You are on page 1of 12

Materials Science and Engineering B65 (1999) 111 122

www.elsevier.com/locate/mseb

Development of an electroplating solution for codepositing Au Sn


alloys
W. Sun, D.G. Ivey *
Department of Chemical and Materials Engineering, Uni6ersity of Alberta, Edmonton, Alta, Canada T6G 2G6
Received 14 May 1999; received in revised form 23 June 1999

Abstract
A relatively stable, weakly acidic, non-cyanide electroplating solution has been developed for deposition of Au Sn alloys over
a range of compositions. The solution consists of Au and Sn chloride salts, as well as ammonium citrate as a buffering agent and
sodium sulphite and ascorbic acid as stabilizers. Electrochemical studies have been conducted to examine the effects of the various
additives and their concentrations on bath stability and plating behaviour. Preliminary electroplating experiments with the
developed solution indicate that uniform, homogeneous deposits can be achieved over a range of compositions, including the
technologically important eutectic and near eutectic values. 1999 Elsevier Science S.A. All rights reserved.
Keywords: Au Sn alloys; Electroplating; Solders

1. Introduction
Goldtin eutectic solders are commonly used in the
optoelectronic and microelectronic industries for chip
bonding to dies. Au Sn solder is classified as a hard
solder with superior mechanical and thermal properties
relative to soft solders, such as the Pb Sn system.
AuSn solder can be applied in a number of ways, i.e.
as AuSn preforms, solder paste, by sequential evaporation and sequential electrodeposition. Compared with
solder preforms and pastes, evaporated solder is cleaner
and provides more precise thickness and positional
control. Thin film deposition technology, however, involves expensive vacuum systems. Electroplating of
AuSn eutectic solder is an attractive alternative in
that it is a low cost process, offering the thickness and
positional control of thin film techniques. AuSn solder layers have been produced sequentially by depositing Au first on a seed layer, followed by Sn [13].
Commercially available Au and Sn baths are utilized
from which several microns of solder can be deposited.
Co-electrodeposition of Au and Sn from a single solution offers the same economic advantage of sequential
plating relative to vacuum deposition techniques, as
* Corresponding author. Fax: +1-403-493-2881.

well as the prospect of depositing the solder in a single


step without oxidation of an outer Sn layer.
The technology for Au and Sn plating is quite well
developed and will be briefly reviewed here.
Electrodeposition of soft Au on electronic devices
and components is generally performed using a bath
containing cyanoaurate (I) ions, because Au cyanide
complexes have the highest stability coefficients, e.g. the

stability coefficients for Au(CN)


2 and Au(CN)4 are
38
56
2 10 and :10 , respectively [4]. Free cyanide ions
generated as a result of the Au deposition process
attack the interface between the resist film and substrate, lifting the resist and depositing extraneous Au
under the resist [5]. Because of this incompatibility,
work has focused on developing non-cyanide baths.
Au(I) sulphite complexes have better compatibility towards positive resists and the added benefit of improved throwing power and deposit thickness
uniformity compared with cyanide baths [6]. In addition, deposits from Au sulphite solutions are bright,
hard and ductile. The Au(I) sulphite complex is subject
to a disproportionation reaction, however, forming
Au(III) and metallic Au, which causes the bath to
decompose spontaneously on standing.
3[Au(SO3)2]3 = 2 Au+[Au(SO3)4]5 + 2 SO23

0921-5107/99/$ - see front matter 1999 Elsevier Science S.A. All rights reserved.
PII: S 0 9 2 1 - 5 1 0 7 ( 9 9 ) 0 0 2 2 3 - 8

112

W. Sun, D.G. I6ey / Materials Science and Engineering B65 (1999) 111122

To prevent decomposition, a suitable stabilizing additive is needed. The first commercial sulphite Au plating solutions were developed in the early to mid 1960s.
The sulphite ion is itself in equilibrium with sulphur
dioxide according to
SO23 +H2O = SO2 (g) +2 OH
Because the above reaction forms hydroxyl ions, the
equilibrium is pH-dependent. Most commercial solutions operate at pH values above :9.5 (e.g. see [7]).
When Au is plated out of solution at alkaline pH, the
excess sulphite remains and can be oxidized to sulphate
at the anode. There have been several attempts to
reduce the operating pH to below neutral for applications involving alkaline-developable photoresists [7
10]. The addition of organic polyamines, such as
ethylenediamine, can be used to lower the pH to acidic
values, allowing controlled evolution of sulphur dioxide
to remove a portion of the excess sulphite [8,9,11].
The possibility of electroplating soft Au from a noncyanide bath containing both thiosulphate and sulphite
as complexing agents has been explored [10,12,13]. The
bath reported by Osaka [10] operates at a pH of 6.0
and a temperature of 60C. The bath is reported to be
stable, although no specific stability data has been
given. Three different Au complexes can exist in this
system.
Au+ +2 SO23 = [Au(SO3)2]3
Au+ +2 S2O23 =[Au(S2O3)2]3

b =1010
b =1026

Au+ +SO23 +S2O23 =[Au(SO3)(S2O3)]3


b =unknown
b is the stability coefficient for the complex. Thallium(I) ions have been added in the form of Tl2SO4 as a
grain refiner to improve the surface morphology of the
deposit.
Phosphates, carbonates, acetates and citrates are
commonly used as buffering and conducting agents for
Au plating baths. In alkaline Au sulphite baths, metals
such as Cd, Ti, Mo, W, Pb, Zn, Fe, In, Ni, Co, Sn, Cu,
Mn and V in various concentrations are used as brightening additives, while Sb, As, Se and Te semi-metals are
also used [14].
There are two types of Sn plating solutions: alkaline
and acidic [15]. Alkaline solutions are based on sodium
or potassium stannate. Hydrogen peroxide or sodium
perborate is used to oxidize any stannite (bivalent Sn)
to the stannate form. Alkaline baths are superior to
acid baths in throwing power. Acidic plating baths
contain Sn in the bivalent form, using metal salts that
are sulphates, fluoborates and fluosilicates. Electrodeposition of Sn from a stannous Sn solution has the
obvious advantage of consuming less electricity (half
the amount at 100% efficiency) compared with a stannate bath. The problems with acidic baths include poor

throwing power and solution instability, with basic tin


compounds precipitating on standing. Various additives, including gelatin, glue, cresol, sulphonic acid and
aromatic hydroxyl compounds, have been used to improve plating quality. When an acidic bath ages, the
bath may change colour to darker yellow and may also
become turbid. The actual chemistry of this change is
relatively poorly understood, but is attributed to the
formation of stannic compounds when stannous Sn salt
is oxidized to stannic Sn in the presence of dissolved air
and elevated temperature. The stannic compounds are
colloidal and very difficult to remove. Oxidation of
stannous Sn can be minimized by maintaining the
solution temperature at 2025C, using an airtight
plating setup and adding a suitable anti-oxidant such as
a phenol compound.
The available information concerning the electrodeposition of AuSn alloys is virtually confined to the
patent literature. One of the problems with AuSn
alloy plating baths is preventing the oxidation of Sn(II)
to Sn(IV) [16]. Oxidation of Sn can be minimized by
using soluble Sn anodes, however, Au is deposited on
the anodes unless they are isolated by semi-permeable
diaphragms.
An early report by Raub and Bihlmaier stated that
AuSn alloys containing up to 30 at.% Sn could be
deposited from baths containing no free cyanide, and
containing the Sn as its stannate complex formed with
KOH [17]. Later claims concerning AuSn alloy plating, however, have been based on the use of alkaline
and acid cyanide electrolytes, where Sn in many cases
has been incorporated with the goal of obtaining
brightening effects rather than producing deposits with
significant amounts of Sn.
Several cyanide based systems have been reported
[1821]. Frey and Hempel [18] developed a bright
AuSn plating bath with a pH of 314, comprised of
potassium dicyanoaurate, soluble Sn(IV), potassium
hydroxide, potassium salt of gluconic, glucaric and/or
glucaronic acid, conductivity salt, piperazine and a
small amount of As. The bath was used to plate small
parts with an alloy containing 525 wt.% Sn. Bright
deposits were obtained for thicknesses greater than 0.1
mm and the solution exhibited long term stability without the use of soluble Sn anodes. Another patent using
potassium dicyanoaurate and tin chloride has been
reported with a deposit composition of 8 wt.% Sn and
thickness of 5 mm claimed [19].
AuSn codeposition from a cyanide system using
pyrophosphate as a buffering agent was studied by
Kubota et al. [20,22]. The basic formula consisted of
K4P2O7, KAu(CN)2 and SnCl22H2O. The mass transfer was investigated to clarify reaction mechanisms
between monovalent Au or bivalent Sn and pyrophosphate ions, by measuring conductivity, kinematic viscosity and limiting current density of the bath

W. Sun, D.G. I6ey / Materials Science and Engineering B65 (1999) 111122

components [2]. Two pyrophosphate ions were complexed with one stannous ion, with excess pyrophosphate acting as a supporting constituent.
Tanabe et al. [21] obtained various Au Sn alloy
compositions by electrodeposition from cyanide baths
containing HAuCl44H2O, K2SnO33H2O, KCN and
KOH. Although a linear relationship was not found
between the Sn content in the bath and the Sn content
in the alloy formed, a relationship was found between
the two alloys which permitted formation of alloys of
desired compositions. The composition of electrodeposited AuSn was shifted by about 10% to the Sn side
in comparison with alloys at thermal equilibrium; thus
exhibiting the j phase in the 25 29 at.% range. AuSn,
AuSn2 and AuSn4 were also electrodeposited.
Gold chloride electrolytes were used in the early days
of Au plating, but today are employed almost exclusively in the electrochemical refining of Au. An extensive investigation [23,24] of the cathodic behaviour of
Au in chloride solutions has shown that the quality of
the cathode deposit is strongly influenced by the relative amounts of Au(I) and Au(III) in the solution. The
reduction of Au(III) chloride to the metal can be expected to involve the formation of Au(I) as an intermediate species. Under plating conditions, Au will be
deposited from both the Au(III) and Au(I) species.
Since Au(I) has a more positive plating potential (1.154
V) than Au(III) (1.002 V) [4], a limiting current density
for Au(I) will be reached first and it can be expected
that the deposits will be of relatively poor quality, i.e.
they tend to be bulky and porous. Gold fines will be
present in the solution as a result of the following
disproportionation reaction:

113

Au and Sn over a range of compositions. The solution


that was chosen as a starting point was the chloride
system reported by Matsumoto [25]. Preliminary experiments were carried out to obtain the stable solution
described in the patent. The solution deteriorated immediately when Sn salt was added to the ammonium
citrate buffered Au solution. To prepare a clear and
stable solution for AuSn alloy plating, a screening test
was conducted to select suitable stabilizers from several
candidates. The effect of the selected stabilizers was
studied in an effort to understand the roles of the
stabilizers and to obtain reference information regarding the current response to cathodic potential change,
open circuit potential, limiting current density and the
potential at which the limiting current density is approached. A proper set of plating parameters can be
derived from this information.
A causticity test for semiconductor wafers in the
developed solution was performed to examine the causticity of the solution towards the wafers to be plated.
Preliminary plating experiments were performed with
the developed solution using metallized InP wafers.

2. Experimental methods
All solutions for this study were prepared with deionized water. Any precipitated products were examined in
an Hitachi S-2700 scanning electron microscope (SEM)
equipped with a Link energy dispersive X-ray (EDX)
spectrometer for composition analysis.
The starting point was a solution based on the Matsumoto patent [25], which is listed below:

3 AuCl2 =2 Au +AuCl4 +2 Cl
200 g l1
Detailed studies of the anodic and cathodic reactions
have shown that the use of low temperatures and
periodic interruption of the current are major factors
that can contribute to reduced Au(I) concentration.
A recent patent [25] reported on a Au Sn plating
bath (pH 37) containing KAuCl4, SnCl2, triammonium citrate, L-ascorbic acid, NiCl2 and peptone. A 7
mm AuSn alloy (20 92 wt.% Sn) layer was plated out
on a 50 mm diameter Si wafer at 20C and a current
density of 0.6 A dm 2 in 30 min using a Pt-coated
non-consumable Ti anode.
In another patent [8], it was claimed that in a Au
sulphite plating bath (pH: 6.5) containing an organic
polyamine, or a mixture of polyamines and an aromatic
organic nitro compound, tetravalent Sn may be added
as sodium or potassium stannate. However, no examples of plating results were given, particularly with
respect to bath stability and deposit composition.
The goal of this work was to develop a stable, weakly
acidic, non-cyanide bath suitable for co-electroplating

20 g l1
13 g l1
30 g l1
1 g l1
5 g l1

ammonium citrate
(H4NO2CCH2C(OH)(CO2NH4)
CH2CO2NH4)
KAuCl4
SnCl22H2O
L-ascorbic acid (HOCH2CH(OH)
(C(H)OC(O)C(OH)C(OH))
NiCl2
Peptone

Since the initial solution was unstable, suitable stabilizers had to be found. Three stabilizers were chosen as
candidates, Na2SO3 (20100 g l 1), Na2S2O3 (20100 g
l 1) and Na2H2EDTA2H2O (540 g l 1). Based on
the screening tests, sodium sulphite was selected as a
Au stabilizer for additional tests. L-ascorbic acid was
chosen to prevent Sn hydrolysis. Solution lifetimes were
measured for different concentrations of additives.
Polarization testing of the cathode in Au solutions,
Sn solutions and Au-Sn solutions was done on an
EG&G Potentiostat/Galvanostat Model 273. Only ca-

W. Sun, D.G. I6ey / Materials Science and Engineering B65 (1999) 111122

114

thodic data is plotted in the subsequent figures. The


cathode was a piece of InP wafer metallized with a
25 nm Ti adhesion layer and a 200 nm Au coating.
Stop-off laquer was used to define a 1 1 cm exposed area of Au coated surface and to cover the
backside of the sample. Platinum foil was used as the
anode and a saturated calomel electrode (SCE) as the
reference electrode. The cathode potential scanning
rate was set at 0.5 mV s 1.
Causticity tests for InP and GaAs were performed
in a plating solution developed from this work. Wafer
pieces, without metallization layers, were weighed
prior to and after immersion for 48 h. Atomic force
microscopy (Digital Instruments Nanoscope E AFM)
was used to check the surface morphology before and
after immersion.
Preliminary plating experiments were carried out on
the developed plating solution. Depositions were obtained at a fixed temperature (20C) under both direct
current (DC) and pulsed current (PC) conditions. The
cathodes were sectioned InP wafers coated with a Ti
(25 nm)/Au (250 nm) blanket metallization. The current density was varied from 1.6 to 3.6 mA cm 2
and the plating time was 60 min. For PC plating, a
constant average current density of 2.4 mA cm 2 was
used, along with a cycle period of 10 ms and an on
time of 2 ms. All electroplated samples were examined in the SEM and deposit composition was
checked by EDX analysis. An accelerating voltage of
20 kV was used for both imaging and composition
analysis; pure Au and pure Sn standards were used
for quantitative analysis.

3. Results and discussion

3.1. Solution preparation


Initial solution preparation results are shown in
Table 1. If Sn chloride is mixed with water, without
any additives, the bivalent Sn chloride salt undergoes
hydrolysis according to:
Sn2 + + 2 H2O= Sn(OH)2 + 2H+
with a solubility product for Sn(OH)2 of 3 10 27
[26]. Solution A in Table 1 contained 30 g l 1 of
L-ascorbic acid, while solution B contained 200 g l 1
of ammonium citrate. Both solutions were weakly
acidic (pH: 6.5), which helps to minimize hydrolysis
preventing hydroxide precipitation. After 1 week
solution A became turbid, while solution B changed
to dark yellow from colourless, but remained clear.
The difference may imply that ammonium citrate is a
complexing agent for Sn2 + ions; however, no
information was found in the literature concerning
the complexing ability of ammonium citrate with
bivalent Sn ions. Although the actual chemistry for
the change in the solutions is not well understood,
the change is attributed to the oxidation of stannous
ions (II) by dissolved air to stannic ions (IV) and the
formation of stannic compounds. Higher temperatures
than room temperature result in increased oxidation
rates. It can therefore be concluded that without any
anti-oxidant additives, Solutions A and B are only
stable for about a week. The behaviour of bivalent
Sn ions in water is very complex. Possible forms of
Sn ions in a chloride solution include [SnCl]+,

Table 1
Solution preparation
Solution c

Solution

Observations

13 g l1 SnCl22H2O dissolved in 30 g l1 L-ascorbic acid solution

Clear solution with pH 1.7


Precipitation after 1 week

13 g l1 SnCl22H2O dissolved in 200 g l1 ammonium citrate solution

Clear solution with pH 6.5


Solution still clear after 1 week but turned
dark yellow

10 g l1 KAuCl4 dissolved in water

Solution turned black and turbid on standing


Precipitated fine black powder

10 g l1 KAuCl4 dissolved in water in darkness

Solution turned black and turbid on standing


Precipitated fine black powder

10 g l1 KAuCl4 dissolved in a 200 g l1 ammonium citrate solution

Clear solution and stable in light

Solution E added to B

Solution turned black and turbid on standing


Precipitated fine black powder

10 g l1 KAuCl4 dissolved in a 800 g l1 ammonium citrate solution and


then solution B added

Same phenomena as solution F

1. 10 g l1 KAuCl4 dissolved in a 800 g l1 ammonium citrate solution


2. 13 g l1 SnCl2.2H2O dissolved in 400 g l1 ammonium citrate solution
3. Solution (2) added gradually to solution (1) with vigorous agitation

Clear solution with dark green colour


Precipitation after a few hours

W. Sun, D.G. I6ey / Materials Science and Engineering B65 (1999) 111122

115

Table 2
Solutions utilized for bath stability tests

Ammonium citrate (g l1)


KAuCl4 (g l1)
Na2SO3 (g l1)
L-ascorbic acid (g l1)
SnCl22H2O (g l1)
Solution stability (days)

S1

S2

S3

S4

S5

S6

S7

S8

S9

S10

200
5

200
5

200
5
60

200
5
60
15
5
15

100
7
60
15
7
11

200
7
60
15
7
9

200
7
30
15
7
3

200
10
60
15
10
7

200
14
60

200

15
5
0

5
4

[SnCl2], [SnCl3] and [SnOH]+ with stability constants


of 14, 15, 50 and 1010, respectively [26].
KAuCl4 is soluble in aqueous solutions and is light
sensitive. Preparation of solutions C and D (Table 2)
shows that KAuCl4 undergoes hydrolysis both in light
and in darkness. The solutions precipitate a fine black
powder, which gradually changes to a gold colour on
standing. The powder was determined by EDX analysis
to be metallic Au. In aqueous solution, AuCl
4 ions are
hydrolyzed to some extent forming (AuCl3)H2O. This
in turn acts as a weak acid forming species such as
AuCl4 n (OH)n (where n varies from 0 to 4 and increases with increasing alkalinity) in alkaline solutions.
AuCl4 +H2O = (AuCl3)H2O +Cl
=AuCl3(OH) +H+ +Cl
The pH value of solution E containing 200 g l 1 of
ammonium citrate falls in the range of a weak acid. The
hydrolysis of KAuCl4 is prevented by the presence of
concentrated ammonium citrate. (NH4)+ hydrolyzes in
water,
(NH4)+ =NH3 +H+
and produces a significant amount of NH3 that dissolves in the solution. NH3 can form complex
Au(NH3)3 + cations with simple Au(III) ions (if any are
present) in the solution. The stability of Au(III) ions in
the solution is further improved. The stability constant
26
however, no stability constant data
for AuCl
4 is 10
3+
for Au(NH3)
is available in the literature.
Preparation of solution F (Table 1) was the first
attempt to make a Au Sn solution. It turned black and
turbid immediately after the Au solution (E) was added
to the Sn solution (B). The exact chemistry responsible
for the instantaneous precipitation of fine black powder
(gold) is not clear because of the lack of relevant
information. Still, it is reasonable to surmise that a
chemical interaction between Au ions and Sn ions
causes the problem. The chemical processes for Au
precipitation when Sn salt and Au salt are mixed can be
AuCl
ion reduction to AuCl
ions, followed by
4
2

AuCl2 ion dissociation.


3 AuCl2 =AuCl4 +2 Au + 2 Cl

30
14
7

Since ammonium citrate is able to complex Au ions,


solutions with more concentrated ammonium citrate
should be more stable. Preparation of solutions G and
H is the result of such an attempt. No improvement
was found for solution G, while solution H was the first
solution that remained clear after preparation. Solution
H was prepared by adding the Au solution gradually
instead of by pouring the entire Au solution in the Sn
solution. This implies that a high concentration of
ammonium citrate is needed to eliminate the chemical
reaction between Au(III) ions and Sn(II) ions. The way
that ammonium citrate works may be 2-fold, i.e. as
either a Au complexing agent or a Sn complexing
agent. Since a very high concentration of ammonium
citrate is needed to stabilize Au or Sn ions, it can be
surmised that it is not a strong complexing agent for
either Au(III) or Sn(II) ions. Solution H has two major
problems in terms of being used as a practical plating
solution. One problem is its short lifetime; the solution
deteriorated by precipitating only a few hours after
preparation. The other problem is the high viscosity of
the solution, due to the high concentration of ammonium citrate. High viscosity results in a slow mass
transport rate and therefore a lower limiting current
density. Although the improvement in solution H relative to the other solutions was minor, the key to
developing a stable AuSn solution seems to lie in
finding a more efficient Au complexing agent to decrease the oxidizing ability of Au ions when mixed with
the reducing agent, bivalent Sn.
Three stabilizers were chosen and tried based on the
above discussion. Na2SO3 and N2S2O3 are Au complexing agents, while Na2H2EDTA2H2O is a Sn complexing agent. Na2SO3 was more effective than Na2S2O3 at
reducing Au precipitation during the addition of Sn
salt. The Na2SO3 containing solution was clear and
stable for several days, while Au precipitation occurred
within a few minutes for the Na2S2O3 containing solution. Na2H2EDTA is a complexing agent for many base
metal impurities in plating baths. It is often used to
reduce contamination of the coating with base metals
where very pure Au deposits are required. The stability
constant for [SnEDTA]2 ions is 1022 [26], which is
fairly high. However, it fails to prevent interaction
between Au and Sn ions; Au precipitates on the wall of

116

W. Sun, D.G. I6ey / Materials Science and Engineering B65 (1999) 111122

the beaker within a few minutes of mixing the Au and


Sn solutions.
The stability constants for [Au(SO3)2]3 and
Au(S2O3)32 are 1010 and 5 1028, respectively [4]. Although the latter has a much higher stability constant,
it does not have a positive influence on solution stability in this case. This phenomenon may be due to its
incompatibility with other chemicals in the solution;
e.g. the buffering agent, ammonium citrate, which tends
to nullify the stabilizing effect of the stabilizer.
The complexes formed by Au(I) and Au(III) in sulphite media are expected to be of the types Au(SO3)32
and Au(SO3)54 , respectively. Pure Na5[Au(SO3)4]5
H2O has been isolated [27], while there is no unequivocal evidence for the presence of an Au(SO3)32 complex
in the solid state. However, the existence of the
Au(SO3)32 complex in sulphite baths has been reported
by Socha et al. [28]. In one method of preparing such
baths [29], for example, the precipitate, which is formed
by adding ammonia to a Au chloride solution, is simply
dissolved in an alkaline metal sulphite solution.
Gold(III) is reduced to Au(I) in the process. In the
1970s, Au(I) amine sulphite complexes stable at pH
levels as low as 4.5 were found. In the preparation of
such baths, the Au may be added in the form of solid
Au(III) amine-sulphite complexes, which are reduced in
situ by warming the bath before use [30]. The complex
Na[(H2N H2C CH2 NH2)AuIII(SO3)2] can be precipitated by mixing solutions of Au(III) chloride and excess
ethylene diamine, cooling and then adding a saturated
solution of sodium sulphite. The Au(I) amine sulphite
complexes, to which they presumably give rise on reduction, do not appear to have been isolated. Laude et
al. [31] patented solutions based on ammonium Au(I)
sulphite that could operate at pH 6 8 in the absence of
polyamines. The solutions were prepared by adding
ammonia to an HAuCl4 solution, which resulted in the
formation of a precipitate. This was followed by dissolving the precipitate in an aqueous ammonium sulphite solution. The precipitate, the nature of which is ill
determined, is called fulminating gold.
The method of preparing Au Sn sulphite solutions
in this work is a little different compared with the
methods described above. The Au is added in the form
of solid KAuCl4 salt that is dissolved in a concentrated
ammonium citrate solution. When Na2SO3 is added to
the solution, no precipitation occurs. It is presumed
that the Au(III) ions have been reduced to Au(I) ions.
The stability of the Au Sn solution was substantially
improved; no Au precipitation occurred when Sn salt
was added. Based on these results, Na2SO3 was selected
for stabilizing Au ions for subsequent work.
The experimental results for the effects of chemical
additives on Au Sn solution stability are summarized
in Table 2. Solutions S1 and S2, which contained no
sulphite, deteriorated immediately when Sn salt was

added. With 60 g l 1 of Na2SO3, Solution S3 remained


clear and stable for 4 days; after which it began to
gradually precipitate fine Au particles. Solution S4 was
the same as S3, except for the addition of 15 g l 1 of
L-ascorbic acid. The solution stability was improved to
15 days. Its stabilizing effect is quite surprising since
L-ascorbic acid was originally added to prevent Sn
hydrolysis. L-ascorbic acid only changed the pH from
: 6.5 to :6.0, since a high concentration of ammonium citrate, a buffering agent, was also present in the
solution.
Comparison of solutions S5 and S6 seems to indicate
that that the concentration of ammonium citrate has
very little influence on bath stability, which may be
because most of the Au ions are present in the form of
a Au sulphite complex. A lower citrate concentration is
favoured for practical plating, since the viscosity is
lower.
Comparison of solutions S4, S6 and S8, which contained gradually increased amounts of Au and Sn salts,
shows that the higher the total salt content, the shorter
the bath lifetime. Because the chemical reaction rate is
proportional to the reactant concentrations, the higher
the total concentration of reactants (Au and Sn ions),
the faster the Au precipitates from solution.
The effect of sulphite on bath stability is clearly
evident by comparing solutions S6 and S7. S7 contained less sulphite and its lifetime was shortened from
9 to 3 days. Since the stability constant for Au sulphite
is fairly low, free sulphite is required.
Another possible alternative to improve bath stability
is to prepare and store the Au and Sn solutions separately and mix them when plating is to be performed.
Solution S9 is a Au solution and S10 is a Sn solution.
If S9 and S10 are mixed at a 1:1 volume ratio, the
overall make-up would be the same as S7. The Au
solution has a lifetime of 78 days, after which Au
precipitation begins. The Sn solution turns light yellow
from colourless after about 7 days, but remains clear
for more than 30 days. The reason that Au still precipitates from solutions containing sulphite is that the
stability constant for the sulphite complex is not that
large and, with time, any free sulphite is oxidized by air
at the liquid/air interface. It would be expected that for
solutions used for plating the lifetime would be even
shorter because of sulphite consumption by anodic and
chemical oxidation and cathodic reduction. The change
in colour of the Sn solution is due to oxidation of
bivalent Sn to tetravalent Sn.

3.2. Polarization measurements


Polarization curves obtained for Au solutions (Fig.
1), Sn solutions (Figs. 2 and 3) and Au-Sn solutions
(Fig. 4) are discussed below and solution compositions
are given in Table 3.

W. Sun, D.G. I6ey / Materials Science and Engineering B65 (1999) 111122

117

Comparison of curves (a) and (b), solutions SAu-1


and SAu-2, respectively in Fig. 1 indicates that the
concentration of ammonium citrate does not markedly
affect the polarization behaviour of the cathode in the
Au solution. The limiting current density (:1 mA
cm 2) does not change significantly. When the cathode
potential is lower than : 0.7 V versus SCE (standard calomel electrode), organic decomposition at the
cathode takes place, plating carbon at the surface of the
cathode. The curves have only 1 wave in the Au plating
region, so it is likely that the cathode reaction is a

Fig. 2. Cathodic polarization curves for solutions SSn-1 (a), SSn-2


(b), SSn-3 (c). The dashed lines H1 and H2 represent the approximate
potentials for hydrogen evolution for curve (a) and curves (b) and (c),
respectively.

single step process.


Au(III)+3 e = Au(0)
A three step process for the reduction of Au(III) in
chloride solutions was proposed by Nicol et al. [23]:
AuCl4 = AuCl3 + Cl
AuCl3 + 2 e= Au(I) (slow step)
Au(I)+ e= Au(0) (fast step)
The slow step is not balanced. Because the role of the
chloride in the reaction is not clear, it is not indicated
on the right hand side of the equation.
Curves (c)(e), for solutions SAu-3 (30 g l 1 sodium
sulphite), SAu-4 (60 g l 1 sodium sulphite) and SAu-5
(100 g l 1 sodium sulphite), in Fig. 1 show that the
addition of sodium sulphite to the Au solution dramatically shifts the polarization curves to lower potentials.
Both the oxidizing ability of the Au ions and the ease of
plating Au have been greatly reduced. The effect of
sulphite concentrations in the solution is not very significant and cannot be correlated in a clear way. The
curves have a single wave for potentials higher than
0.75 versus SCE, where the cathodic reaction may be
Au(I)+ e= Au(0)

Fig. 1. Cathodic polarization curves for solutions SAu-1 (a), SAu-2


(b), SAu-3 (c), SAu-4 (d), SAu-5 (e), SAu-6 (f) and SAu-7 (g). The
dashed line indicates the approximate potential for hydrogen evolution.

The potential for carbon plating is also lowered to


: 0.75 V versus SCE. The suitable range for Au
plating in the three solutions is then about 0.6 to
0.75 V versus SCE, with a limiting current density of
less than 2 mA cm 2. At potentials higher than 0.6
V versus SCE, the Au plating speed is very low, while
at potentials lower than 0.75 V versus SCE, Au
plating is diffusion controlled which is usually not
desirable for practical plating. At this potential, hydrogen evolution also begins, resulting in a lower current

118

W. Sun, D.G. I6ey / Materials Science and Engineering B65 (1999) 111122

Fig. 3. Cathodic polarization curves for solutions SSn-3 (a), SSn-4 (b), SSn-5 (c), SSn-6 (d), SSn-7 (e) and SSn-8 (f). The dashed line indicates the
approximate potential for hydrogen evolution.

efficiency and a local increase in pH, which can cause


Sn hydrolysis and incorporation of Sn hydroxide into
deposits.
Polarization curves (f) and (g) in Fig. 1 show the
effect of Au content in the solution. Solution SAu-6
(curve (f)) contains 7 g/L KAuCl4 while SAu-7 (curve
(g)) contains 10 g/L KAuCl4. A higher gold content (up
to 7 g/L) results in a little higher limiting current
density, which is beneficial for plating speed; however,
solution stability is worse. Where a longer solution
lifetime is preferred, a lower Au content is a better
choice.
Polarization curves (a) (c) in Fig. 2 show the effect
of ammonium citrate on the polarization behaviour of
the cathode in Sn solutions. Curve (a) was obtained
from solution SSn-1, which contains only Sn salt and
L-ascorbic acid. Its open circuit potential is much
higher than that those for curves (b) and (c) obtained
for solutions SSn-2 and SSn-3, which contain ammonium citrate, i.e. 100 and 200 g l 1, respectively. This
might imply that Sn ions form complexes with ammonium citrate, although no information about this is
available in the literature. The cathodic reaction for the
three solutions is expected to be

which contains no sulphite, while curves (b), (c) and (d)


are for SSn-4, SSn-5 and SSn-6 which contain 30, 60
and 100 g l 1 of sulphite, respectively. All the sulphite
containing solutions have two major waves in the Sn
plating region, which correspond to potentials higher
than 1.2 V versus SCE. There is a less distinct wave,
which occurs in the activation control region of the
second major wave. The first wave occurs at potentials
greater than 0.8 V versus SCE, while the second
wave occurs at potentials between 0.8 and 1.2 V
versus SCE. The second wave has a much higher limiting current density ( \ 3 mA cm 2) than the first wave
(B 1 mA cm 2). This means that there are two different cathodic reactions occurring over different potential
ranges as a result of the addition of sulphite. It is clear
from the figures that the addition of sulphite results in
a higher limiting current density and a higher cathodic
potential, which makes Sn plating easier. The limiting
current density is the largest for sulphite additions of 30
g l 1. If too much sulphite is added, e.g. 100 g l 1, the

Sn(II)+2 e=Sn(0)
where Sn(II) ions are in the form of simple Sn ions for
curve (a) and in the form of a Sn complex for curves (b)
and (c).
Hydrogen evolution for SSn-1 occurs at a potential
of : 0.65 V versus SCE, while for SSn-2 and SSn-3
the potential is moved to much more negative values of
: 1.10 V versus SCE. Ammonium citrate keeps the
pH of the solution at :6.5, which greatly suppresses
hydrogen production. Citrate concentration in the bath
does not significantly affect the limiting current density
or the open circuit potential.
Curves (a)(d) in Fig. 3 show the effect of sodium
sulphite on Sn plating. Curve (a) is for solution SSn-3,

Fig. 4. Cathodic polarization curves for solutions SAuSn (a), SAu-4


(b), SSn-5 (c). Curve (d) represents the sum of curves (b) and (c). The
dashed line indicates the approximate potential for hydrogen evolution for curve (c).

W. Sun, D.G. I6ey / Materials Science and Engineering B65 (1999) 111122
Table 3
Solutions utilized for polarization studies
Solution composition
Au solutions

SAu-1
SAu-2
SAu-3

SAu-4

SAu-5

SAu-6

SAu-7

Sn solutions

SSn-1
SSn-2

SSn-3

SSn-4

SSn-5

SSn-6

SSn-7

SSn-8

AuSn solution

SauSn

100 g l1 ammonium citrate


5 g l1 KAuCl4
200 g l1 ammonium citrate
5 g l1 KAuCl4
200 g l1 ammonium citrate
5 g l1 KAuCl4
30 g l1 Na2SO3
15 g l1 L-ascorbic acid
200 g l1 ammonium citrate
5 g l1 KAuCl4
60 g l1 Na2SO3
15 g l1 L-ascorbic acid
200 g l1 ammonium citrate
5 g l1 KAuCl4
100 g l1 Na2SO3
15 g l1 L-ascorbic acid
200 g l1 ammonium citrate
7 g l1 KAuCl4
60 g l1 Na2SO3
15 g l1 L-ascorbic acid
200 g l1 ammonium citrate
10 g l1 KAuCl4
60 g l1 Na2SO3
15 g l1 L-ascorbic acid
15 g l1 L-ascorbic acid
5 g l1 SnCl2H2O
100 g l1 ammonium citrate
15 g l1 L-ascorbic acid
5 g l1 SnCl2H2O
200 g l1 ammonium citrate
15 g l1 L-ascorbic acid
5 g l1 SnCl2H2O
200 g l1 ammonium citrate
30 g l1 Na2SO3
15 g l1 L-ascorbic acid
5 g l1 SnCl2H2O
200 g l1 ammonium citrate
60 g l1 Na2SO3
15 g l1 L-ascorbic acid
5 g l1 SnCl2H2O
200 g l1 ammonium citrate
100 g l1 Na2SO3
15 g l1 L-ascorbic acid
5 g l1 SnCl2H2O
200 g l1 ammonium citrate
60 g l1 Na2SO3
15 g l1 L-ascorbic acid
7 g l1 SnCl2H2O
200 g l1 ammonium citrate
60 g l1 Na2SO3
15 g l1 L-ascorbic acid
10 g l1 SnCl2H2O
200 g l1 ammonium citrate
5 g l1 KAuCl4
60 g l1 Na2SO3
15 g l1 L-ascorbic acid
5 g l1 SnCl2H2O

119

effect on limiting current density is not as great as that


for solutions containing less sulphite. The effect of
sulphite must be related to the interaction between
ammonium citrate and sulphite. When sulphite is
added, some NH3 gas is released from the solution
during mixing. Tin is plated from a complex formed
between Sn ions and NH3 in the solution. The addition
of sulphite to the solution lowers the concentration of
the Sn complexing agent (NH3), facilitating Sn plating.
Polarization curves for Solutions SSn-5, SSn-7 and
SSn-8, which contain different Sn contents (5, 7 and 10
g l 1, respectively), are shown in Fig. 3 (curves (c), (e)
and (f)). With more Sn salt in solution, a larger limiting
current density results.
From Fig. 3, it can be seen that for Sn plating, the
suitable potential range for plating is approximately
0.7 to 1.10 V versus SCE. Comparing this region
with the suitable range for Au plating (0.6 to 0.75
V versus SCE), the overlapping region is quite narrow,
which is disadvantageous for practical plating.
The polarization curve for a AuSn solution
(SAuSn), which is essentially a combination of Solutions SAu-4 and SSn-5 and also the most stable AuSn
solution, is given in Fig. 4 (curve (a)). The polarization
curves for Solutions SAu-4 and SSn-5 are also shown in
Fig. 4 (curves (b) and (c), respectively), while curve (d)
is the summation of curves (b) and (c). Curves (a) and
(d) are qualitatively quite similar. The implication is
that Au plating and Sn plating are somewhat, but not
completely independent plating processes. Four distinct
regions can be identified in curve (a), which are indicated by AB, BC, CD and DE. The first region, AB,
corresponds to the plating of nearly pure Au at potentials higher than : 0.68 V, since there is no significant current density for Sn plating in this potential
range for curve (c) (SSn-5). The steep edge BC from
: 1 to : 3 mA cm 2 is due to the incorporation of Sn
plating. It can be surmised that the Sn content in the
deposits will increase with increasing current density in
this current range. The third region, CD, corresponds
to a potential range of 0.68 to 0.8 V versus SCE
and a current density ranging from : 3.0 to 4.2 mA
cm 2. The total current density increases as the potential is lowered in curve (a). In this region, curve (c)
indicates that the current density for Sn plating increases initially, then remains constant and finally
jumps to a higher level again. In the same potential
range, the Au plating current density continues to
increase as shown in curve (b). It is difficult to predict
precisely how the deposit composition will change with
current density in this potential range; however, it can
be estimated that the deposit composition will remain
approximately constant as both Au and Sn plating
increase as the potential is lowered. In the region DE,
the current density increases rapidly with decreasing
potential, because of the onset of hydrogen evolution or
carbon plating. It is difficult to predict the deposit

120

W. Sun, D.G. I6ey / Materials Science and Engineering B65 (1999) 111122

composition change with increasing current density;


however, poor deposit morphology and high carbon

levels in the deposits can be expected. One can conclude


from Fig. 4 that a suitable working current density or
potential for AuSn plating would be the region BD
(: 14 mA cm 2), since both Au and Sn are plated in
this range.

3.3. Wafer causticity test


Within the measurement accuracy of the balance
utilized (9 0.0002 g), there was no weight loss for
either InP or GaAs wafers after soaking in Solution
SAuSn for 48 h. AFM images, before and after soaking, for GaAs are shown in Fig. 5. No pitting was
detected. Among the two types of wafers, only InP is
subject to attack by hydrochloric acid. Solution S6 is
only weakly acidic. Chlorine ions arise from the addition of a relatively low concentration of Au and Sn
salts. Furthermore, the chlorine ions act as complexing
ligands for Au and Sn ions, where their chemical activity is much lower than for simple chlorine ions.
During practical plating, the working potential for
the wafer will be in the 0.2 to 1.1 V versus SCE
range, which is under the condition of cathodic protection for the wafers. The plating time for eutectic Au
Sn solder for laser chip bonding is generally of the
order of 23 h, which is significantly less than the 48 h
soaking time utilized here.

3.4. Preliminary electroplating experiments

Fig. 5. AFM images of GaAs wafer surface prior to (a) and after (b)
immersion in solution SAuSn.

Fig. 6. Deposit Sn content versus current density for DC and PC


plating from solution SAuSn.

Solution SAuSn was utilized for preliminary plating


tests. The aim here was to demonstrate plating capability and not to analyze the deposits in detail or optimize
the plating process. A small amount of NiCl2 (1 g l 1)
was added as a leveler. Deposit composition results for
a range of current densities are shown in Fig. 6. The
current density range was chosen based on the polarization results. The composition versus current density
tendencies are similar for DC and PC plating, although
PC deposits have consistently higher Sn levels at the
same current density. The Sn content initially increases
with increasing average current density, levels out at
: 37 at.% Sn and then decreases with increasing current density. Tin plating appears to be favoured by a
more negative cathodic potential, which is achieved by
increasing the current density. If the current density is
too high, however, hydrogen evolution becomes significant, decreasing the efficiency of alloy plating. Hydrogen evolution may also cause a local increase in pH,
increasing the susceptibility of Sn ion complexing. Tin
ions will be further stabilized as a result of complex
formation, suppressing Sn plating and reducing tin
concentration in the deposit. The composition plateau
in Fig. 6 can likely be shifted by altering the solution
composition.

W. Sun, D.G. I6ey / Materials Science and Engineering B65 (1999) 111122

121

Fig. 7. SEM top view images for selected deposits from Fig. 6.

Representative microstructures for the deposition


conditions presented in Fig. 6 are shown in Fig. 7. The
deposits show a tendency towards coarser microstructures at higher current densities, particularly for DC
deposits.

4. Conclusions
A relatively stable Au Sn plating bath has been
developed. This is a non-cyanide, weakly acidic system
with ammonium citrate as a buffering agent for hydrogen ions and Sn ions, chloride-ammonium citratesodium sulphite as a mixed complexing agent for Au
and L-ascorbic acid as a Sn hydrolysis inhibitor. Bath
stability depends on the total gold and tin content, as
well as additives and the way the solution is prepared
and stored.
The effects of ammonium citrate, sulphite and Au/Sn
concentration on the polarization behaviour of the
cathode in various solutions have been studied. In Sn
solutions, the addition of ammonium citrate changes
the cathode open circuit potential and the hydrogen
evolution potential to more negative values. The ammonium citrate concentration has very little effect on
the polarization behaviour of both Au and Sn solutions. Addition of sodium sulphite to Au solutions
lowers the open circuit potential substantially, although
the concentration of sulphite has very little effect on the
polarization behaviour. For Sn solutions, the addition
of sulphite moves the open circuit potential to higher
values and increases the limiting current density,

thereby enhancing Sn plating. The shape of the Sn


polarization curves undergoes a significant change with
the addition of sulphite, from single wave to two wave
curves, which is believed to result from the interaction
between sulphite and ammonium citrate. Gold and Sn
plating from a common solution appear to be somewhat independent processes, with a current density of
:14 mA cm 2 being optimal.
Causticity tests showed that InP and GaAs were
essentially immune to corrosive effects from the electrolyte. Preliminary electroplating experiments using
the developed plating solution indicate that coelectroplating of uniform AuSn deposits is feasible.
Acknowledgements
This work was funded by grants from Nortel Technology and the Natural Sciences and Engineering Research Council (NSERC) of Canada.
References
[1] C. Kallmayer, D. Lin, J. Kloeser, H. Oppermann, E. Zakel, H.
Reichl, IEEE/CPMT International Electronics Manufacturing
Technology Symposium, 1995, pp. 20.
[2] C. Kallmayer, D. Lin, H. Oppermann, J. Kloeser, S. Weib, E.
Zakel, H. Reichl, Proceedings of the 10th European Microelectronics Conference, 1995, pp. 440.
[3] E. Zakel, H. Reichl, in: J. Lau (Ed.), Flip-Chip Technologies (ch.
15), McGraw-Hill, 1995, p. 415.
[4] G.G. Stanley, The Extractive Metallurgy of Gold in South
Africa, vol. 2, The South African Institute of Mining and
Metallurgy, 1987, p. 842.

122

W. Sun, D.G. I6ey / Materials Science and Engineering B65 (1999) 111122

[5] H. Honma, K. Hagiwara, J. Electrochem. Soc. 144 (1995)


3469.
[6] J. Traut, J. Wright, J. Williams, Plating Surf. Finish. 77 (1990)
49.
[7] A. Gemmler, W. Keller, H. Richter, K. Ruess, Plating Surf.
Finish. 81 (1994) 52.
[8] R.J. Morrissey, R.I. Cranston, US Patent 5,277,790 (1994).
[9] R.J. Morrissey, Plating Surf. Finish. 80 (1993) 75.
[10] T. Osaka, A. Kodera, T. Misato, T. Homma, Y. Okinada, O.
Yoshioka, J. Electrochem. Soc. 144 (1997) 3462.
[11] A. Meyer, S. Losi, F. Zuntini, Proc. Fachtagung. Galvanotachnik, Leipzig (1970), Swiss Patent 506,828 (1969).
[12] T. Inoue, S. Ando, H, Okudaira, J. Ushio, A. Tomizawa, H.
Takehara, T. Shimazaki, H. Yamamoto, H. Yokono, Proceedings of IEEE 45th Electronic Components and Technology
Conference, May 2124, 1995.
[13] M. Kato, Y. Yazawa, Y. Okinaka, Proceedings of the 82nd
AESF Technical Conference, American Electroplaters and Surface Finishers Society (1995) pp. 813.
[14] D.G. Foulke, in: F.H. Reid, W. Goldie (Eds.), Gold Plating
Technology (ch. 7), Electrochemical Publications, 1974, p. 52.
[15] A.C. Tan, Tin and Solder Plating in the Semiconductor Industry (chs. 8 10), Chapman and Hall, 1993, p. 197.
[16] D.R. Mason, A. Blair, P. Wilkinson, Trans. Inst. Met. Finish.
52 (1974) 143.
[17] E. Raub, K. Bihlmaier, Galvanische Weissgolniederschlage,

[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]

Mitt. Forschungsinst. Probierants. Edelmetalle Staatl. Hoheren


Fachschule Schwab. Gmund, 11 (1937) 59.
T. Frey, W. Hempel, DE 4406434 (1995).
W. Kuhn, W. Zilske, A.-G. Degussa, Ger. DE 4,406,434, August 10, 1995.
N. Kubota, T. Horikoshi, E. Sato, J. Met. Fin. Soc. Jpn. 34
(1983) 37.
Y. Tanabe, N. Hasegawa, M. Odaka, J. Met. Fin. Soc. Jpn.
34 (1983) 8.
N. Kubota, T. Horikoshi, E. Sato, Plating Surf. Finish. 71
(1984) 46.
M.J. Nicol, E. Schalch, Report No. 1848, National Institute
for Metallurgy, Johannesburg, South Africa, 1976.
M.J. Nicol, E. Schalch, Report No. 1844, National Institute
for Metallurgy, Johannesburg, South Africa, 1976.
S. Matsumoto, Y. Inomata, JP 61 15,992 [86 15.992], January
24, 1986.
W. Chang, K. Li, Concise Handbook of Analytical Chemistry,
Beijing University Publishing House, 1981, p. 121.
B. Oddo, Q. Mingoia, Gazz. Chim. Ital. 57 (1927) 820.
J. Socha, S. Safarzynski, T. Zak, J. Less-Common Met. 43
(1975) 283.
T.P. Smith, US Patent 3,057,789 (1962).
F. Zuntini, G. Aliprandini, M.J. Gioria, A. Meyer, S. Losi,
US Patent 3,787,463 (1974).
P. Laude, E. Marka, F. Zuntini, US Patent 4,192,723 (1980).

You might also like