Вы находитесь на странице: 1из 8

Technical Note

Rotating cylinder electrode study of the effect of activated


polyacrylamide on surface roughness of electrodeposited copper
C. Fabian
a,b,

, M.J. Ridd
b
, M. Sheehan
a
a
Department of Chemical Engineering, James Cook University, Townsville, QLD., 4811 Australia
b
Department of Chemistry, James Cook University, Townsville, QLD., 4811 Australia
Received 24 April 2006; received in revised form 1 August 2006; accepted 2 August 2006
Available online 12 September 2006
Abstract
The development of a new organic additive based on polyacrylamide is reported for copper electrowinning (EW) and
electrorefining (ER) in which a rotating cylinder electrode (RCE) is applied as a novel method of determining surface roughness.
The effect of hydrodynamics of the electrolyte (Cu
2+
=36 g/L; H
2
SO
4
=160 g/L and Cl

=25 mg/L) on the surface roughness of the


copper deposit was first validated against the literature. The effect of the preparation of polyacrylamide (PAM) on the surface
roughness was then assessed. PAM (MW 15 million Dalton) was dissolved at 50 C in full-strength and systematically half-
strength standard electrolyte and in water and alkaline solutions and dosed to the electrolyte prior to copper electrowinning for 6 h
using the RCE at 50 C. The lowest surface roughness (6.59 m0.49) was obtained when PAM was prepared in 16-fold diluted
electrolyte giving a block copolymer hydrolysis product named activated polyacrylamide (APAM) in this work.
2006 Elsevier B.V. All rights reserved.
Keywords: Copper electrowinning; Polyacrylamide; Rotating cylinder electrode; Surface roughness
1. Introduction
At commercial scale, copper electrorefining is
conducted at about 62 C2 to minimise anode
passivation and copper electrowinning between 40 C
and 55 C. The main process variables are current den-
sity, electrolyte fluid flow described by the diffusion
layer thickness (), temperature, deposition time,
chloride ions and organic additives. These process vari-
ables affect the topography (surface roughness) of the
copper deposit. Therefore, the measurement of the sur-
face roughness, followed by cross-section examination
and crystallite size determination, is the most suitable
response variables to control the copper deposit quality.
The Mt. Gordon operations produced 50,000 tonnes/
year of copper cathodes using a ferric ionacid leach of
chalcocite ore followed by solvent extraction and elec-
trowinning (Baxter et al., 1999; Richmond and Christie,
1999). Non-ionic polyacrylamide (Magnafloc 800HP
from Ciba, MW 15 million Dalton) was dosed as floc-
culant to the hydroclassifiers and pinned bed clarifiers
within this process. Richmond and Christie (1999) also
indicated that the pregnant leach solution contained about
(g/L): sulfuric acid=10; cupric ions=30; ferrous ions=30
and ferric ions=23 at about 50 C. It was considered
likely that occasionally polyacrylamide (PAM) or its
hydrolysis products passed through the solvent extraction
stage and then reached the electrowinning plant where it
Hydrometallurgy 84 (2006) 256263
www.elsevier.com/locate/hydromet

Corresponding author. Department of Chemical Engineering, James


CookUniversity, Townsville, QLD., 4811Australia. Tel.: +61297177067;
fax: +61 2 9543 7179.
E-mail addresses: cfabian2@bigpond.net.au,
ces.fabian@ansto.gov.au (C. Fabian).
0304-386X/$ - see front matter 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.hydromet.2006.08.001
could act as an organic additive. Thus, the electrolyte
potentially contained two organic additives: PAM and
Guarfloc66 (Guar), the industry-standard additive inten-
tionally used to smooth copper deposits. It was considered
that the combination of these organic additives occasion-
ally assisted Mt. Gordon to produce highly smooth copper
cathodes. However, it was unclear, whether this smooth-
ness was the effect of PAM by itself or its combination
with Guar. This Technical Note develops the use of a RCE
to measure surface roughness of electrodeposited copper
and assesses the best means of preparing or hydrolysing
PAM as an additive to minimise surface roughness. A
second publication will compare the effect of APAM and
Guar on surface roughness more comprehensively.
In the copper electrodeposition industry, it is im-
perative to dose very small concentrations of additives
including chloride ions to control the physical and che-
mical properties of the copper deposit. It is also widely
recognized in the industry that the adsorption of these
additives at electrode surfaces significantly influences the
current-potential relationships (Bonou et al., 2002; Kelly,
1999; Moffat et al., 2000, 2004; Onicio and Muresan,
1991; Wu and Barkey, 2000). These organic additives are
typically classified as levelling agents or suppressors and
grain refiners or accelerators. Levelling agents/suppres-
sors increase the over-potential relative to the additive
free-electrolyte by conferring preferential adsorption at
the protrusions in copper EW and ER. These inhibitors
control the horizontal growth on vertically aligned
electrodes to produce smooth deposits. Grain refiners/
accelerators decrease the over-potential for deposition by
inducing newnucleates on the surface. It is widely known
that chloride ions also depolarize the electrode (Gabrielli
et al., 2004; Ilgar and O'Keefe, 1997; Wu and Barkey,
2000) by enhancing the rate of charge-transfer, or the
formation of dendrites rather than new nuclei.
Under neutral conditions non-ionic PAM is stable;
but under acidic or basic conditions it undergoes hydro-
lysis. Moreover, the hydrolysis products under strong
and weakly acidic solutions; and alkaline solutions
differ (Caulfield et al., 2002). PAM is highly soluble in
water but its solution viscosity limits its concentration to
very low values. It has a unique position for industrial
applications due to its adsorption properties (Pefferkorn,
1999) but there is limited literature describing the
behaviour of PAM in strong acid solutions such as
copper electrolytes in 1.51.8 M sulfuric acid. Hydro-
lysis of non-ionic PAM, shown in Fig. 1 (left), results in
the transformation of the amide group (CONH
2
) of the
PAM molecules to produce carboxylate (COO

) and
ammonia (Ghannam, 2002). Hydrolysed PAM, shown in
Fig. 1 (right), is frequently referred to as anionic PAM
and consists of acrylamideacrylic acid copolymers.
Panzer et al. (1984) and Halverson et al. (1985)
investigated the structural changes of PAM macromole-
cule from weak acid (pH 2) and alkaline hydrolysis using
13
C NMR and derived a sequence length of carboxyl
groups in the PAM chain. These authors described
structural characteristics within the polymer chain, using
the notation A for the acrylamide group in the polymer
and B for the acrylic acid group. It was determined that
after 24 h at 110 C and pH 2, there was 57% hydrolysis
predominantly producing blocks of continuous acrylic
acid or acrylamide segments, i.e., AAA and BBB
functional groups. But under alkaline conditions there
was 56% hydrolysis with the products containing ABB,
BAB, ABA triads. The B segments from alkaline
hydrolysis were well distributed along the chain with an
average sequence length of 1.4; whilst the average
sequence length of carboxyl groups from pH 2 acid
hydrolysis was 14 (Panzer and Halverson, 1988).
In this work the RCE is used to electrowin copper
under typical electrowinning conditions. The surface
roughness of the electrowon copper is compared to that
of Ilgar and O'Keefe (1997) who used a parallel plate
electrode under similar hydrodynamic conditions. Thus,
a brief literature review on the hydrodynamics of the
RCE is discussed below.
A review of the literature on the electrodeposition of
metals shows that the primary and mass transfer-limited
Fig. 1. Non-ionic polyacrylamide (left) and anionic or hydrolysed polyacrylamide (right).
Table 1
Experimental conditions used to assess surface roughness
This work (Ilgar and O'Keefe, 1997)
Copper, g/L 36 36
Sulfuric acid, g/L 160 150
Chloride ions, mg/L 25 20
Active cathode area, cm
2
27.91 12
Electrolyte volume, L 3.75 0.2
Inter-electrode distance, mm 40 30
257 C. Fabian et al. / Hydrometallurgy 84 (2006) 256263
current distribution are uniform on a RCE (Grabowski
and Newman, 1993; Newman and Thomas-Alyea,
2004). The fluid flow at the RCE is described using
the Reynolds and Taylor numbers. However, the Rey-
nolds number (Re) has been expressed differently in the
literature due to variations in the selection of the radius,
diameter and/or inter-electrode distance as the charac-
teristic length for the RCE. The critical Reynolds num-
ber for the RCE is about 200 (Gabe, 1995) and therefore
turbulent flow is achieved at low speeds of rotation.
Moreover, the ohmic potential drop and concentration
change at the electrode can be accurately calculated
even in turbulent flow when the RCE is used (Newman
and Thomas-Alyea, 2004). In this work, the Re number
is defined as Re=d
1
2
/ 2v (Fabian et al., 2006) and vary
from about 80 to 280. The Taylor number, Ta=0.25
(d
2
d
1
)
1.5
d
1
0.5
/ according to Mizushina (1971)
varies from about 480 to about 1740 or Ta=[Re
2
.(r
2

r
1
) / r
o
] according to Newman (1991) varies from1.47.10
6
to 22.10
6
. Therefore the qualitative interpretation of these
results indicates that the hydrodynamic regime for the
RCE in the present work is laminar with vortices.
If the cathodic current density is much less than the
limiting mass transfer controlled current density, the
Fig. 2. Rotating cylinder electrode
1
N
bottom section.
258 C. Fabian et al. / Hydrometallurgy 84 (2006) 256263
limiting diffusion current density, i
L
is given by Fick's
first law, Eq. (1), with a surface concentration of zero
(Manzanares and Kontturi, 2003).
i
L
u
nFDC
b
d
1
where , cm, is the diffusion layer thickness and other
terms have their usual meaning.
Mass transfer between concentric cylinders, the inner
of which is rotating with an angular speed has been
studied extensively (Barkey et al., 1989; Eisenberg
et al., 1954; Newman and Thomas-Alyea, 2004). The
limiting current density equation developed by Eisen-
berg et al. (1954) was slightly modified by Arvia and
Carrozza (1962) as described in Eq. (2).
i
L
0:0791nFC
b
d
i
v

0:30
U
0:70
d
o
d
i

Sc
0:644
2
where
i
L
is the limiting current density,
mA/cm
2
,
Sc is dimensionless Schmidt
number (/ d)
C
b
is Cu concentration in bulk
solution mol/cm
3
U is the peripheral velocity,
m/s
d
i
is the diameter of the inner
(rotating) RCE, m
d
h
is the hydraulic diameter
(d
o
d
i
)
d
o
outer diameter (stationary), m
If i
L
is determined experimentally, Eqs. (1) and (2)
could be used to calculate the diffusion layer thickness,
, provided the physicochemical properties of the elec-
trolyte are known.
2. Experimental
Table 1 summarises the standard operating condi-
tions using the RCE. The PAM used was non-ionic with
a high MW (MW=15 million Dalton Ciba Magnafloc
800 HP). The concentration of PAM in the preparation
media was 1 mg/mL which was gently stirred at 50 C to
effect dispersion. An Eppendorf pipette was used to
dose 3.75 mL of the PAM solution product in 3.75 L of
electrolyte held at constant temperature in another water
bath to achieve a final concentration of 1 mg/L.
A RCE was locally constructed and is similar in
design to that described by Barkey et al. (1989). Fig. 2
shows the design of the bottom sections of the RCE. The
RCE consisted of a 4.45 cm diameter 15 cm overall
height 316 L stainless steel rod (Sandvik). This electrode
was bounded tightly with 3 mm thickness PTFE sleeves
to suppress edge effects and to leave a 2 cm active height
of stainless steel exposed or an active area of 27.91 cm
2
.
A dimensionally stable anode (DSA) sourced from
Eltech Systems Corporation (USA) was used for the
anode of the RCE. The RCEwith the DSAconcentrically
arranged around it was located within a 5 L Pyrex beaker.
Eight PTFE baffles (200205 mm) within the beaker
maintained the DSA mesh at a distance of 40 Mm from
the RCE surface through a PVC lid. The entire RCE EW
cell rested in a thermostatted water bath. The rate of
rotation of the RCE was controlled using a Movitrac
controller and 0.37 kW motor (RF27DT71D4) using a
gear box sourced from SEW Eurodrive. Direct current to
the RCE EW cell was applied using a 0-18VDC, 020 A
digital power supply sourced from Cole-Parmer Extech
Equipment Pty. Ltd, USA.
The RCE was polished with 1 m and 0.25 m
diamond paste and had an initial surface roughness of
0.150.04 m. Once the copper was detached from the
RCE, the RCEwas immersed in 10%nitric acid for about
60 s, polished with 0.25 m diamond paste, washed with
distilled water, immersed again in the nitric acid solution
Table 2
Effect of diffusion layer thickness on surface roughness
EW cell type (Ilgar and O'Keefe, 1997) This study
Parallel plate electrode RCE RCE
RCE rpm 25 10
Reynolds number 3082 1233
Diffusion L.
thickness, , m
65 175
a
87 107
EW time, h 5.5 or 4 5.5 or 4 4 4
Current density,
mA/cm
2
25 and 35 25 and 35 30 30
Electrolyte
temperature, C
30 and 40 30 and 40 50 50
Coulombs, C/cm
2
500 500 434 434
Number of
fine needles
N.A.
b
N.A.
b
None 50
Surface roughness,
Ra, m
3.9
0.50
6.40.50 &
12.50.50
4.31
0.16
5.51
0.41
a
Natural convection (stationary electrolyte).
b
Not reported.
Table 3
Experimental conditions used to assess the effect of polyacrylamide
preparation on surface roughness
Copper, g/L 36
Sulfuric acid, g/L 160
Chloride ions, mg/L 25
RCE active Deposition area, cm
2
27.91
Electrolyte volume, L 3.75
PAM/APAM conc. in electrolyte, mg/L 1
Electrowinning time, h 6
Current density, A/m
2
300
Electrolyte temperature, C 50
RCE , m (10 rpm) 107
259 C. Fabian et al. / Hydrometallurgy 84 (2006) 256263
for another 60 s and completely washed with distilled
water. This sequence was repeated after each test and the
stainless steel maintained its brightness.
2.1. Surface roughness measurements
The surface roughness was collected using a Mahr
Perthometer M1 with a PFM Drive Unit, (Mahr GMBH,
Germany) using a 2 m stylus tip radius. It was calib-
rated with its PGN-3 (Ra=0.79 m) standard roughness
profile. The surface roughness measurement was con-
ducted before the copper deposit was detached from the
RCE along the height of the electrode. The RCE was
placed horizontally on a V-block standard tool and a
bull's-eye level was also placed on top of the PFM drive
unit to control its horizontality. This measurement pro-
cess was repeated 8 times around the cylinder electrode
on equi-distantly distributed lines.
3. Results and discussion
3.1. Effect of fluid flow at the RCE on surface roughness
Preliminary tests were conducted to evaluate the
RCE and surface roughness measurement by comparing
results produced with those of Ilgar and O'Keefe (1997).
The electrolyte compositions used in both studies are
similar (Table 2), except for minor differences between
the sulfuric acid and chloride ion concentrations, and
closely resembles that used in commercial practice.
Table 4
Effect of polyacrylamide preparation on surface roughness
Test No. 1 2 3 4 5 6 7 8
Sulfuric acid conc., g/L Alkaline Water 10 20 40 80 160
No. of replicates 1 1 3 4 1 2 1 2
pH Preparation media 8.5 8.5 6 1.5 1.25 1.14 1 b1
APAM preparation temp., C 50 50 25 50 50 50 50 50
APAM preparation time, h 3 2 2 2 2 2 2 2
Electrode voltage drop, V 1.46 1.42 1.45 1.35 1.4 1.38 1.36 1.33
Surface roughness, Ra m 8.47 7.03 7.09 6.59 7.11 7.09 7.29 7.32
Ra standard deviation, m 0.93 0.43 0.44 0.47 0.41 0.85 0.49 0.61
Test 4 one-way ANOVA sig.
a
0.00 0.54 0.04 0.32 0.11 0.06 0.00
Dendritesb0.1 mm length 10 Nil Nil Nil Nil Nil Nil Nil
S. steel spotsb11 mm Nil Nil Nil Nil Nil 1 Nil Nil
a
One-way ANOVA significance for Test 4, Tuckey, HSD.
Fig. 3. Plot of the effect of PAM preparation media on surface roughness of the copper deposits.
260 C. Fabian et al. / Hydrometallurgy 84 (2006) 256263
Table 3 presents the results of the effect of the RCE
rotation rate on surface roughness compared with those
obtained by Ilgar and O'Keefe (1997) in the absence of
organic additives. The values of diffusion layer thick-
ness were obtained using Eqs. (1) and (2), and using the
experimental current density and physicochemical
properties of the electrolyte presented elsewhere (Fabi-
an, 2005; Fabian et al., 2006). It can be seen in Table 3
that the measurement of average surface roughness
obtained in this testwork is consistent with those results
obtained by Ilgar and O'Keefe (1997). However, the
value for of 175 m at 30 C and 40 C for natural
convection (stationary electrolyte) reported by Ilgar and
O'Keefe, is about twice as large as the of 97 m at
45 C and 115 m at 65 C obtained for natural con-
vection in this study which agree more closely with the
value of 78 m used by Drews et al. (2003) for natural
convection at 25 C. The comparability of the surface
roughness and the trend of the surface roughness towards
higher surface roughness at higher in both studies lead
us to conclude that the results of the RCE are comparable
to those that are seen on parallel plate electrodes.
3.2. The effect of polyacrylamide hydrolysis conditions
on surface roughness
Given the hydrolysis of PAM can lead to a range of
hydrolysis products and adsorption characteristics,
control of the preparation of PAM (and its associated
hydrolysis products) is important to achieving a smooth
copper deposit surface. Thus, PAM was prepared in both
acid solutions (including diluted solutions of the full-
strength electrolyte) and alkaline solutions. Experiments
were designed to evaluate the effect of PAM preparation
media on the surface roughness of the copper deposit
obtained from copper EW. Table 4 describes the results
of the effect of the preparation media of polyacrylamide
on the surface roughness of EW copper.
The one-way ANOVA between groups was analysed
using SPSS software (version 11, 2003). Fig. 3 shows
the plot of the surface roughness of the copper deposits
produced as a function of preparation media. The error
bars show the 95% confidence interval (CI) obtained
from at least 8 replicate measurements of each test sam-
ple, sometimes repeated 2, 3 or 4 times.
The PAM used in Test 4 was prepared with 16-fold
diluted electrolyte (pH=.5) and produced a surface with
a significantly lower roughness (6.590.47 m) than
PAM prepared in water, full-strength electrolyte and
alkaline hydrolysis. This PAM is referred to as APAM
or activated polyacrylamide. The significance of the
difference in surface roughness between each test and
Test 4 is also presented in Table 4. In acid solutions
(Tests 4 to 8), the mean surface roughness increases as
the concentration of sulfuric acid increases indicating an
effect of acid concentration on the hydrolysis and ad-
sorption of PAM. Since surface roughness is directly
related to adsorption, it is assumed that APAM confers
the highest adsorption, although further test work would
be necessary to confirm this hypothesis. However, the
C
13
NMR spectra (Fabian, 2005) indicates that the pro-
portion of PAM hydrolysed is less than 10%.
PAM prepared at a pH of 8.5 at 50 C for 3 h (Test 1)
produced the highest surface roughness (8.470.93 m)
of any test. It appears that under alkaline conditions sur-
face roughness increases with preparation time as PAM
hydrolysis increases. As discussed above, hydrolysis in
acid and alkaline solutions produce random copolymer.
Fig. 4. Effect of PAM degradation in 16-fold diluted electrolyte and
full-strength electrolyte on surface roughness.
Table 5
Effect of APAM degradation in 16-fold DE and full-strength
electrolyte
Test No. 9 10 11
pH Preparation media 1.5 1.5 1.5
Preparation temperature, C 50 25 50
Residence time in 16-fold DE, h 2 24 24
Residence time in electrolyte, h 0 0 24
Sulfuric acid concentration, g/L 10 10 10 and 160
APAM conc. in electrolyte, mg/L 1 1 1
EW time, h 6 6 6
Electrode voltage drop, V 1.34 1.39 1.41
Current density, mA/cm
2
30 30 30
Electrolyte temp., C 50 50 50
Diffusion layer thickness, m (10 rpm) 107 107 107
Surface roughness, Ra, m 6.23 6.89 7.46
Ra standard deviation, m 0.41 0.68 0.86
No. peaks per cm 82.63 78.13 77.75
Peaks per cm standard deviation 10.6 9.61 9.90
Dendrites formation Nil Nil Nil
261 C. Fabian et al. / Hydrometallurgy 84 (2006) 256263
3.3. Affect of ageing activated polyacrylamide on
surface roughness
Tests were undertaken to determine the effect of
ageing PAM in 16-fold diluted electrolyte (DE) and in
full-strength electrolyte. The study in full-strength elec-
trolyte was designed to simulate the effect of in-situ PAM
hydrolysis in the EWcells. Tests were also undertaken to
verify the effect of ageing of PAM in 16-fold DE at room
temperature for over 24 h to determine if APAMcould be
stored prior to use. This test was also designed to verify
whether additional ageing in the electrolyte at 50 C
following the 24 h in 16-fold DE could reduce surface
roughness. Therefore, Test 9 was run with fresh APAM
whilst Test 10 was run with PAM aged for 24 h at 25 C
and Test 11 was run after a further 24 h at 50 C without
adding any fresh APAM. Table 5 and Fig. 4 show the
experimental details and results for this testwork.
APAM prepared at 50 C for 3 h showed the lowest
surface roughness followed by APAM prepared at 25 C
for 24 h. The surface roughness produced from Test 3
was the highest and indicated that additional hydrolysis
in the electrolyte at 50 C for 24 h increased surface
roughness presumably through the formation of less
adsorbent copolymers, such as imides. Tests 1 and 2
were not statistically different indicating that APAM
could be prepared for up to 24 h without affecting
greatly the surface roughness. However, based on the
data shown in Fig. 4, it is recommended to prepare fresh
APAM every 6 or 12 h for continuous dosing.
4. Conclusions
It has been shown that the surface roughness obtained
with the RCE built for this testwork agrees with the results
obtained by Ilgar and O'Keefe (1997) and that the effect of
organic additives on surface roughness can be accurately
quantified using the RCE. It is clear fromthe literature that
hydrolysed polyacrylamide can be obtained by dissolving
polyacrylamide in both alkaline and acid solutions. How-
ever, the reaction products of polyacrylamide hydrolysed
in acid solution are different from those obtained from
alkaline solutions. It was shown that when high MWPAM
(15 million Dalton) was prepared in 16-fold diluted
electrolyte (pH=1.5) at 50 C for 2- h the surface
roughness of the copper deposit was statistically lower
than whenPAMwas either prepared in water, full-strength
electrolyte or alkaline solution at pH 8.5. We conclude
that the reason for the differences in surface roughness is
due to the differences in the PAM copolymers formed
during hydrolysis. Our data suggests that the formation of
a block copolymer of PAM/PAA in weakly acidic media
results in better adsorption of the polymer at the copper
metal/electrolyte interface to produce a smoother copper
deposit than that which occurs when a randomcopolymer
is formed under alkaline conditions.
Acknowledgements
We wish to thank Mount Gordon Operations of West-
ern Metals Copper Ltd. and to the Australian Research
Council for funding this study.
References
Arvia, A.J., Carrozza, J.S.W., 1962. Mass transfer in the electrolysis of
CuSO
4
H
2
SO
4
in aqueous solutions under limiting current and
forced convection employing a cylindrical cell with rotating elec-
trodes. Electrochim. Acta 7, 6578.
Barkey, D., Muller, R., Tobias, C., 1989. Roughness development in
metal electrodeposition I. Experimental results. J. Electrochem.
Soc. 138 (8), 21992207.
Baxter, K., Kaiser, C., Richmond, G.D., 1999. Design of the Mt.
Gordon Chalcocite Project, ALTA Copper Sulfides Symposium.
ALTA Metallurgical Services Ltd, Melbourne. 21 pp.
Bonou, L., Eyraud, M., Denoyel, R., Massiani, Y., 2002. Influence of
additives on Cu electrodeposition mechanisms in acid solution:
direct current study supported by non-electrochemical measure-
ments. Electrochim. Acta 47 (26), 41394148.
Caulfield, M.J., Qiao, G.G., Solomon, D.H., 2002. Some aspects of the
properties and degradation of polyacrylamides. Chem. Rev. 102
(9), 30673083.
Drews, T., Ganley, J., Alkire, R., 2003. Evolution of surface roughness
during copper electrodeposition in the presence of additives.
J. Electrochem. Soc. 150 (5), C325C334.
Eisenberg, M., Tobias, C., Wilke, C., 1954. Ionic mass transfer and
concentration polarization at rotating electrodes. J. Electrochem.
Soc. 101 (6), 306319.
Fabian, C., 2005. Copper Electrodeposition in the Presence of Guar or
Activated Polyacrylamide. PhD Thesis, James Cook University,
Australia.
Fabian, C., Mandin, P., Ridd, M.J., Sheehan, M., 2006. Hydrodynamic
modeling of copper electrodeposition at a vertical rotating cylinder
electrode. In: Woods, R., Doyle, F.M., Kelsall, G.H. (Eds.),
Electrochemistry in Mineral and Metal Processing VII. The
Electrochemical Society Inc., Denver, CO.
Gabe, D., 1995. Rotating electrodes for use in electrodeposition
process control. Plating Surf. Finish. 9, 6976.
Gabrielli, C., Mocoteguy, P., Perrot, H., Wiart, R., 2004. Mechanism of
copper deposition in a sulfate bath containing chlorides. J. Elec-
troanal. Chem. 572 (2), 367375.
Ghannam, M., 2002. Wetting behavior of aqueous solutions of
polyacrylamide over polyethylene substrate. J. Chem. Eng. Data
47 (2), 274277.
Grabowski, A., Newman, J., 1993. Current and potential distributions
on a cylinder electrode. J. Electrochem. Soc. 140 (6), 16251631.
Halverson, F., Lancaster, J., O'Connor, M., 1985. Sequence
distribution of carboxyl groups in hydrolyzed polyacrylamide.
Macromolecules 18 (6), 11391144.
Ilgar, E., O'Keefe, T., 1997. Surface roughening of electrowon
copper in the presence of chloride ions. In: Dreisinger, D. (Ed.),
262 C. Fabian et al. / Hydrometallurgy 84 (2006) 256263
Aqueous Electrotechnologies: Progress in Theory and Prac-
tice. The Minerals Metals and Materials Society, Warrendale,
pp. 5162.
Kelly, J., 1999. Copper Deposition in the Presence of Mixed
Surfactants. PhD Thesis, Columbia University, 135 pp.
Manzanares, J.A., Kontturi, K., 2003. Diffusion and migration. In:
Calvo, E.J. (Ed.), Interfacial Kinetics and Mass Transport. Wiley-
VCH, pp. 81121.
Mizushina, T., 1971. The electrochemical method in transport
phenomena. Adv. Heat Transf. 7 (2), 87161.
Moffat, T., et al., 2000. Superconformal electrodeposition of copper in
50090 nm features. J. Electrochem. Soc. 147 (12), 45244535.
Moffat, T.P., Wheeler, D., Josell, D., 2004. Electrodeposition of copper
in the SPS-PEG-Cl additive system. J. Electrochem. Soc. 151 (4),
C262C271.
Newman, J., 1991. Electrochemical Systems. Prentice-Hall Interna-
tional, London.
Newman, J., Thomas-Alyea, K.E., 2004. Electrochemical Systems.
John Wiley & Sons Inc., Hoboken, New Jersey.
Onicio, L., Muresan, L., 1991. Some fundamental aspects of levelling
and brightening in metal electrodeposition. J. Appl. Electrochem.
21, 565574.
Panzer, H., Halverson, F., 1988. Blockiness in hydrolyzed polyacryl-
amide. In: Moudgil, B., Scheiner, B. (Eds.), Flocculation
Dewatering, Proc. Eng. Found. Conf., Palm Coast Florida, USA,
pp. 239249.
Panzer, H., Halverson, F., Lancaster, J., 1984. Carboxyl sequence
distribution in hydrolyzed polyacrylamide. Polym. Mater. Sci.
Eng. 51, 268271.
Pefferkorn, E., 1999. Polyacrylamide at solid/liquid interfaces. J. Colloid
Interface Sci. 216, 197220.
Richmond, G., Christie, M., 1999. The commissioning and operation
of a copper sulfide pressure oxidation leach process at Mt. Gordon.
ALTA Copper Sulfides Symposium. ALTA Metallurgical Services
Ltd., Melbourne. 9 pp.
Wu, Q., Barkey, D., 2000. Faceting and roughening transitions on
copper single crystals in acid sulfate plating baths with chloride.
J. Electrochem. Soc. 147 (3), 10381045.
263 C. Fabian et al. / Hydrometallurgy 84 (2006) 256263

Вам также может понравиться