plating baths based on methane sulphonic acid and metal methanosulphonates has been reviewed by Guhl and Honselmann [ 181 and Balaji and Pushpavanam [ 191. Methanesulphonic acid (MSA) based plating formulations have gained wider popularity over the past decade especially in electroplating related to the electronics industry. They are replacing the conventional fluoroborate baths, are relatively less harmful and their waste can easily be treated. In addition they are stable in acidic, neutral and alkaline solutions and do not undergo any appreciable hydrolysis, regardless of the temperatures used. Since MSA is less aggressive in the shop environment, no special requirements are necessary. The baths produce less wastage and sludge formation and cause less oxidation of metallic species present in the solution. It is also possible to use inert anodes and thus the initial investment or capital can be minimised. The baths are not aggressive to ceramic and glasses and produce nodular-free deposits that are the main requirement. A comparison of the properties of methanesulphonic acid with sulphuric acid [ 191 indicates that MSA is less oxidising in nature. This property is very much desired in the case of plating of those metals that exist in multivalent states. Hence, these baths offer the advantage of using insoluble anodes. MSA is a strong acid (pK, = -2) and is completely ionised at concentrations of 0.1 mol dm-3in aqueous solutions. Although MSA baths have a lower equivalent conductance than hydrochloric or sulphuric acids, this is compensated by a higher solubility of its salts [ 191. This enables the bath to be operated at very high current densities as in high speed plating and continuous plating which in turn increases productivity. MSA also offers higher solubility for surfactants and other organic additives.

6 0 0 0 0 0 0 0 0 0 0 0 0 0 6 0 0 0 0 0 0

c
0 0 0 0 0 0

24

0 0 0 0

-.

,-

.-

a
0

0 0
0

a a a
0

Finally, MSA baths are less toxic than counterparts such as fluoroboric and fluorosilicic acids, they generate less pollution and are safer to handle. Above all, MSA is biodegradable, ultimately forming sulphate and carbon dioxide and it is recyclable up to 80%. In general, the environmentally benign nature of MSA, especially when compared to the HF complex acids, makes it an environmentally advantageous electrolyte. Balaji et al. [ 191 also discuss the particular advantages of MSA based electrolytes in tin plating. Two types of electrolytes have commonly been used. One based on phenolsulphonic acid popular in Japan and the other a halide-based electrolyte popular in the USA. The former generates no chloride or fluoride sludge but has the drawback of poor conductivity. Conversely, the latter achieves high productivity due to excellent electrolysing properties, but requires the disposal of voluminous amounts of sludge accumulated in the bath. By comparison, the new MSA bath with special additives enables uniform plating over a wide range of current densities and generates virtually no sludge. The versatility of this bath allows the same bath composition and additives to be used for barrellstill plating and high speed reel-to-reel plating. Despite the above advantages, there are a few disadvantages with MSA baths. They offer a high plating efficiency at current densities above 30 A dm-2,,but a lower efficiency at lower current densities. The coating morphology and preferred crystal orientation also exhibits a greater variation with operating conditions, especially below 6 A dm-2.This means that the baths are usually operated at current densities exceeding about 10 A dm-*. The following discussion is concerned with recent investigations to elucidate the advantages of MSA electrotinning processes in more detail. 4.1 Comparison of Halogen and Methanesulphonic Acid Electrotinning Processes

a
0
0

a
0 0
0

a
0

e
0 0

The characteristics of a typical halogen and a patented methanesulphonic acid based electrotinning process have recently been compared in a quantitative manner by Yau [201 The halogen electrotinning process, developed by E. I. Du Pont de Nemours & Co. and Weirton Steel Co., has been used successfully in the production of tin plate in continuous steel strip plating lines for more than 50 years. Because of increasing environmental concerns about the ferrocyanide containing sludge generated by the halogen process, users of the process are considering other, environmentally friendlier processes. The MSA based process has been used for years in the electronics industry and has also been developed for continuous steel strip plating lines. In addition to being influenced by the proprietary additives in MSA baths, the plating efficiency and the coating quality also depend upon the operating conditions of line speed and current density. It was therefore necessary for Yau [ 191 to compare the two processes under the operating window of a halogen line, where the steel strip travels at speeds of between 2 and 10 m / s and where the current density can vary from 11 to 65 Ndm2to encompass the product mix. The hydrodynamic conditions created by a fast moving steel strip cannot be simulated easily in the laboratory. However, an approximation to these conditions could be obtained by the use of a rotating cylinder electrode.

a e a
0

a
0
0

0 0

25

. ..

. -.. .-.- .. .

:.

._.. .^.. .-. .

_ . . ~.

The MSA bath consisted of 50 ml/l stannous methanesulphonate, 30 ml/l methanesulphonic acid, 100 mV1 grain refiner and 20mV1 anti-oxidant. (These were all proprietary products, Ronastan TP, Lea Ronal). The halogen bath consisted of 16.5g/l stannous fluoride, 44.3 g/1 stannic chloride, 43.4 g/1 sodium bifluoride, 21.5 g/1 sodium hydroxide, 3g/l sodium ferrocyanide and 2ml/l brightener known as addition agent 20. In a halogen line, the steel strip is plated only on the lower side as it travels through the horizontal plating cells. As a result, the upper side of the strip is subject to corrosion by the plating solution. The dissolved iron ions greatly increase the oxidation rate of the stannous ions in the halogen bath. A rotating disc electrode was used to evaluate bath corrosiveness as well as its stability in the presence of dissolved iron. The results showed that the MSA bath was very stable with a low oxidation rate of stannous ions, but was much more corrosive towards the steel substrate. The effects of operating conditions, such as line speed and current density onplating eflciency, coating morphology, and preferred crystal orientation were investigated using the rotating cylinder cathode. Comparison of the plating efficiencies obtained from the rotating cylinder cathode and production identified a window of test conditions that was representative of production conditions. Within this window, the MSA process offered a higher plating efficiency at higher current densities, but different microstructure and preferred crystal orientation compared with those of the halogen process. The coating morphology in the MSA bath also exhibited a greater variation at low current densities, resulting in a narrower operating window. In both processes the grain size increased in directions parallel as well as noma1 to the steel substrate. This three dimensional grain growth was accompanied by a continuing change in the crystal orientation, suggesting that recrystallisation takes place during these two electrotinning processes. The above conclusions concerning plating efficiency, coating morphology and preferred crystal orientation can be seen more easily in the following figures. Fig. 24 compares the plating efficiencies in the halogen bath from the rotating cylinder experiments with those obtained from a production trial. It can be seen that the current density had the same effect on the plating efficiency, i.e. the plating efficiency decreased at the same rate as the current density increased. Moreover, the plating efficiencies from production at 3-9 m/s linear speed were comparable with those from the rotating cylinder experiments with a rotation speed of 500-2,000 rpm. Therefore, it was concluded that the operating conditions in the halogen line could reasonably be simulated by the rotating cylinder approach. Fig. 25 compares the plating efficiency of the two processes at two representative rotation speeds of 1,000 and 2,000 rpm. It can be seen that the plating efficiency in the MSA bath could be up to 10% higher than that of the halogen bath at high current densities, i.e. above 30 A/dm2. The plating efficiency in the halogen bath was, however, higher at low current densities, i.e. below 12 A/dm2.

0 0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

26

100

0 0
80

e a
0 0 0 9 0
0

5
.m

60

w
.m

40

d
J

z
20 0 0

--++RCC

. s

Line Trial 1200 fpni Line Trial 600 fpm RCC 2000 rpm
1000 rpm

- 5 RCC 500 rpni B -

10

20

30

40

50

60

70

Current Density, A/din2

Fig. 24. Comparison between plating efficiencies obtained in a production line trial with those obtained with a rotating cylinder cathode in the halogen bath [20].
100

e
0
0

s
5

80

e
0

.- 60 i w

-z

40

e
0

20

0 0 10
20

e
0

30

40

-50

60

70

Curent Density, Ndni

e
0 0
0 0

Fig. 25. Comparison between plating efficiencies obtained for the MSA bath operating within the window representative of production conditions [20].

e
e a
0

Fig. 26 shows a scanning electron micrographs of the coating plated in the MSA bath as a function of the electrode rotation speed and current density. In general the morphology of the coatings could be divided into two groups diagonally. In the lower right half (high rotation speeds and low currents), the deposits were compact and fine grained. In the upper left half (low rotation speeds and high currents), the deposits were predominantly columnar dendritic. The halogen bath exhibited the same characteristics except there was less variation in the morphology in the fine grain region. These effects were discussed quantitatively in terms of the juxtapositiodratio of the applied current to the diffusion limited current. Increasing the stannous ion concentration, the rotation speed or the temperature all increase the latter current and this widens the current range for a good coating.

e
Q
27

W 0

c W 4

0 0

SO0

1000

2000

3000

Rotation Speed, rpm

Fig. 26. Effect of current density and rotation speed on the morphology of tin deposited from the MSA bath [20]. Fig. 27 shows the preferred orientation of coating plated at 18 A/dm2 in the MSA bath as a function of the rotation speed. Such a preferred orientation is essentially equivalent to the percentage of crystal grains with a particular orientation parallel to the steel surface. At 0 rpm, the dendrite is characterised by the preferred orientation of (220) with three times the theoretical intensity, followed by (220}, (420), ( 1lO}, etc. As the rotation speed increased to 500 rpm, (200) diminished and (220) became the preferred orientation, accompanied by the emergence of (32 1) and ( 112). As the rotation speed further increased to 3,000 rpm, (112) gradually became the preferred orientation, primarily at the expense of (220). Compared with the coating morphology (Fig. 26), the diminution of (200) coincided with the disappearance of dendrites. Fig. 28 shows the equivalent data for the halogen bath. It is clear that the distribution of preferred orientations and the changes with rotation speed are different.

0 0 0
0 0 0 0 0 0 0 0 0 0

28

,i

i'

gp"
. +
U

g
0

-s 8
s
. c
h)

* 3z
r
Inverse Pole Figure Intensity Ratio
w
P

0
Y

H ,

Inverse Pole Figure Intensity Ratio


C-L

0,
r O

g
O r H W P m m 4 w

5.
I

cn

c l . a

\c

=
l l l l l l l

l
l l

+i!
CD

+i!
CD

a
p.

C I I I I I I I I I I
01

92
5 . 3

x L a
g
0

a 4
3
1
w

U 1

3
1 1 1 1 1 1 1 1 1 1
0

g.
Za

E
0
h

5. a
n
W

U
h

n
W

h ,

p.

0 h ,
v
w

c-r l. -i t 4
.

0 t4
w

i2 w
J

I
..

5'

.g

5' cc
CD

=f

times the theoretical intensity at the expense of other orientations except for (321) and (420 }, both of which remained constant.

_-

10

9
0

B (312)

Cl(41l)

.g
E
U U

a (420)
7

B31 (2)
6

n(iiz)
(301)
BiB (211)

(220)
rn(101)

D20 (0)
2

i
I

10

IS

20

25

30

Plating Time, seconds

Fig. 29. Effect of plating time on crystal orientation of tin deposited at 42 Ndm2and 2,000 rpm in the MSA bath [20]. Fig. 30 shows scanning electronmicrographs of the cross-sections from cryofractured samples. The grains were about 0.4 pm thick and 1 pm in diameter after two seconds of plating and subsequently increased progressively to 4 pm thick and 5 pm in diameter after 20 seconds. The grain size increased not only in the direction normal to the electrode surface, but also laterally, parallel to the substrate. While the increase in thickness was proportional to the plating time, the increase in diameter gradually slowed, resulting in near-equiaxed grains. The same pattern of grain growth was observed with coatings plated from the halogen bath. This three-dimensional grain growth, accompanied by continuous changes in preferred orientation, suggested that the deposited tin recrystallises during the plating process.

20 seconds

10 seconds

2 seconds

Fig. 30. Three-dimensional crystal growth of tin deposited at 42 Ndm2and 2,000 rprn in the MSA bath [20].
30

0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

0 0
0
4.2 Effect of Process Variables on Electrotinning in Methanesulphonic Acid

0 0
0

0
0

e
0 0 0 0 0 0 0 0

In later work, Yau has considered the effect of some of the process parameters discussed above in more detail [21]. The plating bath (pH 0.4) had a similar composition to that discussed above. The effects of agitation, current density and temperature on the plating efficiency and coating morphology were examined using the same rotating cylinder and rotating disc techniques. A good correlation was found between the plating efficiency and coating morphology. Moreover, both of them were affected by the diffusion limited current density (iL)of the stannous ion. A compact crystalline deposit was obtained with > 95% efficiency in the plating bath. When the current density exceeded i, , not only did the plating efficiency begin to decrease, but also the coating changed to micronodular and then to dendritic.
Fig. 3 1 exemplifies the effect of rotation speed and current density on the coating morphology at 52 "C. The coating morphology could be divided diagonally into three groups as previously: (I) compact crystaZZine at the lower right region of high rotation speeds and low current densities, (11) dendritic at the upper left region of low rotation speeds and high current densities, and (111) micronodular between the compact crystalline and dendritic regions. This scheme falls into the general scheme proposed by Winand [22]. Furthermore, the coating morphology was found to be well correlated with the plating efficiency. As shown in Fig. 32, the plating efficiency, excluding those at the extremely low current density of 60 A/m2, was above 90% in the compact crystalline region, between 80 and 92% in the micronodular region, and much lower on average in the dendritic region.

a
0 0

e
0
Y

r:

1800

8
GO0

a
0

a
0 0

GO

'

500

1000

2000

Rotation Speed, rpm

0 0

Fig. 3 1. Effect of current density and rotation speed on coating morphology at 52C determined by a rotating cylinder cathode [21].
31

100

80

rf-"

5
.-

60

$
2 .U

40

20

Dendritic

Micronodular

Polycrystalline

Coating Morphology

0 0 0 0 0 0 0 0 0 0 0
0

Fig. 32. Correlation between plating efficiency and coating morphology determined at four temperatures and current densities above 600 A/m2 [21]. Under a given set of hydrodynamic conditions, i.e. rotation speed and temperature, the current density above which the morphology changes from compact crystalline to micro nodular is referred to as the critical current density, ic. As seen in Fig. 33, ic was more than doubled, either as the rotation speed increased from 500 to 2000 rpm at a given temperature, or as the temperature increased from 27 to 60 "C at a given rotation speed. It is commonly believed that dendrites start to form once the current density reaches a certainfraction of the difflusion limited current density, iL,and a value of 0.4 has been cited for tin in a phenolsulphonic acid bath [23].
5000
I
'

I .

4000

0 0 0 0 0 0 0 0 0 0

"E 3000

2 c, .-

2000

1000

1000

2000

3000

Rotation Speed, rpm

Fig. 33. Effect of rotation speed and temperature on the critical current density, ic [21].

The preceding discussion makes it clear that a knowledge of the diffusion limited current density is paramount for any understanding of the morphology detennining

32

0 0 0 0 0 0 0 0

0
0

0
0 0

processeses. Yau [211 therefore attempted to calculate this current using the theory for the rotating cylinder electrode. The following equation was used:i,
= 0.079

nF d -0.3 V 0.7

-0.344

D 0.644

c b

0
0 0 0

0 0 0

where d is the diameter of the cylinder, V is the periferal speed of the cylinder, v is the kinematic viscosity, D is the diffusion coefficient of stannous ions and c b their bulk concentration. The effects of speed (V) and concentration (Cb) on i, are explicitly shown in equation (l), whereas the effect of temperature is not so obvious. This appears through an effect of temperature on the kinematic viscosity and the diffusion coefficient. Therefore, in order to examine the effect of temperature and calculate i, using equation (l), the kinematic viscosity (v) of the electrolyte and the diffusion coefficient (D) of the stannous ion were determined as a function of temperature in the range 27-60C. The kinematic viscosity was measured using a glass capillary viscometer. As shown in Fig. 34, the kinematic viscosity of the electrolyte decreased from 0.98 to 0.5 1 (10-6 m2/s) as the temperature increased from 24 to 60 "C. This temperature dependence was found to be simolar to that of water and correlated well with the GuzmanAndrade equation [23] which was used to interpolate the kinetic viscosity within the temperature range.

a
0 0
0

0
0

a a
0 0

0
0 0
0

20

40

60

80

100

Temperature, C

Fig. 34. Effect of temperature on the kinematic viscosity of the plating bath [21].
To determine the diffusion coefficient of the stannous ion, potentiodynamic tests were

0
0
0

a
Q

conducted using the rotating disc technique. As shown by the cathodic polarisation curves obtained at 27 "C in Fig. 35, the cathodic current density was very low at potentials above -0.3 V. At -0.4 V, the current density reached a plateau of 5 to 7 Nm2, attributed to residual oxygen or stannic ions. However, once the potential reached -0.48V, the current density increased quickly as nucleation of tin crystals commenced. At about -0.65V7 the current density reached another plateau as a result of attaining the diffusion limited current density of stannous ions. This diffusion-limited current density increased as the rotation speed increased. At a more negative potential, depending on the rotation speed, the current density increased again as proton reduction started to become significant.

33

....

0.0
-0.2

8 -0.4
n

2
>
c .

-0.6

5 g
Y

.E! -0.3 . c
-1.0 -1.2
-1.4 0.01

0.1

10

100

1000

10000

Current Density, ~ / m ~

Fig. 35: Effect of rotation speed on the cathodic polarisation curves at 27 "C [21].

The diffusion limited current density obtained from the rotating disc can be related to the rotation speed o (rads) by the Levich equation:213 i L = 0 . 6 2 2 n F D o 112 v 116 C b

(2)

As shown in Fig. 36, iLwas linear against o1I2 the four temperatures tested and the for diffusion coefficient of the stannous ion could be obtained from the slope. As shown in Fig. 36, the diffusion coefficient increased from 0.83 to 1.58 x lO-' m2/s as the temperature increased from 27 to 60 "C.
2

-7
E
X

n 0 ?4

0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

0
0
0 20

40

60

80

100

Temperature, C

Fig. 36. Effect of temperature on the diffusion coefficient of the stannous ion in the MSA plating bath [21].

With the determination of the kinematic viscosity of the electrolyte and the diffusion coefficient of the stannous ions as functions of temperature, the diffusion limited current, iL,of stannous ions was calculated using equation (1). As depicted in Fig. 37, while iLincreased as the Reynolds number (Re = dV/v) increased with increasing temperature at each rotation speed, i, was always found to be less than icycontrary to

34

0 0 0 0 0 0 0 0 0

0 0

e
0 0

expectations, and the difference increased as the temperature increased at each rotation speed. Although it was possible that the Eisenberg correlation underestimated the limiting current, because the actual surface area after deposition was greater than the geometric area; it was also pointed out that iChL the MSA bath could be greater for than the 0.4 observed for a phenolsulphonic acid bath mentioned earlier [23].
5000

e
0

. .

4000

e
0
E
..:

2000
. .

a &

0
0

1000

a a a
0
0

Fig. 37. Comparison of the critical current density ic and the calculated diffusion limited current density iL[2 11.

a
0

e
e e
0 0

In view of the possibility that the accuracy of the rotating disc data was being affected by interference from the hydrogen evolution reaction, further investigations were undertaken to elucidate this problem in more detail. Clearly, the rate of this reaction depends on the nature of the substrate. Although the exchange current density for the hydrogen reaction is two orders of magnitude lower on tin than on iron, the vast majority of the deposition reaction takes place on a tin surface because the substrate is covered by tin within a very short period of time. In order to simulate this effect and study both the plating and hydrogen reactions, a copper rotating disc electrode was employed which had been flash coated with tin prior to the experiments. This was accomplished using a potentiodynamic scan from the open-circuit potential (around OV vs. SCE for copper) to -0.7V at 2000 rpm. A second potentiodynamic scan was then made in the test solution. As shown in Fig. 38, the open-circuit potential of the flash coated tin was -0.45V, i.e. much lower than the -0.02V found on copper in the plating bath (Fig. 35). In order to study the hydrogen reaction in isolation from the plating reaction, experiments were also performed in tin-free solutions.
Fig. 38 shows that the cathodic polarisation curves exhibited a distinct region where the current was limited by diffusion of stannous ions and this increased with increasing rotation speed. However, as the rotation speed increased, the plateau of limiting current became less well defined. This was attributed primarily to the increase in surface area, as the coating became thicker and dendrites started to form. In the tin-free solution, the polarisation curves were independent of the rotation speed, suggesting that the hydrogen reduction was proceeding under kinetic control and was not affected by mass transfer. It can be seen that the current density in the tin-free solution was two orders of magnitude lower than that from the plating bath until the potential reached below -0.9V. Therefore it was concluded that when the current
35

a
0 0 0

e
a

density was less than iL(or above -O.SV), the current from hydrogen evolution could be assumed to be negligible and hence the plating efficiency approached 100%. Conversely, when the current density exceeded i L , the plating efficiency would decrease significantly as the proton reduction became increasingly significant.
0.0
-0.2

0 1

0
G

-0.4

0 0 0 0

-1.2

1 j
0.01
0.1

I I

-1.4

10

100

1000

10000

Current Density, Un r 2

Fig. 38. Cathodic polarisation curves on the tin flash-coated RDE in the plating bath and the tin-free solution at 27C [211.

In light of the above information, further studies were undertaken to explore the relationship between the plating morphology and current density at different ratios with respect to iL,i.e. at various i /iLvalues. It can be seen from Fig. 38 that iL= 1000A/m2 at 500 rpm and 27 "C. Gavanostatically controlled experiments were conducted on the rotating disc electrode at various current ratios with respect to iL, but with a constant total charge of 1800 C1 m 2 . As shown in Fig. 39, as the current density increased from 600 to 1000 A/m2, the coating remained compact crystalline whilst the potential decreased from -0.65 to -0.75V. As the current density increased to 1200 A/m2, the coating changed to micro nodular and the potential dropped to -0.92V. As the current density hrther increased to 2000 A/m2 and beyond, the morphology changed again to dendritic, but the potential stayed at -0.95V. This phenomenon had been observed at other combinations of temperature, rotation speed and stannous ion concentration. It was concluded from these experiments that although the hydrodynamic conditions in the rotating cylinder cathode and rotating disc electrode studies were different, the same basic plating characteristics were demonstrated by both techniques.
The experiments reported above were repeated, but in the absence of the grain refiner in the electrolyte. As shown in Fig. 40, a thin layer of discrete nuclei were formed regardless of current density between 600 and 1400 A/m2. However, these nuclei did not coalesce, resulting in large crystals of tin. As the current density increased, the number of crystals increased but the size of the individual crystals decreased. It was also noticed that the potential remained at -0.45V regardless of the current density.

0 0 0 6 0 0 0 0 0 0 0 0 0 0 0 0
0 0 0 0 0

36

a
0. 0 0
2
0.0

a
.
'

z
U Y
.CI

-Oe2

.5 -0.4

P
Y

3 .

';3 -0.6

' +/\I/\
E
0
. ) c

-0.8

-1.0
0 1000
2000

3000

Current Density, A/m2

Fig. 39. Effect of current density on potential and coating morphology in the plating bath. (RDE, 27 "C, 500 rpm, 18200 C/m2,iL=1000 A/m2) [21].

0.0

-0.2

%
d

-0.4

P a .a G
U

c .

-0.6

'

-0.8

-1.0

500

1000

1500

2000

Current Density, N m 2

Fig. 40. Effect of current density on potential and coating morphology in the plating bath without additive. (RDE, 27 "C, 500 rpm, 18200 Urn2,iL=1000 A/m2) [21]. The work of Yau [21] described above is important because it attempts to correlate all the electrochemical and hydrodynamic processing variables during plating with the plating morphology. It also allows these parameters to be related to the technology of high speed plating. In particular, the effects of plating speed (agitation) and temperature on the diffusion limited current density can be simulated. This in turn is regarded as the key parameter in determining bath efficiency and the specific morphology observed at any particular applied current value.

37

4.3

Methanesulphonic Acid Electrolyte in the Presence of Polyethylene Glycols

Organic additives are commonly used in the electrodeposition of tin from acid solutions. Surfactants and grain refiners are added to plating baths to influence the deposit morphology, brightness and grain refinement. These addition agents are an important component of plating baths for the deposition of tin with desirable properties and morphology suitable for electronic and other practical applications. While the use of organic reagents in electrodeposition is commonplace, the specific activity of the reagents is only generally understood in terms of adsorption at the cathode surface during the deposition process. This fact was discussed previously in the current review (Section 3.1). The manner in which small amounts of organic species affect the morphology of the deposit remains unclear. Adsorption on the metal surface is believed to be the main generic process. Numerous experimental works have discussed the effect of additives on metal electrodeposition. However, a major limitation of these studies has been the inability to characterise accurately the adsorption of ionic and molecular species at the electrode surface during the electrodeposition. Very recently, Zavarine et al. [25] have studied the adsorption behaviour and the effect of additives on the tin electrodeposition process using electrochemical and surface enhanced Raman spectroscopy (SERS) techniques. The latter is a vibrational spectroscopy that provides 1O6 times stronger Raman signals of the species adsorbed on a metal surface as compared to normal Raman spectroscopy. In addition, SERS is highly surface selective, that is, only the molecules (or parts of them) that are in direct contact with the surface are enhanced and show up in the spectra. SERS has therefore been widely used to obtain information on what species are adsorbed, what their orientations are and how strong is the adsorption bond. These studies, however, have been mainly limited to relatively small and simple molecules, which are normally not used in electroplating. One exception is the work of Hope and Brown [26] who studied the adsorption of poly(ethy1ene-glycol) at a polarised copper electrode and found adsorption to be potential dependent. Unfortunately, direct measurement of the adsorption behaviour of organic molecules on a tin electrode by SERS is very difficult because the enhancement effect is limited to gold, silver and copper electrodes. Nevertheless, Zavarine et al. [25] argued that valuable insights could be gained by correlating the adsorption behaviour measured by SERS on gold or silver electrodes with the electrochemistry of tin and the morphology of tin deposits. As a model system, a combination of Triton X-100 derivatives (polyethylene glycols with or without hydrophobic head groups) and phenolphthalein were investigated. The latter was chosen because it is a commonly used grain refiner. Such systems are known to produce smooth and reflective coatings of both pure tin and tidlead [27]. Schematic structures of the Triton X-100 family and phenolphthalein are depicted in Fig. 41. In the case of Triton X itself, n = 8 and the number of ethylene oxide groups in the tail of the molecule is 10. Various head groups of the form C,-Ph- were studied (where n = 8 and 18) and various ethylene oxide tail lengths (range 5-150 ethylene oxide groups). The electrolyte consisted of 1M methane sulphonic acid and 0.5 M tin methanesulphonate. The working electrode in the electrochemical studies was a rotating copper electrode and in the SERS investigations gold or silver electrodes.
38

0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
C

0 0 0 0 0

e
0 0 0 0 0 0 0 0 0

0
0 0 0
0
bead
tail

a
0
0

IgepalS20
on

0 0
0
phenolphthalein

0 0
0 0
0

Fig. 41. Schematic structures of the Triton-X family of surfactants and phenolphthalein [25] Typical voltammograms for the electrodeposition of tin with and without Triton X-100 are shown in Fig. 42. The onset of reduction current occurred at ca. -0.42V. Without Triton X-100, a current plateau was seen in the curve and as the potential decreased, the current density started to increase rapidly again. Such behavior is usually associated with dendritic crystal growth, which greatly increases the surface area of the electrode. The addition of Triton X-100 to the solution suppressed the dendrite growth of the tin deposit and the cyclic voltamrnogram revealed two distinct peaks. The current in the first peak at - 0 . W did not depend on tin the concentration in the solution and was assumed to be associated with adsorption of the additive on the growing tin layer, which strongly inhibited the tin reduction current. This process could be reversible, as seen from the inset in Fig. 42. The second peak at -0.75V depended strongly on the tin concentration indicating that it was a diffbsion limited current. Curves of a similar qualitative shape were obtained for other polymers in the series and also a block copolymer of ethylene oxide and propylene oxide containing 85 ethylene oxide groups in the tail of the molecule. Fig. 43 provides some indication of the range of surface morphologies obtained including that in the presence of phenolphthalein.

<

a
0 0

a
0
0 0 0
0

0 5M SnMSA+ 1M MSA

-10

2 5 g/l Triton X-1 00

E, -20
+ .-

5 I n

-30
U
+
C

0
0 0 0 0 0
.

!?
0
.

-40
no additives

-50
-1 2

-1 0

-0 8

-0 6

-0 4

Potential, V vs AgIAgCI

Fig. 42. Voltammograms for reduction of tin with and without the addition of Triton X-100 ~51.
39

0 0 0 6 0 0 0 0

0 0 0
0 0

Fig. 43. Surface morphology of tin deposit obtained with solutions of (a) Triton X-100, (b) block copolymer of ethylene oxide and propylene oxide and (c) Triton X-100 and phenolphthalein [25].

0 0 0
0 0 0

40

0
0

0 0
0

The SERS spectra of the Triton X class of surfactants could be exemplified by the spectra of Igepal CA520 where the number of ethylene oxide groups in the tail ofthe molecule is 5. Fig. 44 shows the SERS spectra of this compound obtained for a saturated solution in 0.1 M sulphuric acid at the open-circuit potential of a gold electrode (+0.23V) and at -0.W vs. sat. calomel from which two different surface orientations of the molecule were deduced. These are depicted in Fig. 44.

0 0 0
0

d63+?

+0.2

0
0

0 0 0 0

Fig. 44. Suggested orientation of Igepal 520 on a gold electrode surface at (a) +0.23V and (b) -0.5V VS. SCE [25]. In conclusion, Zavarine et al. claimed that their work had demonstrated that new insights had been obtained regarding the adsorption of organic additives used in electroplating. Triton X- 100 is adsorbed on the electrode surface in a wide potential range. The hydrophobic head group serves as an anchor to the metal surface. At negative potentials, around -0.5V, the molecule turns its hydrophobic tail to the surface increasing the effective surface coverage. The head groups of Triton X compounds are not required for adsorption but significantly enhance it. Polyethylene glycols are also adsorbed to the metal surface, but the effective coverage is relatively low. This results in a rough deposit. Triton X-100 facilitates adsorption of phenolphthalein and interacts weakly with it through some kind of host-guest interaction. The coadsorption of Triton X and phenolphthalein modifies the grain size and the deposit structure to impart a unform and smooth surface as compared to Triton X alone.

0
0 0

0 0 0 0 0
0

4.4

Phenolsulphonic Acid Electrolyte in the Presence of Ethoxvlated a-Naphthosulphonic Acid

0
0
0

Various organic compounds have been used in tin electroplating baths for the purpose of enhancing throwing power, grain refinement, surface brightness and anti-oxidative properties of stannous solutions. Also, extensive efforts have been directed to find more effective organic compounds for tin electroplating and to examine the structural properties of the deposits [27-331. Although the initial deposition process is critical in determining the overall quality of the deposited layers, little attention has been paid to establishing the way organic molecules act on the kinetics of the initial tin electrodeposition process at the iron electrode, which is one of the most important industri a1 materials.

41

0
Very recently, Lee et al. [341 have investigated the effects of an organic compound, ethoxylated 3-(a-napthol) sulphonic acid, also known with the commercial name ENSA-6, on the polarisation behaviour, kinetic parameters of the electrode reaction, interfacial properties and surface morphology when tin is electrochemically deposited at pure iron electrodes in acid stannous sulphate solutions. The plating solution consisted of 0.14 M SnS04,0.033 M PSA as supporting electrolyte and 0.013 M ENSA as an additive. An iron rotating disc was used in the work. The electrode was polished successively with alumina slurries and then cleaned ultrasonically with distilled water before the experiments. Potentiodynamic polarisation, electrochemical quartz crystal microbalance, scanning probe microscopy and electrochemical impedance spectroscopy techniques were employed to study the system. The structure of an ENSA molecule in an extended and coiled configuration is shown in Fig. 45.

0
0

0 0
0

0
r

0
0

a
0 0 0 0
0

0 0 0

0
0 0
Fig. 45. Structure of ENSA in (a) extended and (b) coiled configuration [34].

0
Fig. 46 shows a series ofpotentiodynamic curves recorded for tin deposition at the iron disc electrode rotating at 1200 rpm in solutions containing various concentrations of ENSA while the potential was swept at 2mV/s from the open circuit potential (-0.4OV to -0.45V). Shown inset is an enlarged version of the circled region. In general terms, the higher the ENSA concentration, the more distinctive the current peaks and the limiting currents became. Limiting currents for tin deposition have been reported to be smaller when the surfactant concentration approached the critical micelle concentration (CMC) which was interpreted to control the mass transfer of Sn2+[34]. It was noted that the rate of nuclei formation at the surface, which appeared to be related to the peak current at -0.54V7 seemed to be the lowest at 1x M ENSA. This suggested that ENSA molecules.contro1 the mass transfer of tin for nuclei formation by keeping Sn2+off the electrode surface by forming sheaths that
42

0 0 0
0 0 0 0

0
0

0
0 0 0 0
hold these ions, thus presenting an effective barrier to Sn2' reduction. A rather high current was observed at the high ENSA concentration of 0.13 M. The facilitation of' mass transport of tin in this case was postulated to result from a transformation of the Sn2'-ENSA aggregates on the electrode surface in which pillar-shaped complexes are formed in an ordered manner. This configuration may direct the Sn2' towards the surface and facilitate its mass transport, resulting in a higher peak current. The rather ill-defined current peaks at lower ENSA concentrations appeared to be due to insufficient control of the mass transfer of Sn2'.

0 0
0

0
0
N

Y Y

a
0

E .
4

60.0m

..........................................
I I

5
L

40.0m

5).

0
20.0m

e
0

0.0 1

......................

........ ....... 1.3~ 10-2MENSA

backsround

-0.4

-0.6

-0.8

-1 .o

-1.2

E vs. Ag/AgCI(V)

e
0

e
0

Fig. 46. Linear sweep voltarnmograms at a rotating iron electrode (1200 rpm) conducted at 2 mV/s in solutions containing 0.14 M Sn2', 0.033 M PSA and (a) 0 M, (b) 1.3 x 104 M, (c) 1.3 x 10-3 (d) 1.3 x 10-2 M and (d) 1.3 x 10-' M ENSA [34]. M, Fig. 47 displays open-circuit potential data measured at the iron electrode when the concentration of ENSA was varied in a 0.033 M PSA solution with and without 0.14 M Sn2'. A change of 35 mV was noted per 1000-fold change in the ENSA concentration. However, when only the [Sn2'] was varied in a control experiment without NSA and PSA present, a change of 3 1 mV/decade in the OCP was observed (not shown), in excellent agreement with the Nemst equation (n =2 ). Thus it was concluded from Fig. 46 that an increase in [ENSA] by three orders of magnitude causes the effective concentration of the potential determining species, i.e., Sn2', to decrease to only about 1/10 of the original concentration at the electrode surface, resulting in a change in the electrode potential by only 35 mV. This reasoning confinned that some form of complexation between ENSA and Sn2' was in operation. The trends displayed above could also be discerned from galvanostatic (chronopotentiometric) measurements. Fig. 48 shows galvanostatic polarisation curves recorded at 20 mNcm2 at the rotating Fe-RDE in acidic plating solutions containing variable concentrations of ENSA. These results were interpreted in terms of an instantaneous deposition of tin when no additives are present changing to a progressive mechanism in the presence of ENSA whilst the potential was maintained at a value of about -0.75V. The nucleation mechanism has been reported to affect the quality of the deposits; electrodeposits obtained via progressive nucleation having better quality than those by instantaneous nucleation [35-371. Bulk reduction of Sn2' then takes place at higher overpotentials. Thus, Zarger nuclei are formed when no or a low concentration of ENSA is present, which leads to the poor quality of the tin bulk
43

e a
0
0 0 0

8
0
0

0
0

8
0
~

deposition. A controlled number of smailer sized nuclei are fonned on the surface, which controls the rate and quality of bulk deposition at higher polarisation. This observation was consistent with that observed during the potentiodynamic experiments (Fig. 45), in which there was a small peak at around -0.55V followed by a constant potential to about -0.78V.
'

-0.40 7

,.

.
-0.41

4... i :

-0.42

vj

-0.43

a'
0

-.4 04

1 ' I ' I ' I . I . I

0
-0.45

-2.5

-2.0

-1.5

-1.0

-0.5

0.0

0.5

1.0

l 1.5

log([ENSA] / (1.30 x 10'* M ) )

Fig. 47. Open-circuit potentials measured at a rotating iron electrode as a function of the concentration of ENSA in a 0.033 M PSA solution. Open symbols (with) and closed symbols (without) 0.14 M Sn2+[34].
(PSA+ENSA) backgound
'

-1.0-0.9

PSA solution

0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
0 0 0 0 0 0

3
2
2
>

-08-

-0.7-0.6-

-05-04l ' l ' l ~ ~ ' l ~ l ~ l ' l ~ i ' l ' l ' l

-2

10

12

14

16

18

20

22

time, s

Fig. 48. Chronopotentiometric curves for tin deposition at a constant current density of 20 mA/cm-2at a rotating iron electrode (1200 rpm) in solutions containing 0.14 M Sn2+, M, 0.033 M PSA and (a) 0 M, (b) 1.3 x 10-4M, (c) 1.3 x 10-3 (d) 1.3 x 10-*M and (e) 1.3 x 10-' M ENSA [34]. To investigate the changes caused by the potential transition, atomic force microscopy (AFM) was used to obtain images of the electrode surface before and after the current step in the tin plating solution containing 1.3 x 1OP2 M ENSA.The data is shown in Fig. 49. Both the images A1 and B1 were taken ex situ in the air from the surface after the galvanostatic deposition for 0.5s and 2s, respectively, at a stationary electrode. The AFM images were clearly different in that only a few grains of about 240 nm were found in image A l , while the whole surface was covered with large grains in image B 1. Thus, ENSA molecules affect the nucleation process around the peak current region by adsorption at the iron electrode, leading to the formation of tin grains across the whole surface of the electrode after the potential transition.
44

0 0 0
0

0 0 0 0

.:.- . . .. ..

. .. .

"

a
0

0 0 -

i
I

0 0
0
-0.50
-0.45 -0.40

* I
I

0
0

0.0

0.5

1.o

1.5

2.0

Time, s

a
0

Fig. 49. AFM images taken during galvanostatic polarisation electrolysis in a stock solution containing 0.0 13 M ENSA at a stationary Fe electrode (roughness 1.7 nm). AFM image taken after (Al) 0.5 and (Bl) 2s of electrolysis at 20 mA/cm2 [34].

a
0
0

a
0

AFM images were also taken after tin was deposited by stepping the potential to -0.52V as shown in Fig. 45 with and without ENSA and it was shown that the root mean square roughness (R-) was smallest when [ENSA] was 0.013 M at 7.5 nm. The Rmsvalues ranged from - 18 ntn when ENSA was 0.13 M to as large as several tens of nanometers in the absence of ENSA. This indicated that ENSA inhibits the instantaneous nucleation and the effect is maximal when [ENSA] is close to its CMC.
Resulting from the above extensive studies and other experimental investigations, the following conclusions were drawn from the work: -

0 0
0

0
0

1. ENSA molecules enhance tin plating quality by affecting both the nucleation and bulk deposition processes. The nucleation step is influenced by the adsorption of aggregated ENSA molecules at the surface. Also, the presence of ENSA in the interlayers of the tin deposit increases the overpotential for bulk deposition of the tin.
2. A critical concentration of ENSA may be defined for its optimum effectiveness as an additive. At this concentration and above, ENSA molecules appear to interact intimately with neighbouring ones, leading to the fonnation of compact aggregates at the surface requiring a large overpotential for the reduction of Sn2+ due to complexation with ENSA.

0 0 0 0
0 0

a
0

3. The compact structure is evidenced by smaller double layer capacitance values of the tin layer-electrolyte interface. While the aggregates formed without Sn2+ and were separated from each other, those in the presence of Sn2+form a larger and thicker blanket. This appears to control the access of Sn2+to the electrode surface.

e
0
45

4.5

Elimination of Tin Whiskers by Reducing Stress Levels in Electrodeposits

Zhang and Abys [39] have described a plating process that is claimed to be a unique electroplating tin chemistry. Smooth, satin bright tin deposits that have stable, large grain structures in the region of 5 pm are produced. The chemistry is capable of operating at elevated temperatures over a wide range of current densities and is thus applicable to rack, barrel and reel-reel operations. Extensive bath life studies have shown that the deposit appearance and material properties, including grain structures, are stable in relation to the age of the electroplating chemistry. In addition, the grain refiners used are highly stable and have few breakdown products as the chemistry ages. This tin electroplating process has been utilised in plating coatings for connectors, solder bumps, PWPs and components for semiconductor applications. In particular, it is claimed that the process minimises the formation of tin whiskers. After careful examination of both literature and commercially available processes, it became clear that an electroplating tin chemistry that produces coatings with zero or low internal stress and low organic inclusions was highly desirable. It was claimed that a typical satin bright plating bath formulation and the operating conditions required to fulfil these objectives would contain 80 g/1 tin as metal fiom stannous methanesulphonate; 200 ml/l MSA, 15 ml/l additive A (proprietary wetting agent); 10 mV1 additive B (proprietary grain refiner); cathode current density 100 ASF; anodekathode ratio 1:1; temperature 55C; and agitation 50 c d s . The satin bright tin deposits showed superior behaviour in the context of whisker reduction and reflow operations. Deposits had excellent ductility, solderability and grain structure. These properties remained constant after extensive aging of the electroplating chemistry. Whiskers were absent fi-om samples aged under ambient conditions and long term accelerated aging under high humidity and high temperature conditions. Tin coatings plated on copper also showed they were easily reflowed without any solderability failures. These properties were attributed to the large well polygonised grains and the very low organic inclusions. These were both believed to be instrumental in reducing the internal stress that is the root of whisker formation.

4.6

Production of Tin Methanesulphonate

The high rate electroplating of tin on a moving steel strip is generally carried out in cells with dimensionally stable anodes. To obtain a matt tin deposit, a concentrated acidic tin methanesulphonate solution containing a small concentration of sulphuric acid is used. The concentrated tin methanesulphonate solution is prepared by dissolution of tin particles in the presence of oxygen in a special column. In order to gain a better insight into the mechanism of the reactions pertaining, Greef and Janssen [40] have studied the electrochemical dissolution processes. These included the reduction of oxygen, the reduction of hydrogen peroxide and the reduction of hydrogen. It was concluded that on tin, oxygen is almost entirely reduced to water and that hydrogen peroxide cannot corrode tin directly, but its decomposition products, e.g. oxygen, can. The exchange current density and the charge transfer coefficient for the investigated reactions were estimated. It was concluded that the dissolution (corrosion) of tin is determined by the kinetic parameters of the oxygen reduction reaction and by the mass transfer of dissolved oxygen towards the electrode and the mass transfer of tin

46

0 0 0 0 0 0 0 0 0 0 0 0 6 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

0
0

away from the electrode surface. Hydrogen evolution can be neglected during the dissolution of tin in the oxygen saturated solutions employed.

0 0 0
5.
Electrolytes and Additives in High Speed Plating High speed electroplating equipment and processes are well known in the industry and generally consist of directing the work to be plated into the electroplating cell from one end, allowing the work to proceed through the electroplating cell and exit thereafter the cell at the other end. The electroplating solution is removed or overflows the electroplating cell into a resevoir and the solution is pumped from the resevoir back into the electroplating cell to provide vigorous agitation and solution circulation.and exits the cell thereafter the cell at the other end. Many variations of these electroplating cells can exist, but the general features are as described. There are a number of features that the electroplating solution should possess for improved operation in this type of equipment or processing :-

0
0
0

0
0

a
0
0

a
0 0 0
0

1. The solution must be able to electroplate the desired deposit at the high speeds required and over an extended range of current densities but with minimal amounts of foaming. 2. The plating bath should have a high conductivity, a low corrosivity and high antoxidant properties, but with a low environmental impact. 3. The deposit should be lustrous and fine grained, even at the high current densities required for high speed plating. 4. The deposit should have good solderablity and be capable of meeting the solderability requirements specified for such deposits. 5. The solution should be stable and the additives must withstand exposure to the strong solution as well as to the introduction of air, which wouldtake place as a result of the vigorous solution movement in high speed plating machines. 6. The solution should remain clear and fiee of turbidity, even at elevated temperatures such as 120-130F or higher.
Various plating bath compositions comprising an alkane or alkanol sulphonic acid (normally methanesulphonic acid), a tin salt and various auxiliary additives are known. These range from small organic molecules to large polymeric surfactant molecules. Plating bath compositions containing mixtures of aryl and alkyl sulphonic acids are also known. Various addition agents have been proposed which enhance the quality of the tin plate. They can include condensates of hydrophobic organic compounds with alkylene oxides such as, for example, alpha naphthol 6 mol ethoxylate (ENSA 6); alkylbenzene alkoxylates such as the Tritons; derivatives of N-heterocycles such as 2-alkylimidazolines; aromatic aldehydes such as naphthaldehyde; and 2,4,6substituted phenols in which at least one of the. substituents includes a secondary, tertiary or quaternary nitrogen atom. The following is a summary of recently proposed plating bath compositions which incorporate sulphonic acids.

e
0 0

0 0
0 0
0

0
0

0
0
47

a
0

0
0
5.1
Sulphonic Acid + Alkylene Oxide Condensation Compound

Toben et al. [41] have described an electrolyte system and process for depositing tin (or tidlead alloys) upon a substrate by high speed electroplating, which includes a base solution of an alkyl or alkylol sulphonic acid; at least one of a solution soluble tin compound; together with an alkylene oxide condensation compound of either (1) an aliphatic hydrocarbon having less than 8 carbon atoms and at least one hydroxy group; or (2) an aromatic organic compound having no more than twenty carbon atoms in one or two independent or joined rings optionally substituted with an alkyl moeity of eight carbon atoms or less. A preferred hydrocarbon is an alcohol, such as butyl alcohol. The alkylene oxide compound may be ethylene oxide, wherein preferably between 6 and 28 moles of ethylene oxide are used to form the condensation compound. The aromatic organic compound can be benzene, naphthalene, phenol, toluene, bisphenol A, styrenated phenol, or an alkylated derivative. 5.2 Sulphonic Acid + Stannous Alkyl Sulphonate + Reaction Product of Polyalkylene Glycol and Phenolphthalein or Derivatives of Phenolphthalein

0
0

a
0
0

0
0 0 0

0
0

Hong and Yung [42] have described a plating bath for plating tin (or tin alloy) onto metal substrates in a high speed plating process. The electrolyte bath contains an alkyl sulphonic acid and a solution soluble tin compound. The bath also contains an organic compound that is the reaction product of polyalkylene glycol and phenolphthalein or derivatives of phenolphthalein.

0
0
0 0

An example of a suitable alkyl sulphonic acid is methane sulphonic acid and an example of a solution soluble tin compound is tin methanesulphonate. The concentration of the alkyl sulphonic acid in the bath is selected to provide a bath pH of about zero to about three. The solution also contains at least one organic additive that is the reaction product of a polyalkylene glycol and phenolphthalein and derivatives of phenolphthalein. In the reaction product, the number of alkylene oxide moieties (e.g. -C,H,O- where y = 2x) is about 10 to about 62. The alkylene oxide moieties can be substituted or unsubstituted. In the embodiment of the invention wherein the polyalkylene glycol is ethylene glycol, the alkylene moieties are ethylene oxide moieties.
The condensation product is formed by reacting either a polyalkylene glycol with the phenolphthalein derivative. In one embodiment, the condensation product is formed before it is added to the electrolytic plating bath. In a second embodiment, the condensation product is formed in the electrolytic plating bath itself. The solution also contains an antioxidant. Examples of suitable antioxidants include a sodium salt of benzaldehyde sulphonic acid or derivatives thereof. The background research relating to this invention has been discussed in Section 4.3.

0 0

0
0 0

a
0
0
0

0 0
0

0
48

0 0
0

0 0

5.3

Sulphonic Acid + Substituted Phenol

0
0

0
0

0
0 0

In the case of strip tinplate manufacture, it is desirable from a commercial point of view to have a system capable of giving satisfactory tin deposits over as wide as possible a range of current densities in order to accommodate all variations in speed of production and to minimise the incidence of current density defects. It is generally accepted that it is the addition agent that has the greatest effect on tin plate quality and little work has been done on how acids affect plate quality such as plating width and brightness. Recently, however, ODriscoll [43] has claimed that certain acids have the capability of improving the performance of addition agents at high current densities whilst others are able to improve performance at low current densities. When combined, synergistic effects are produced which give even wider plating ranges and bright plates. This synergistic effect is hrther enhanced when the acid combination is used in conjunction with specific additives.
Accordingly, the above invention provides a combination suitable for use in a process for electroplating surfaces with tin, which exhibits all the aforementined benefits in a wide plating range, has good quality deposits and enhanced environmental benefits. It comprises:(a) (b) (c) (d) (e) (f)

0 0 0

0
0 0
0 0 0

An unsubstituted or substituted alkyl benzenesulphonic acid One or both of sulphuric and sulphamic acid One or more addition agents A tin source An antioxidant (optional) Water

0 0
0
0

The preferred para alkyl benzene sulphonic acid is para toluene sulphonic acid. The addition agents are chosen so as to enhance the synergistic action of the mixtures of acids (a) and (b). The preferred additives are mono, di- and tri-substituted phenols (each optionally alkylated or alkoxylated) having at least one substituent containing a secondary, tertiary or quaternary nitrogen atom, or mixtures of two or more such components. The preferred phenols are 2,4- or 2,6- di substituted or 2,4,6- tri substituted phenols. Typical antioxidants include, 1,2-dihydroxy benzene; 1,2,32 trihydroxy b enzene; 1, -dih ydroxyb enzene-4-sulphonic acid; etc . 5.4 Alkali Metal, Alkaline Earth Metal, Ammonium or Substituted Ammonium Salts of Alkyl or Alkanol Sulphonic Acid

0
0
0

Gillman et al. [44] have described the use of alkali metal, alkaline earth metal, ammonium and substituted ammonium salts of alkyl and alkanol sulphonic acids as additives in pure metal (and metal alloy) sulphate electroplating baths for improved appearance, oxidative stability and wider current range. It is claimed that in addition to tin plating, other metal plating processes can be improved, including Cu, Cr, Cd, Fe, Rh, Fe-Zn and Sn-Zn plating.

49

0 0
5.5

Sulphonic Acid + Carboxyalkvlated Polvalkyleneimine Compounds

0 0
0

Crosby [44] has recently dicussed the use of carboxylated polyalkyleneimine compounds as additives in tin plating processes. These are polyalkyleneimine compounds contain one or more carboxyalkyl substituents. The additives are claimed to be particularly effective for increasing the current range for the production of quality uniform deposits from low metal concentrations.
5.6

0
0 0
0

Toluene Sulphonic Acid + Ammonium Salts and/or Magnesium Salts

Crotty [46] has proposed a composition and process for electroplating tin (and tin alloys) onto a substrate at relatively high current densities. The electrolyte comprises toluene sulphonic acid and a source of ammonium ions and/or magnesium ions. Previous attempts to use toluene sulphonic acid have encountered difficulties due to the low solubility of the compound and the relatively low conductivity of the solution. However, the inventor has discovered that by combining, in aqueous solution, this acid with ammonium salts and /or magnesium salts, the solubility of the toluene sulphonic acid can be increased substantially. It is claimed that excellent plating results can be achieved.

0
0
0

0 0 0

6.

Electroplating and Electrotinning Quality

0
0

6.1

Testing of Tin Plating Using a Rotating Cone Electrode


-

In the process of etching surplus copper from printed circuit boards, electrochemically deposited tin is used as a protective layer. For this purpose, a bright or semi-bright, ductile, non-porous tin layer, prevents copper etching or partial copper etching in the hot etching ammoniacal copper(I1) chloride solution. After the etching of surplus copper from a circuit board, tin is stripped away in hydrochloric acid solution with addition of copper(I1) chloride and tin(I1) chloride.
As a protection layer, electrochemically deposited bright tin is obtained from a tin sulphate electrolyte in the presence of organic additives. Also, additives that include a brightening agent and a leveling agent are included. The influence of the additives on the properties of the tin layer related to brightness, porosity and thickness distribution is considerable. During the operating time of the electrolyte, the additives are partly decomposed and partly incorporated into the deposit and must be renewed. However, the rate of this process requires quantification and therefore constant monitoring. Recently, Reichenbach [46] has described the use of a rotating cone electrode specifically designed for this purpose. In these studies the electrolyte was composed of tin sulphate, sulphuric acid and three different additives: a brightening agent; a leveling agent (described as IRA); and a surface tension agent (described as AN1 1). The electrode is shown schematically in Fig. 50. The anode-cathode distance is 2.5 cm.

0
0

e
0

0 0 0
0

0
0 0
0

0
0
50

Cathode
I

1i
I

Anode

Fig. 50. Rotating cone electrode [46].

Testing of current density distribution on the rotating electrode was carried out with total current variation from 0.5-2A at a rotation rate of 260 rpm in a copper and tin containing solution. After plating, the copper deposit and tin layer were carefully peeled off the cone. The thickness of the tin deposit along the surface of the rotating cone was measured by p backscatter. The current density distribution was calculated from the copper thickness data, assuming 100 % efficiency. Fig. 5 1 shows the dependence of current density distribution on distance from the high current density (cone apex) end of the electrode. (This distribution was not found to be a function of electrode rotation speed in the range 100-800 rpm at similar applied current values). It can be seen from Fig. 5 1 that a considerable range of current densities is produced along the surface of the electrode when the applied total current is 2A. This value was therefore chosen to test the effects of temperature and additive compositions or concentrations. A rotation rate of 260 rpm was also used.

4.0 -

Total current, A
0

0.5
1.5 2.0 .
260 rpm

3.5 3.0 -

0 1.0

*
0

25 .
2.0 -

1.5 1.0 -

0.5 I I I I I I I I I I

Fig. 5 1.Dependence of current density distribution on distance from the high current density . end of the rotating cone electrode at various total currents [46].

51

0
Fig. 52 shows that the appearance of the tin coating depended upon additive concentration and current distribution. With increased brightener concentration (0.5-3 ml/L), bright and very bright tin in the high current density region was obtained. The varied concentrations of additives, IEZA (1-3 ml/L) and AN 11 (5-30 ml/L) affected the decreased bright appearance of tin in the high current region. With overdoses of brightener, the bright area of tin in the high current density (apex) region was predominant. High concentrations of brightener (2-2 ml/L) shifted the bright area to the low current density region. In addition to these experiments it was also shown that the effect of temperature in the range 20-26 "C was to decrease the brightness in the high current density region.

0 0 0
0 0

3ri;Fer

1
I

Additive Additive AN 11 mUL

Rotating cone electrode

Appearance

dull bright dull bright dull bright

5.0

5.0

lmO
10.0

1.5

2.0

dull bright dull

--tI

20.0

bright very bright dull very bright

0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
0 0 0 0

30.0

I
3.0 .3.0

30.0

dull bright very bright

Fig. 52. Appearance of tin deposit on the rotating cone electrode at various levels of additives ~461.

In conclusion, it was claimed that the rotating cone electrode proved to be a very good tool for testing the appearance of tin coatings [46].

52

0 0

0
0

6.2

Surface Morphology and Appearance of Electrodeposited Tin Films

a
0 0 0

a
0

0 0 0
0
0

As discussed above, the visual appearance of electrodeposited tin coatings is an important characteristic. This has often been classified as matte, satin-bright and bright, among other descriptions. These qualitative classifications are obviously subject to interpretation and often lead to misunderstanding. A quantitative assessment based on some physical measurement would be beneficial for quality control and would provide the basis for understanding the influence of a plating process and operating parameters on the visual appearance. Furthermore, this analysis would be extremely useful if it could be correlated to other parameters, e.g. wear resistance, solderability, etc. These questions have recently been addressed in some detail by Xu et al. [47]. Because of the importance of this field in the context of the current review, the results from this complex investigation are discussed in some detail below. The appearance of an object can be described by two factors, viz. colour and geometric attributes. Colour arises from the wavelength-dependent absorption and reflection of light. The geometric attribute is often determined by gloss, lightness, haze, luster, surface uniformity and directionality. A goniometric curve, produced by measuring reflected light at different angles of observation for a fixed incident angle of light, is required to fully describe these geometric attributes. Fig. 53 is a schematic representation of typical goniometric curves for matte and bright surfaces. A bright surface reflects a larger proportion of light specularly relative to a matte surface, where the spacial distribution of reflected light is broader. The specularly reflected light contributes to the glosiness of the substrate, whilst the diffuse reflection leads to reflection haze. A detailed measurement of a goniometric curve is tedious, however, requiring expensive instrumentation and is therefore not very practical. An alternative Method to describe the gloss of a surface is the gloss reflectance factor [48], which measures the ratio of specularly reflected light to light reflected from a gloss standard.

a
0 0 0 0

a
0 0 0

a a
0
0
Fig. 53. Goniometric curves of matte and bright surfaces [47]. The shape of a goniometric curve is dependent on the type of surface under investigation and the incident angle of the light used for this measurement. This is illustrated in Fig. 54, where the measured gloss is plotted as a function of the surface appearance or visual gloss at three different incident angles.
53

a
0 0

0 0

0 0 0 0

Surface appearance-visual

gloss

(3

Fig. 54. Measured gloss vs. visual gloss ('appearance') at various light incident angles [47].

It can be seen that there is a very good response between the measured gloss and the visual appearance for a highly 'reflective' surface at 20" and for a low 'reflective' surface at 85". Conversely, the measured gloss is rather insensitive for highly 'reflective' surfaces when measured at 85" and for low 'reflective' surfaces when measured at 20". To quantitatively describe the appearance of surfaces with high gloss, therefore, a small light incident angle should be used, e.g. 20, as under these conditions a small change in the appearance of the surface will yield a large change in the measured gloss. On the other hand, a high light incident angle (85") is required to sensitively distinguish between surfaces with low gloss. Typical light incident angles are 20,60 and 85" and specific angles have also been adopted for special uses, such as 45" for ceramic and 75" for paper. The parameters affecting surface glossiness have not been investigated in detail. However, it is generally thought that surface morphology, surface roughness, grain size, grain structure and orientation significantly affect surface glossiness. Xu et al. [47] set out to establish the relative importance of these factors in determining surface appearance. Specifically, the influence surface roughness and grain structure was investigated in detail. Electroplated tin films classified as bright, satin-bright and matte were plated to a thickness of 3.0 pm using proprietary chemistry [49,50]. These films were deposited on a 2.5 pm thick nickel underlayer plated from a proprietary semi-bright nickel process [5 11. Gloss measurements were made using a variety of spectrophotometers, gloss reflectometers and haze gloss meters. Preliminary experiments showed that a measurement geometry with a light incident angle of 60" provided one of the best (most sensitive) test conditions for differentiating bright, satin-bright and matte electrodeposits. The initial focus of the work was concerned with the correlation between glossiness and surface roughness. The surface roughness of the plated deposits is shown in Fig. 55, where typical 500 pm line span profiles for bright, satin-bright and matte tin are illustrated. These measurements were made using a surface profiler and the arithmetic roughness was calculated by averaging the deviation from the mean line. For comparison, the line profiles for brass and Nihrass substrates are also included in Fig. 55. As can be seen, the bright tin plating slightly reduced the roughness of the

0 0 0

e,

0 0 0 0

0 0 0 0 0
0 0 0 0 0 0 0 0 0

54

0
0

0 0
0 0
Nihrass substrate, exhibiting a leveling effect, whilst the satin-bright and matte bright dramatically increased the roughness of the surface.

0
0

a> C .c
0)

VI VI

Bright Sn

2
0

a
0
0
0

hlatte Sn
v)

100

200

300

400

500

600

Across the surface, pm

Fig. 55. Surface roughness of various substrates [47]. A strong correlation was demonstrated between glossiness and surface roughness with the smooth'surface being glossier than the rough surface as expected. This is clearly shown in Fig. 56 where the plot of gloss reflectance factor vs. roughness is exponentially related. It is interesting to note that satin-bright was relatively close to matte in terms of surface roughness and glossiness, whereas a rather large difference existed between bright and satin bright. The authors commented that it would have been instructive to have plated tin with a surface roughness between 200 and 600 A which should have exhibited a gloss factor between 200 and 700. Moreover, it would also have been usehl to have plated deposits with a surface roughness >1200 A and determine whether they would fall into the categories commonly referred to as dull or burnt deposits.

e
a
0

e
0
0

0 0
0 0
0
4 400
300 200
100
\

satin bright

'\
'\
\

, , V I .

" ,. bi .z .
I I I

I I

a
0
0 0 0 0

-z;r

The grain structure and grain orientation parallel to the surface were studied using XRD. Fig. 57 shows the 8/20 scans for bright, satin-bright and matte films. The diffraction peaks marked 's' were the result of the Nihrass substrate. The deposits all exhibited XRD patterns typical for (p); however, very different crystalline orientations were observed. The bright tin showed predominantly { lOl} , { 112) and { 103) orientations, whilst the satin-bright tin consisted of {220}, (21 l}, {301},
55

a
0

0 0
{321}, {312}, {431), (440) and (521) orientati ns. It was interesting to note that there were no common orientations between the bright and satin-bright films.

0 0'

Satin bright

0 0'
6 0 0 0 0 0 6 0 0 0 0 0 0 0 0 0 0 0 0
0

Matte

Bright
8

'

'

'

40

50

60

70

80

90

1 3

28

Fig. 57. X-ray diffraction spectra of electroplated tin films [47]. The matte film, by contrast, exhibited essentially a combination of the crystal faces of the other two and a gloss factor lower than the satin-bright. Apparently, the addition of crystal orientations typical for bright tin to that of satin-bright tin does not increase its gloss, but rather results in a matte deposit. This indicates that there is no simple and direct correlation between specific crystal orientations, their intensity and surface appearance, as discussed by Sun et al. [521 and any quantification based on these parameters could be misleading. The data, however, did indicate that the number of crystal orientations seemed to correlate with the glossiness of the surface. The bright tin had only three preferred orientations, while the satin-bright and matte films had 8 and 11 crystal orientations, respectively. This observation is consistent with the general assumption that an increase in the number of crystal faces results in a 'rougher' surface and decreased gloss. Grain size is another important parameter that was considered in the analysis. Scanning electron microscopy was used to provide evidence of grain size and surface morphology. The results are summarised in Fig. 58. The bright tin exhibited a fine structure with a much smaller grain size than either the matte or the satin-bright tin. It was interesting to note that the matte tin had a significantly smaller grain size relative to the satin-bright tin, although its surface was less glossy. Once again, this confirmed that grain size is only one of the parameters that affect surface appearance by modifying the surface roughness. In this regard it was also noteworthy that the 'texture' (i.e. preferred orientation and morphology) of the matte tin was significantly different fiom that of the satin-bright tin.

I 0 urn

briaht

satin

matte

Fig. 58. SEM images of various electroplated tin films 471.


56

0 0 0

0
0

0 0

0
0 0 0

Atomic force microscopy was used to provide topographic mapping of the electroplated samples. The images are shown in Fig. 59. Consistent with the SEM results, the AFM images showed bright tin as a fine grained structure with grain size in the range 50 to 1OOnm. The satin-bright and matte surfaces, on the other hand, had a much larger grain size of several microns. The arithmetic roughness, calculated from the AFM images over a 100 x 100p.marea, was 19nm for bright, 113nrn for satin-bright and 176nm for matte tin.

e
0

0 0 0

0
0

a
Fig. 59. AFM images of the electroplated films [47].

0
0

0
0

Fig. 60 is a plot of gloss reflectance measured at 60" vs. surface roughness via AFM. Once again, as shown in Fig. 56, there exists a strong exponential correlation between glossiness and surface roughness; however, the AFM data is better in distinguishing the surface topography of the satin-bright and matte samples.

800

a
0
0 0 0
0 200
400

600 ,800 1000 1200 1400 1600 1800

Surface roughness, A

0
0

Fig. 60. Gloss at 60" vs. surface roughness measured by AFM [47].
57

0
In conclusion, Xu et al. deduced from their work that the appearance of electroplated tin can be quantitatively described by measuring the gloss reflectance at a light incident angle of 60". The grain structure, orientation and grain size, have an indirect effect on gloss by modifying the surface morphology and particularly the surface roughness. The relationship of the gloss with the various parameters is shown schematically in Fig. 61.

0
0

0
Sn Electroplating

0
Nucleation and Growth
, I
I

0 0
0 0 0
0

Grain Orientation

Fig. 6 1. Factors influencing the surface appearance of electroplated tin [47].

0
0
7. Conclusions and Recommendations

The current review highlights aspects of the electrodeposition and electroplating of tin in which either progress in understanding or significant developments have been made during the past few years. It is clear that the topic continues to receive much attention in both scientific publications and the patent literature. The following is a summary of general observations considered by the reviewer to have particular merit. Recommendations for future research effort are also provided :-

0
0

0
0

A) In the field of fundamental investigations, attention continues to be focused on the problem of obtaining a better understanding of the mechanism@) of electrocrystallisation of tin on a variety of substrate materials. This work pertains for both the early and later stages of electrodeposition. The mechanisms are described in terms of either instantaneous or progressive nucleation processes and furthermore in terms of whether two or three dimensional nuclei are thought to be involved. Attempts have also been made to link these mechanisms with preferred morphologies. However, the elucidation of the relationship between a particular electrochemical mechanism and the exact morphology observed, is a goal that requires far more effort in the future.

0
0

e
0
0

a
0

0
0
58

0 0
0 0
In the realms of fundamental studies, a new galvanic technique to obtain tin nanoaggregates on an insulating surface has been described. The nanoaggregates show a dendritic structure composed of polycrystalline grains with size down to 20 nm. These experiments pave the way for possible applications of tin thin films obtained directly on insulating substrates by electrodeposition, i.e. galvanic rather than electroless routes. A considerable proportion of the current review has been concerned with the technologically important field of the role of additives in the electrodpositiodplating process. A range of compounds have been studied using a variety of electrochemical techniques and much evidence for their effectiveness as, for example, grain refiners, etc., has continued to be accumulated. However, most of the basic studies have been made in isolation with little regard to the important question of comparative effectiveness and the role of the chemical nature of the additives themselves. Much literature therefore contains references to individual additives having completely unrelated structures.

0 0 0
0

0 0

0
0 0

e a
0
0

0 0

(D) Despite the application of numerous and very sophisticated electrochemical and surface analytical techniques, any detailed understanding of the mechanisms by which additives affect the morphology of the electrodeposit, continues to be lacking. Although various possiblilities exist, including the requirement for the additive to be adsorbed on the substrate, or complexed in some way with the depositing species, only very basic conclusions appear to have been made regarding the exact mechanism(s) pertaining during the electrodeposition process. It is suggested that much greater effort is required to elucidate these mechanisms and in particular, to explain the apparent specificity of the compounds in question. (E) The patent literature contains a considerable number of references in which various additives are proposed constituents of the electrolyte formulations, particularly for high speed plating. Again, the chemical systems in question have little in common, although all are claimed to be highly effective. Information relating to the mechanism of operation of these compounds is also lacking. It is suggested that, in view of the diversity and complexity of these systems, they have been discovered mainly by empirical reasoning. Nevertheless, they must be regarded as highly significant in the context of the subject as a whole. (F) One of the constituents of recently proposed plating baths that is common to many electrolyte formulations, however, is methane suiphonic acid (or other sulphonic acids). An appreciable patent and scientific literature has appeared in which these acids form part of the basic electrolyte. Additives may also be derived from these acids, e.g. a-naphthosulphonic acid, and these have been shown to exert beneficial effects on deposit morphologies. The mechanism by which this particular additive is thought to participate has very recently been investigated in some detail, although further work is required.

0 0

0
0

0
0

0
0

0 0

0
0 0
59

0 0

0
(G) Equally important with bath composition in determining the coating morphology of electrodeposited tin are the processing variables. These include the electrode potential, the current density and the hydrodynamic conditions at the electrode surface. With the help of rotating disc and rotating cylinder electrodes, considerable progress has been made in the last few years to establish how these variables determine the ultimate morphology developed in an electrotinning process. This work is technologically very important. The morphologies are commonly referred to as either dendritic (observed at high current densities and/or low rotation speeds); micronodular . (observed at intermediate current densities and/or intermediate rotation speeds); or polycrystalline (observed at low current densities and/or high reotation speeds). It has been shown that one factor that can be used to predict the preferred morphology of the electrodeposit is the ratio of the applied current density to the limiting current density. The smaller this ratio, the higher the current efficiency and the greater the likelihood of compact deposits being produced. So called critical current ratios have been proposed above which signifcant changes in deposit morphology occurs.

0 0

0 0
0

0
0

0
0

(H) The preoccupation with utilising the plating current and hydrodynamics as the independent variables primarily responsible for determining the deposit morphology stems fi-om the need to control commercial plating processes, where tin is deposited at constant current densities on high speed plating lines. These processes can be simulated in the laboratory using a rotating cylinder electrode. However, fundamental electrocrystallisation theory teaches that the electode potential is the independent variable that determines nucleation and subsequent growth characteristics of the deposit and not the current value, since this parameter is simply a measure of the overall rate of the deposition. In the opinion of the reviewer, future studies of a more commercial nature need to appreciate this fact much more carefully and it is interesting to note that the role of the electrode potential has begun to feature more strongly in some of the most recent publications. (I) The importance of the plating current and hydrodynamics in the commercial world of tin plating has also been demonstrated in the field of quality control. Using a rotating cone electrode it has recently been shown that the effects of these parameters, together with changes in bath composition, can be determined quickly and easily. The cone geometry produces a range of current densities along the electrode surface, with the highest current being imposed at the cone apex and decreasing currents between the cone apex and the cone base. Thus, the effects of current density on the deposit morphology can be observed for a given electrolyte, temperature and hydrodynamic condition, in a single experiment by observing the changes in morphology along the electrode surface. This technique is therefore particulary useful when a large range of exprimental parameters need to be screened in a short time,

a a e
0

e
0

0 0

0
0

0 0 0 0
0

0
0

a
60

0 0

0
0

0 0 0 0 0

0
0 0 0
0

(J) The appearance of electrodeposited tin has been classified commercially as burnt, matte, satin-bright and bright, amongst other descriptions. This 'appearance' classification is qualitative and subject to interpretation. It is therefore desirable to attempt a more quantitative assessment of appearance based on physical measurements. Recently, it has been shown that the appearance of electroplated tin can be quantitatively described by measuring the gloss reflectance at a light incident angle of 60". The grain structure, grain orientation and grain size, all have an indirect effect on the gloss by modifying the surface morphology and, in particular, the surface roughness. If this relationship is accepted, then it is concluded that it would be useful to consider surface roughness as the factor controlling appearance in fundamental studies of the electrocrystallisation and electrodeposition of tin. This in turn would permit the appearance of the deposit to be understood and therefore controlled much more effectively by carefully adjusting the electrochemical and hydrodynamic parameters.

a
8
0
0

a
0 0 0 0 0 0
0

e
0
0
0 0

0
0
0
61

, _ .,

_.

_-

z _ ,

..--..,. ".~-. . -.--

- - , . . - ...

.,

8. REFERENCES
1. Tin Electrodeposition on Carbon Electrodes. From Nuclei to Microcrystals. E. Gomez. E. Guaus, F. Sanz and E. Valles, J. Electroanal. Chem. 456 (1999) 63-71. 2. Microstructure o Pure Tin Electrodeposited Films. T. Teshigawara, T. Nakata, K. f Inoue and T. Watanabe, Scripta mater. 44 (2001) 2285-2289. 3. I Fleury, Patent PCT/FR00/02757, 2000. ? 4. Preparation o Dendritic Tin Nanoaggregates by Electrodeposition. T. Devers, I. f Kante, L. Allam and V. Fluery, J. Non-Crystalline Solids 321, (2003) 73-80.
5 . In-situ

STM Studies on Underpotential Deposition o Sn on Au(lO0). J.W. Yan, J. f

Tang, Y. Y . Yang, J. M. Wu, et al., Surface and Interface Analysis, 32 (2001) 49-52.

f 6. Adsorption Behaviour o Laprol2402C on a Tin Electrode. A. Survila, Z. Mockus and S. Kanapeckaite, Trans. Inst. Met. Finish. 80 (2002) 85.
7. Electroplating o Tinfrom Acid Gluconate Baths. S. S. Abd El Rehim, S. M. f Sayyah and M. M. El Deeb, Plating and Surface Finishing, 87, no. 9, (2000) 93-98.

8. Initial Stages o Tin Electrodeposition from Sulphate Baths in the Presence o f f Gluconate. J. Torrent-Burgues, E. Guaus and F. Sanz, J. Appl. Electrochem. 32 (2002) 225-230.

f 9. Electrodeposition-Dissolution o Tin onto Low Carbon Steel Substrate in Acidic Electrolyte Bath Containing Quaternary Ammonuim Salts. R. C. B. Da Silva, M. D. Capela and T. M. C. Nogueira, Bull. Electrochem. 15 (1999) 561-565.
f f 10. Influence o M l 2 Organic Additive on the Electrodeposition o Tinfrom Acid Sulphate Solution. S. Bakkali, T. Jazouli, M. Cherkaoui, et al., Plating and Surface Finishing, 90, no. 1, (2003) 46-49.

0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
v
A

11. Electrodeposition Tinfrom Sulphate Electrolyte with Organic Additives. G. I. Medvedev and N. A. Makrushin, Zhur. Prikl. Khim., 74, no. 11,2001, 1787-1790. 12. Electrodeposition Tinfrom Sulphate Electrolyte with Organic Additives. G. I. Medvedev and N. A. Makrushin, Russian Journal of Applied Chemistry, 74 (1 1) (2001) 1842-1845. 13 Electrodeposition Tinfrom Sulphate Electrolyte in the Presence o Synthanol, f Formalin and Propargyl Alcohol. Medvedev, N. A. Makrushin and A. N. Dubenkov, Russian Journal of Applied Chemistry, 75 (2) (2002) 219-222. 14. A Study o the Kinetics o Tin Electrodeposition from Sulphate Electrolyte with f f Organic Additrives. Medvedev and N. A. Makrushin Zhur. Prikl. Khim 75 (8) (2002) 1234-1236.

62

0 0 0 0 0 0 0 0 0 0

0 0 0
0 0 0
15. Selection of Organic Substances for Production of Tin Lustrous Coatings. Medvedev, N. A. Makrushin and A. N. Dubenkov, Zhur. Prikl. Khim, 75, no. 11, 2002, 1834-1838. 16. The electroplatingfrom Sulphate Electrolyte Containing Synthanol, Formalin and Coumarin, G. I. Medvedev, L. A. Nekrasova and N. Makrushin, Russian Journal of Applied Chemistry, 76, no. 3, 2003, 387-390. 17. Synergistic Effects of Organic Additives on the Discharge, Nucleation and Growth Mechanisms of Tin at Polycrystalline Copper Electrodes. F. J. Barry and V. J. Cunnane, J. Electroanal. Chem. 537 (2202) 151-163. 18. Metal methaneosulphonate: Applications and Opportunities. D. Guhl and F. Honselmann, Metalloberflache 54, no 4, 2000, 34-37. 19. Methanesulphonic Acid in Electroplating and Related Metal Finishing Industries. R. Balaji and M. Pushpavanam, Trans. Inst. Met. Finish. 81, no 5,2003, 154-158. 20. A Comparative Study of Halogen and Methanesulphonic Acid Electrotinning Processes. Y-H. Yau, Plating and Surface Finishing, 86, no. 8, 1999,48. 21. The Effect of Process Variables on Electrotinning in a Methanesulphonic Acid Bath. Y-H. Yau, J. Electrochem. Soc. 147 (2000) 1071-76. 22. R. Winand, Hydrometallurgy, 29 (1992) 567. 23. D. R. Gabe, Plat. Surf. Finih, 82 (1995) 69. 24. E. N. Andrade, Nature 125 (1930) 309. 25. Spectroelectrochemical Study of the Effect of Organic Additives on the Electrodeposition of Tin. I. S. Zavarine, 0. Khaselev and Y. Zhang, J. Electrochem. Soc. 150 (2003) C202-C207. 26. G. A. Hope and G. M. Brown, in Proceedings of the 6thInternational Symposium on Electrode Processes, A. Wieckowski and K. Itaya, Eds., P V 96-98, p 215. The Electrochemical Society Proceedings Series, Pennington, NJ (1996). 27. P: A. Kohl, J. Electrochem. Soc. 129 (1982) 1196. 28. G. S. Tzeng, Plat. Surf. Finish. 82 (1995) 67. 29. E. E. Famdon, F. C. Walsh and S. A. Campbell, J. Appl. Electrochem. 25 (1995) 574. 30. J.-W. Kim, J.-Y. Lee and S.-M. Park, Langmuir 20 (2004) 459. 3 1. C. J. Van Velzen, M. Sluyters-Rehbach and J. H. Sluyters, Electrochin. Acta, 32 (1987) 815. 32. R. Moshorontou, I. Tsangaraki and C. Kotsira, Plat. Surf. Finish. 81 (1994) 53.
63

0 0 0 0

a
0
0
0

0 0 0 0

e
0

0
0 0

a
0 0

0 0 0 0

a
0
0

33. T. C. Franklin, Ibid.,81 (1994) 34. Effects of Ethoxylated a-Naphtholsulphonic Acid on Tin Electroplating at Iron Electrodes. J.-Y. Lee, J.-Woo Kim, B.-Y. Chang and S.-M. Park, J. Electrochem. SOC. (2004) C333-C341. 151 35. A. Aragon, M. G. Figueroa, R. E. Gana and J. H. Zagal, J. Appl. Electrochem. 22 (1992) 558. 36. G. Gunawardena, G.J. Hills, I. Montenegro and B. R. Scharifker, J. Electroanal .Chem. 138 (1982) 225. 37. B. R. Scharifker and G. J. Hills, Electrochim. Acta, 28 (1983) 879. 38. M. Palomar-Pardve, Ma. T. Ramirez and B. R. Scharifker, et al. J. Electrochem. Soc. 143 (1996) 1551. 39. A Unique Electroplating Chemistv. Y. Zhng and J. A. Abys, Circuit World 2511 (1998) 30-37. 40. Electrochemical dissolution of Tin in Methanesulphonic Acid Solutions. R. A. T. de Greef and L. J. J. Janssen, J. Appl. Electrochem. 31 (2001) 693-702. 41. Tin, Lead or Tin/Lead Alloy Electrolytes for High Speed Electroplating. M. P. Toben, N. D. Brown, D. J. Ester1 and R. A. Schetty, Patent EP 0319997, June 1989. 42. Tin Electroplating Process. S. H. Hong and Y. Zhang, Patent EP 1006217, June 2000. 43. Tin Plating Electrolyte Compositions. C. H. ODriscoll, Patent US 62 17738, April 2001. 44. H. D. Gilman, B. Fernandez and K. Wikiel, Patent US 2002014414, Feb 2002. 45. Tin Plating. J. N. Crosby, Patent EP 1260614, November 2002. 46. Testing of Tin Plating by use of a Rotating-Cone Electrode. D. Reichenbacher, Plat. Surf. Finish. 46 (1999) 111-114. 47. Surface Morphology, Appearance and Tribology of Electrodeposited Tin Films. C. Xu, Y. Zhang, C. Fan, P. Chiu and J. A. Abys, Ibid., 87 (2000) 88-92. 48. ASTM E284-93a: Standard Terminology of Appearance. 49. Y. Zang, J. A. Abys, C. H. Chen and T. Siegrist, Plat. Surf. Finish. 85 (June1998) 105

50. Y. Zang and J. Abys, Circuit World, Oct 1998.

64

0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 6 0 0 0 0 0 0 0 0 0 0

a
0 0

e
0 0 0

51. J. A. Abys, C. Fan and I. Kadija, Patent US 5,916,696.


52. M. Sun, K. Liao, J. Swanson, G. Federman and F. Bottos, Proc. 1999 AESF Continuos Steel Strip Plating Symposium (May 1999).

e
0

e
0

a
0

e
0 0 0 0 0 0 0 0 0

e e
0 0 0 0 0 0 0

65

9.

Acknowledgements

The reviewer wishes to acknowledge with gratitude the following for the use of figures in this report :. /

Elsevier, Journal of Electroanalytical Chemistry, Figs. 1-4; Scripta Materialia, Figs. 5-8; & Journal of Non-Crystalline Solids, Figs. 9-1 1. Maney Publishing, Transactions of the Institute of Metal Finishing, Figures 12 & 13. American Platers and Surface Finishers Association, Plating and Surface Finishing, Figs.l4,21-23,24-30, 50-52 and 53-61. Kluwer Publications, Journal of Applied Electrochemistry, Figs. 15-17. Central Electrochemical Research Institute, Bulletin of Electrochemistry, Figs. 7, 8 and 13. Electrochemical Society, Journal of the Electrochemical Society, Figs. 30-49.

0 0 0 0 0 0

0 0 - 0

0
0 0 0 0 0

66

You might also like