Sunteți pe pagina 1din 6

Surface & Coatings Technology 204 (2009) 353358

Contents lists available at ScienceDirect

Surface & Coatings Technology


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / s u r f c o a t

The effect of saccharin addition and bath temperature on the grain size of
nanocrystalline nickel coatings
A.M. Rashidi a,b, A. Amadeh b,
a
b

School of Engineering, Razi University, P.O. Box 67149-67346, Kermanshah, Iran


School of Metallurgy & Materials Engineering, University College of Engineering, University of Tehran, P.O. Box 11155-4563, Tehran, Iran

a r t i c l e

i n f o

Article history:
Received 8 June 2008
Accepted in revised form 28 July 2009
Available online 6 August 2009
Keywords:
Bath temperature
Electroplating
Grain size
Nanocrystalline nickel
Saccharin
Theoretical model

a b s t r a c t
Nanocrystalline nickel coating was synthesized by direct current electrodeposition from a Watts bath at the
current density of 100 mA/cm2 and pH = 4. The effect of saccharin addition (010 g/l) and bath temperature
(4565 C) on the average grain size of the deposits was investigated by XRD technique. The results showed
that the average grain size decreased from 426 nm to 25 nm as the saccharin concentration increased from 0
to 3 g/l, while further increase in saccharin concentration had no signicant effect. Theoretical model also
indicated a non-linear function for dependence of grain size on saccharin concentration, which was in
accordance with experimental results. The experimental results showed that the increases in the bath
temperature had no considerable effect on the average grain size of the deposits. A theoretical formula was
also established for the temperature dependence of the grain size.
2009 Elsevier B.V. All rights reserved.

1. Introduction
Electrodeposition has received considerable attention in recent
years as a feasible and economically viable technique for producing
nanocrystalline coatings [15]. This process is a powerful method for
fabrication of many highly precise products and synthesizing metallic
nanomaterials with controlled shape and size [6]. However, the
performance of electrodeposition for application of nanocrystalline
coatings is actually related to electroplating conditions and the
composition of plating bath. It has been demonstrated [710] that
under certain suitable conditions, only electrodeposition can, indeed,
yield the production of nanocrystalline nickel coatings. So, importance
has been gradually attached to study in this eld in recent years.
The impressibility of microstructure of nickel electrodeposits from
electrodeposition process parameters such as electrolyte type [9],
deposition technique (direct or pulse plating) [1114], pulse parameters [2,3,7,1417] and types of waveform in pulse current
electroforming [18], applied current density in direct current electroplating [5,911,19,20], pH of electrolyte [21,22], substrate conditions [9,23], addition of micro and nano-sized particles to the bath
[2436], emulsion electroplating with dense CO2 [37], the presence,
concentration, and combined effect of additives such as saccharin, 2-

Corresponding author. Tel.: +98 21 66493046; fax: +98 21 66480290.


E-mail addresses: rashidi1347@razi.ac.ir (A.M. Rashidi), amadeh@ut.ac.ir
(A. Amadeh).
0257-8972/$ see front matter 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.surfcoat.2009.07.036

butyne-1,4-diol, etc. [8,10,12,15,38,39], and the rotation speed of


cylindrical electrode [40] have already been investigated. Nevertheless, the results of electroplating experiments obtained by different
researchers are difcult to optimize certain conditions for production
nanocrystalline nickel coating, because in some cases, the reported
data are inconsistent or even different. For instance, Pin-Qiang et al.
[10] found that the grain size of nickel deposits decreased from 50 nm
to about 20 nm by increasing the current density from 50 mA/cm2 to
100 mA/cm2 while the effect of current densities higher than
100 mA/cm2 was negligible. Unlike to these results, a continuous
increasing in the grain size versus current density has also been
recognized in direct current electrodeposition of nickel coating
[5,9,11,17]. On the other hand, Aruna et al. [19] reported that the
current density had no signicant effect on the grain size of nickel
electrodeposits.
A review of literatures shows that saccharin has often been added
to nickel plating bath from the 80s [41] in order to improve the
ductility, brightness, and at later periods as a grain rener agent. The
grain rening by increasing the saccharin concentration in Ni plating
bath was investigated by Erb et al. [8], Pin-Qiang et al. [10] and Xuetao
et al. [16]. Although the addition of saccharin decreased the grain size
of the coating, but the saccharin content corresponding to leveling off
in grain size reduction with increasing the saccharin concentration is
different in these references. Furthermore, in earlier studies
[8,10,15,16] there is still a lack of simple and general mathematical
equations for dependence of the grain size of deposit on saccharin
concentration. The quantitative relationships can assist in reducing

354

A.M. Rashidi, A. Amadeh / Surface & Coatings Technology 204 (2009) 353358

the number of experiments required for optimization the process


parameters. It is therefore important to establish the mathematical
models for the inuence of electrodeposition process parameters on
the grain size of deposits and ascertain their certitude with respect to
experimental results. In the present paper, not only the experimental
results of the inuence of saccharin concentration on the grain size of
nanocrystalline nickel coatings was presented, but also a theoretical
model was developed to predict the variation of the grain size of the
coating as a function of saccharin concentration. The proposed model
can also be applied for other additives used in electroplating of other
metallic coatings than nickel.
Another problem in optimization of electrodeposition conditions
of nanocrystalline nickel coating is the effect of bath temperature,
because the plating temperature can play an important role on the
grain size of the coatings as demonstrated by Natter and Hempelmann
for nanocrystalline gold, copper and aluminum deposits [4244]. It is
well known that electrochemical processes such as electroplating are
inuenced by different diffusion processes: the diffusion of metal ions
in the electrolyte, the movement of ad-atoms on the electrode surface
and also the mobility of grain reners [35]. All of these processes,
therefore, depend on the bath temperature and should inuence the
structure of the deposits. It must be noted that in some cases, a
deviation more than 5 C from optimum temperature is sufcient to
harm the plating quality, deposition rate, and other properties of the
coating [45]. Although, only a few published papers deal with the
effect of bath temperature on nickel deposition [4653], most often
this factor have been investigated with respect to the texture, current
efciency, polarization behavior, electrochemical mechanisms, the
consumption rate of additives and/or microhardness. Unfortunately,
these studies have not been focused on the grain size and do not
present enough information about the relationship between the bath
temperature and the grain size of nickel deposits. This article presents
both the experimental data as well as a theoretical model of the
inuence of bath temperature on the grain size of nanocrystalline
nickel deposits.

2. Experimental procedure
Nanocrystalline nickel coatings were deposited from a Watts baths
by direct current (DC) electroplating at the current density of
100 mA/cm2. The basic composition of the bath was 300 g/l nickel
sulfate (Ni2SO46H2O), 30 g/l nickel chloride (NiCl26H2O), 30 g/
l boric acid (H3BO3). The pH of the bath was adjusted to 4.0 0.2 by
addition of drops of HCl (1 N) or NaOH (1 N). A nickel sheet of 99.99%
purity with dimensions of 100 mm 50 mm 5 mm was used as
anode and pure annealed copper plate with dimensions of
20 mm 15 mm 2 mm as cathode (substrate) materials. Prior to
deposition, the copper substrates were mechanically polished with
silicon carbide papers of 400, 600, 800, 1200 grits and alumina
suspensions of 8, 1 and 0.25 m, then rinsed with distilled water and
activated in 10% H2SO4 solution at room temperature for 30 s. The
deposition time was adjusted to achieve an average thickness of
100 m based on the Faraday's law. The experiments were carried out
in the baths containing different amounts (010 g/l) of sodium
saccharin as grain rener and stress reliever agent at 55 C and/or in
the baths with 5 g/l saccharin content at various bath temperatures
(45 C65 C).
The microstructure of the coatings was studied by scanning
electron microscope (SEM), transmission electron microscope (TEM),
and X-ray diffraction (XRD). XRD investigations were carried out
using a Philips X'Pert-Pro instrument operated at 40 kV and 30 mA
with CoK radiation ( = 1.789 ) at a scan rate of 0.05 s 1 in the
range of 40130 and 0.02 step size. The average grain size of
nanocrystalline nickel coatings was calculated from XRD patterns
using modied WilliamsonHall relation expressed in [54,55]. An

annealed nickel sample with an average grain size of 30 m was also


used as reference sample.
3. Results and discussion
Fig. 1 shows the XRD patterns of nickel coating produced at the
current density of 100 mA/cm2, bath temperature of 55 C and pH = 4
from the bath containing 5 g/l saccharin. For comparison, the XRD
pattern of reference sample (annealed nickel) has also been shown in
Fig. 1. It can be observed that the crystal structure of the coating is
pure fcc nickel and no characteristic peaks of other phases have been
recorded.
3.1. Effect of saccharin concentration
3.1.1. Experimental results
The XRD pattern of (111) and (222) reecting planes of nickel
coatings deposited from Watts baths containing different concentration of sodium saccharin (Cs) were presented in Fig. 2. The peak
broadening by increasing the amount of sodium saccharin is evident
in this gure indicating the grain rening.
From the peak broadening of XRD patterns, by means of a modied
WilliamsonHall relation tting procedure, the values of the mean
grain size as well as the standard deviation (sigma) have been
determined for each sample. The evolution of the grain size of
nanocrystalline nickel coatings versus saccharin concentration in
electroplating bath has been depicted in Fig. 3. It shows that a small
amount of saccharin has a considerable effect on the grain rening of
nickel deposits. The maximum grain rening has almost been occurred
at 3 g/l saccharin as a reduction in grain size by a factor more than 16. It
is well known that the grain size of deposits is strongly inuenced by
the presence of additive in electroplating bath [56] and the ability of
different additives in grain rening of nickel deposits have also been
reported by several researchers [8,10,15,16,39,57].
Numerous mechanisms have been suggested regarding the
inuence of additives on the grain size of electrodeposition coatings,
discussed in detail in a comprehensive review by Franklin [58] and also
by Oniciu and Muresan [59]. In brief, the important role of an additive
as a grain rener is its effect on (i) blocking the surface by formation of
complex compounds which increases the frequency of nucleation and
decreases the surface diffusion of nickel ions adsorbed on cathode
surface, and hence retards the crystalline growth [8,15,6062],
(ii) hydrogen evolution and/or absorption [63,64], and (iii) change

Fig. 1. XRD patterns of annealed (reference) sample and nanocrystalline nickel


coatings electrodeposited at i = 100 mA/cm2, T = 55 C, pH = 4 from a bath containing 5 g/l saccharin.

A.M. Rashidi, A. Amadeh / Surface & Coatings Technology 204 (2009) 353358

355

Fig. 2. Peak broadening of (111) and (200) reections as a function of saccharin content
for nanocrystalline nickel coatings electrodeposited at i = 100 mA/cm2, T = 55 C,
pH = 4.

Fig. 3. Comparison of experimental data and theoretical model for variation of grain size
of nanocrystalline nickel deposits as a function of saccharin concentration.

in cathodic overpotential1 [64,65]. The effect of additives is often


manifested by changes in polarization characteristics of cathode. The
currentpotential curves for Watts-type bath presented by Nakamura
et al. [65] indicated that the addition of 5 mM saccharin to the
electrolyte shifted the reduction potential of nickel ions to higher
potential by about 100 mV due to its inhibition effect on the reduction
of nickel ions. The higher overpotentials result in higher nucleation
rate, and thus in grain renement [66]. Hence, it can be presumed that
the change in overpotential caused by saccharin addition, accelerates
the nucleation and reduces the grain size of deposits.
For comparison, the data reported by Erb et al. [8,15] has also been
presented in Fig. 3. As it can be observed, the grain size of the coatings
at a given saccharin concentration in the present investigation is
greater than that reported in Refs. [8,15]. This difference arises from
difference in operating conditions. Erb et al. [8,15] used pulse current
electrodeposition to fabrication nanocrystalline nickel deposits from
Watts bath at 65 C and pH = 2. In pulse electroplating in comparison
to direct current plating, it is possible to apply higher current densities
which is accompanied by higher overpotentials resulting in higher
nucleation rate, and thus in grain renement [67]. Fig. 3 also indicates
that the saccharin concentration corresponding to the initiation of
curve plateau is about 3 g/l. This value conrms the values reported in
Refs. [10,16] but is different from that reported by El-Sherik et al. [8].
As seen in Fig. 3, beyond 3 g/l of saccharin, the grain size of the coating
is approximately independent of saccharin concentration. This
phenomenon can be attributed to the leveling off in the overpotential
[68] and/or saturation of adsorption sites on the cathode surface [69]
with increasing the saccharin amount in the bath.

In the absence of additive, the reduction reaction can occur in all


available active sites, but in the presence of additive, some of the
active sites are occupied and blocked. The occupied surface area (Aoc)
can be expressed in terms of surface coverage () as [45]:
1

Aoc = A

where A is the total surface area of the cathode. Therefore, unoccupied


surface area (Auc) which is free for reduction reaction is equal to:
Auc = AAoc = A1

If the applied current be denoted by I, then the amount of true


current density (i) of deposition is given as:

i =

I
Auc

By introducing Eq. (2) into Eq. (3), it can be obtained:

i =

i
1

in which i = AI is the current density in the absence of additive in the


bath. The surface coverage () is also related to the bulk solution
concentration of additive (Cs) in a general form of [45]:
=

k Csm
1 + k Csm

where k and m are constants. Thus, Eq. (4) can be rewritten as:
3.1.2. Theoretical approach
A simple model of the relationship between saccharin concentration and average grain size of nickel deposits can be derived by
considering the surface blocking (adsorption) mechanism, Langmuir
isotherm [45,70], and general relationship between the average grain
size of deposit and true current density of reduction [20]. For
simplicity, it is assumed that (i) the main mechanism of the inuence
of additives on the grain size of deposits is surface blocking
(adsorption) mechanism, (ii) all assumptions made in derivation of
Langmuir isotherm [45,70] and general relationship between the
grain size and current density [20] are valid, and (iii) the applied
current density is constant.

The cathodic overpotential is considered as a positive quantity in through paper.

i = i1 + k Cs

On the other hand, a simple and general relationship between the


average grain size of deposit (d ) and true current density (i) can be
expressed as [20]:

d = k i

where k and n are constants. Finally, by substituting Eq. (6) in


Eq. (7), for a given system at a constant current density, the
dependence of average grain size on the saccharin concentration is
obtained as:

m
d = ks 1 + kCs

356

A.M. Rashidi, A. Amadeh / Surface & Coatings Technology 204 (2009) 353358

where ks = kin is constant. All constant parameters in Eq. (8) can be


derived experimentally by curve tting. The numerical relationships
obtained by non-linear curve tting with Matlab software are:

7 0:14
d = 24:1 + 2000Cs


7 0:09
d = 15:1 + 2000Cs

10

for the present experimental results and the data reported by Erb et al.
[8,15], respectively. In Fig. 3, the experimental data has also been
compared with mathematical curve computed from Eqs. (9) and (10).
It is evident that the theoretical model is consistent with both the
present experimental results and the data reported in [8,15] for
electrodeposition of nickel coating. The model (Eq. (8)) was also
checked by curve tting for the data reported by Natter and
Hempelmann [44] who used butanediamine additive for gold
deposition and benzoic acid additive for aluminum deposition. The
results were presented in Fig. 4. Good agreement between their
experimental results and our theoretical model is evident in this
gure indicating that the general model is also applicable for other
additives used in electroplating of other metallic coatings.
3.2. Effect of bath temperature
3.2.1. Experimental results
It has been demonstrated [20] that the plot of normalized intensity
(IN,hkl = Ihkl / Ip,hkl) of each reecting planes versus diffraction angle

Fig. 4. Comparison of experimental data reported in [44] and theoretical model for
variation of the grain size of nanocrystalline, a) gold and b) aluminum deposits with
additive concentration.

(2) presents a better perspective of XRD peak broadening. For this


reason, the normalized intensity of (111) reection for nickel coatings
deposited at various bath temperatures as well as the reference
sample (annealed nickel) were presented in Fig. 5. It is qualitatively
clear that the peak width increases by increasing the bath temperature from 45 C to 55 C (Fig. 5a), but further increase in the bath
temperature decreases the peak width (Fig. 5b). Based on the change
in peak width, it seems that the grain size of deposits reduces as the
plating temperature increases up to 55 C and then increases by
further increase in the bath temperature.
The quantitative variation of the average grain size of the coatings
versus bath temperature has been shown in Fig. 6a. This gure
indicates that nanostructured nickel deposits with a mean grain size
between 24 nm and 32 nm can be obtained at all the investigated
temperatures. It is also evident that the bath temperature has a minor
effect on the average grain size of the coatings with respect to current
density (Fig. 3 of Ref. [20]) and saccharin addition (Fig. 3). Meanwhile,
considering the error bars of the measurements, it can be observed
that the variation of plating temperature does not have a considerable
inuence on the mean grain size of deposits.
3.3.2. Theoretical approach
According to the pattern presented by Dini [66,71], it is generally
expected that the grain size of the deposits increases by increasing the
bath temperature and has been experimentally observed for some
nanocrystalline deposits [4244].
The general effect of bath temperature on the grain size of the
deposits arises from the dependence of cathodic overpotential on
plating temperature. As shown by Volmer and Weber [72], the energy
of grain nucleus formation depends on the cathodic overpotential. A
large cathodic overpotential reduces the energy of nucleus formation,
and therefore increases the nucleus densities (the number of nucleus
per surface area) and renes of coating grains. Consequently, since the
cathodic overpotential decreases with increasing the bath temperature [44,66], it is expected that the grain size of deposits will increase.
Nevertheless, it should be noted that increasing the bath temperature
has two contradictory effects; (i) an increase in critical size of nucleus

Fig. 5. Normalized (111) reection peaks of ref. sample and nanocrystalline nickel
coatings deposited at (a) 45, 50 and 55 C and (b) 55, 60 and 65 C.

A.M. Rashidi, A. Amadeh / Surface & Coatings Technology 204 (2009) 353358

357

reaction, F is the Faraday's constant (96,500 As/mol). Here, for


simplicity, it is assumed that (i) all assumptions used for derivation
of Eq. (11) [7577] are valid for electroplating processes, and (ii) in
electrodeposition, the cathodic current density is enough high. So, the
Tafel equation can be used:
=

 
RT
i
ln
zF
io

12

where io (A/cm2) is the exchange current density. Introducing


Eq. (12) into Eq. (11), it can be obtained:

 B
kc
i
Ist = ko exp
io
RT

13

in which B = nc + b = . On the other hand, the experimental data


[74,78] and theoretical modeling [79,80] have indicated that the
exchange current density (io) is a function of temperature in the
general form of:


Bn
io = An exp
RT

14

where An and Bn are constant. Thus, Eq. (14) can be rewritten as:

Ist = ko

i
An

B



k BBn
exp c
RT

15

From the other point of view, a general relationship between the


average grain size of deposit (d ) and the stationary nucleation rate
(Ist) can be expressed as [20]:

d = kd

Fig. 6. Comparison of experimental results and theoretical model for variation of the
grain size of nanocrystalline, a) nickel (present work) and b) aluminum deposits [44]
with bath temperature.

due to a decrease in the thermodynamic driving force of crystallization process which leads to lower nucleus densities, and (ii) an
increase in the kinetic driving force that can lead to higher nucleation
rate [73]. At the bath temperatures in which the size of critical clusters
is in atomic dimensions, every active site can act as a critical nucleus.
So, the thermodynamic barriers for nucleus formation are negligible
and the grain size of deposit is controlled by kinetics variables.
Therefore, in such conditions, according to Arrhenius equation, the
nucleation rate increases by increasing the bath temperature [74]. It
seems that low dependence of the mean grain size of electrodeposited
nanocrystalline nickel on the bath temperature in the present work,
can be attributed to the balance between thermodynamic barriers and
kinetics variables in investigated bath temperature range.
Based on atomistic theory of electrocrystallization [7577], a
general formula can be derived to demonstrate the relationship
between the grain size and bath temperature. According to atomistic
theory of electrocrystallization, the stationary nucleation rate (Ist) can
be expressed as [75]:




kc
nc + bzF
exp
Ist = ko exp
RT
RT

11

where ko and kc are constants, R is the ideal gas law constant (8.314 J/
mol K), T is the absolute temperature (K), nc is the number of atoms
constituting the critical nucleus, (V) is overpotential and b is equal
to 1 or (the cathodic charge transfer coefcient) depending on the
mechanism of nucleus formation, z is the charge number of the

 1 = 3
i
Ist

16

where kd is constant. Finally, the combination of Eqs. (15) and (16)


yields to:
 
B

d = kf exp k
RT

17

n
in which kf = AkdB i1 = 3B and Bk = kc BB
.
3
n
Eq. (17) indicates that the increase in bath temperature, at a given
current density, may increase or decrease the grain size of deposits
depending on the sign of Bk. Assuming that Bk and kf are independent
of bath temperature, the validity of Eq. (17) was checked by curve
tting for our experimental data as well as the data reported in [44].
The results were presented in Fig. 6. As seen, the calculated curves
based on Eq. (17) for nanosized grains are consistent with the
reported experimental data [44] as well as with our ndings at the
plating temperatures above 50 C, but at the temperatures below
50 C a discrepancy was observed. The reasons of this discrepancy
could be as:

The efcient current density enhances by increasing the bath


temperature [48]. This can be probably due to the lowering of (a)
the passivation of cathode surface by precipitation of nickel
hydroxide [46,8082], (b) evolution of hydrogen [5] and dissolution of some nickel clusters [83] by increasing the bath
temperature.
The efcacy of saccharin additive improves by increasing the bath
temperature.
The nucleation processes are controlled by kinetic driving force
due to atomic dimensions of critical clusters.
Therefore, Eq. (17) should be modied and more electrochemical
experiments are still needed to reveal the explicit dependence of
nucleation rate and grain size of deposits on the bath temperature.

358

A.M. Rashidi, A. Amadeh / Surface & Coatings Technology 204 (2009) 353358

4. Conclusions
Analysis of the results led to the following conclusions:
1. Addition of saccharin to the electroplating bath reduced the grain size
of nanocrystalline nickel coatings up to a concentration of 3 g/l.
Further increase in saccharin content had no signicant effect on
grain renement.
2. Theoretical approach predicted a non-linear dependence of the
grain size on saccharin concentration, which conrmed the
experimental results.
3. The bath temperature had no signicant effect on the average grain
size of nickel deposits.
4. Theoretical model predicted an exponential dependence of the
grain size of deposits on bath temperature, which conrmed by the
experimental results.
Acknowledgements
This research was supported by the University of Tehran and Razi
University. The authors would like to thank them for nancial support
of this work. Dr. S. F. Kashani-Bozorg is also thanked for his fruitful
help.
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]

S.C. Tjong, H. Chen, Mater. Sci. Eng. R 45 (2004) 1.


L. Wang, Y. Gao, T. Xu, Q. Xue, Mater. Chem. Phys. 99 (2006) 96.
N.S. Qu, D. Zhu, K.C. Chan, W.N. Lei, Surf. Coat. Technol. 168 (2003) 123.
U. Erb, Can. Metall. Q. 34 (1995) 272.
F. Ebrahimi, Z. Ahmed, J. Appl. Electrochem. 33 (2003) 733.
D. Bera, S.C. Kuiry, S. Seal, JOM 56 (2004) 49.
A.M. El-Sherik, U. Erb, J. Page, Surf. Coat. Technol. 88 (1996) 70.
A.M. El-Sherik, U. Erb, J. Mater. Sci. 30 (1995) 5743.
I. Bakony, E. Tth-Kdr, L. Pogny, Czirki, I. Gercs, K. Varga-Josepovits, B.
Arnold, K. Wetig, Surf. Coat. Technol. 78 (1996) 124.
D. Pin-Qiang, Y. Hui, L. Qiang, Trans. Mater. Heat Treatment 25 (2004) 1283.
I. Bakony, E. Tth-Kdr, T. Tarnczi, L.K. Varga, Czirki, I. Gercs, B. Fogarassy,
Nanostruct. Mater. 3 (1993) 155.
E.A. Pavlatou, M. Raptakis, N. Spyrellis, Surf. Coat. Technol. 201 (2007) 4571.
C. Kollia, N. Spyrellis, Surf. Coat. Technol. 37 (71) (1993) 7.
E. Tth-Kdr, I. Bakonyi, L. Pogny, Czirki, Surf. Coat. Technol. 88 (1996) 57.
R.T.C. Choo, J. Toguri, A.M. El-Sherik, U. Erb, J. Appl. Electrochem. 25 (1995) 384.
Y. Xuetao, W. Yu, S. Dongbai, Y. Hongying, Surf. Coat. Technol. 202 (2008) 1895.
K.L. Morgan, Z. Ahmed, F. Ebrahimi, Mat. Res. Soc. Symp. Proce. 634 (2001)
B3.11.1, Materials Research Society.
K.P. Wong, K.C. Chan, T.M. Yue, J. Appl. Electrochem. 31 (2001) 25.
S.T. Aruna, S. Diwakar, A. Jain, K.S. Rajam, Surf. Eng. 21 (2005) 209.
A.M. Rashidi, A. Amadeh, Surf. Coat. Technol. 202 (2008) 3772.
C.A. Schuh, T.G. Nieh, T. Yamasaki, Scr. Mater. 46 (2002) 735.
F. Ebrahimi, G.R. Bourne, M.S. Kelly, T.E. Matthews, Nanostruct. Mater. 11 (1999)
343.
F. Ebrahimi, Z. Ahmed, Mater. Charact. 49 (2003) 373.
C.T.J. Low, R.G.A. Wills, F.C. Walsh, Surf. Coat. Technol. 201 (2006) 371.
A.F. Zimmerman, D.G. Clark, K.T. Aust, U. Erb, Mater. Lett. 52 (2002) 85.
A. Mller, H. Hahn, Nanostruct. Mater. 12 (1999) 259.
C.S. Lin, K.C. Huang, J. Appl. Electrochem. 34 (2004) 1013.
I. Naposzek-Bilnik, A. Budniok, B. osiewicz, L. Pajk, E. giewka, Thin Solid Films
474 (2005) 146.

[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]
[40]
[41]
[42]
[43]
[44]
[45]
[46]
[47]
[48]
[49]
[50]
[51]
[52]
[53]
[54]
[55]
[56]
[57]
[58]
[59]
[60]
[61]
[62]
[63]
[64]
[65]
[66]
[67]
[68]
[69]
[70]
[71]
[72]
[73]
[74]
[75]
[76]
[77]
[78]
[79]
[80]
[81]
[82]
[83]

S.C. Wang, W.C.J. Wei, J. Mater. Res. 18 (2003) 1566.


A. Abdel Aal, M. Bahgat, M. Radwan, Surf. Coat. Technol. 201 (2006) 2910.
A. Robin, R.Q. Fratari, J. Appl. Electrochem. 37 (2007) 805.
J. Panek, A. Budniok, Surf. Coat. Technol. 201 (2007) 6478.
I.R. Aslanyan, J.-P. Bonino, J.-P. Celis, Surf. Coat. Technol. 200 (2006) 2909.
S.T. Aruna, K.S. Rajam, Scr. Mater. 48 (2003) 507.
M. Stroumbouli, P. Gyftou, E.A. Pavlatou, N. Spyrellis, Surf. Coat. Technol. 195
(2005) 325.
L. Chen, L. Wang, Z. Zeng, T. Xu, Surf. Coat. Technol. 201 (2006) 599.
H. Wakabayashi, N. Sato, M. Sone, Y. Takada, H. Yan, K. Abe, K. Mizumoto, S.
Ichihara, S. Miyata, Surf. Coat. Technol. 190 (2005) 200.
E.M. Oliveira, G.A. Finazzi, I.A. Carlos, Surf. Coat. Technol. 200 (2006) 5978.
G. Georgiev, I. Kamenova, V. Georgieva, E. Kamenska, R. Hempelmann, H. Natter, J.
Appl. Polym. Sci. 102 (2006) 2967.
E. Moti, M.H. Shariat, M.E. Bahrololoom, Mater. Chem. Phys. 111 (2008) 469.
G.L. Flx, J.D. Pollack, Anal. Chem. 52 (1980) 1589.
H. Natter, R. Hempelmann, Electrochim. Acta 49 (2003) 51.
H. Natter, R. Hempelmann, J. Phys. Chem. 100 (1996) 19525.
H. Natter, R. Hempelmann, Z. Phys. Chem. 222 (2008) 319.
M. Paunovic, M. Schlesinger, Fundamentals of Electrochemical Deposition, WileyInterscience Publication, John Wiley & Sons, Inc, 1998.
D.R. Turner, J. Electrochem. Soc. 100 (1953) 15.
O.O. Schaus, R.J. Gale, W.H. Gauvin, Plating 58 (1971) 801.
A.H. DuRose, Plat. Surf. Finish. 64 (1977) 48.
J. Amblard, I. Eplboin, M. Froment, G. Maurin, J. Appl. Electrochem. 9 (1979) 233.
M. Holm, T.J. O'Keefe, J. Appl. Electrochem. 30 (2000) 1125.
O. Aaboubi, J. Amblard, J.P. Chopart, A. Olivier, J. Phys. Chem. B 105 (2001) 7205.
A.A. Rasmussen, P. Mller, M.A.J. Somers, Surf. Coat. Technol. 200 (2006) 6037.
D. Mockute, G. Bernotiene, R. Vilkaite, Surf. Coat. Technol. 160 (2002) 152.
T. Ungr, J. Gubicza, G. Ribrik, A. Borbly, J. Appl. Crys. 34 (2001) 298.
T. Ungr, I. Dragomir, . Rvsz, A. Borbly, J. Appl. Crys. 32 (1999) 992.
D.A. Vermilyea, J. Electrochem. Soc. 106 (1959) 66.
R. Weil, H.C. Cook, J. Electrochem. Soc. 109 (1962) 295.
T.C. Franklin, Surf. Coat. Technol. 30 (1987) 415.
L. Oniciu, L. Muresan, J. Appl. Electrochem. 2 1 (1991) 565.
J.P. Bonino, P. Pouderoux, C. Rossignol, A. Rousset, Plat. Surf. Finish. 79 (1992) 62.
T.C. Franklin, Plat. Surf. Finish. 84 (1994) 62.
D. Mockute, G. Bernotiene, Surf. Coat. Technol. 135 (2000) 42.
A. Ciszewski, S. Posluszny, G. Milczarek, M. Baraniak, Surf. Coat. Technol. 183
(2004) 127.
B. Szeptycka, Russ. J. Electrochem. 37 (2001) 684.
Y. Nakamura, N. Kaneko, M. Watanabe, H. Nezu, J. Appl. Electrochem. 24 (1994)
227.
J.W. Dini, Electrodeposition: The Material Science of Coatings and Substrates,
Noyes Publications, 1993.
W. Kim, R. Weil, Surf. Coat. Technol. 38 (1989) 289.
C.C. Roth, H. Leidheiser, J. Electrochem. Soc. 100 (1953) 553.
J. Edwards, Trans. Inst. Metal Fin. 41 (1964) 169.
I. Langmuir, J. Am. Ceram. Soc. 40 (1918) 1361.
J.W. Dini, Plat. Surf. Finish. 75 (1988) 11.
M. Volmer, A.Z. Weber, Die Kinetik der Phasenbildung, Steinkopff, Dresden, 1939.
J.W.P. Schmelzer, Mater. Phys. Mech. 6 (2003) 21.
J. Vazquez-Arenas, R. Cruz, L.H. Mendoza-Huizar, Electrochim. Acta 52 (2006) 892.
A. Milchev, Electrocrystalization: Fundamentals of Nucleation and Growth,
Kluwer Academic Publishers, 2002.
A. Milchev, S. Stoyanov, R. Kaischew, Thin Solid Films 22 (1974) 267.
A. Milchev, S. Stoyanov, R. Kaischew, Thin Solid Films 22 (1974) 255.
L. Cifuentes, J. Simpson, Chem. Eng. Sci. 60 (2005) 4915.
A.J. Bard, L.R. Faulkner, Electrochemical Methods, Fundamentals and Applications,
John Wiley & Sons, Inc, 2001.
K.J. Vetter, Electrochemical Kinetics, Academic Press, New York, 1967.
F. Lantelme, A. Seghiouer, A. Derja, J. Appl. Electrochem. 28 (1998) 907.
A.G. Munoz, D.R. Salinas, J.B. Bessone, Thin Solid Films 429 (2003) 119.
P.V. Brande, A. Dumont, R. Winand, J. Appl. Electrochem. 24 (1994) 201.

S-ar putea să vă placă și