You are on page 1of 10

MA TE RI A L S CH A R A CT ER IZ A TI O N 62 ( 20 1 1 ) 1 6 4 1 7 3

available at www.sciencedirect.com

www.elsevier.com/locate/matchar

Characterization of electrodeposited nickel coatings from


sulphamate electrolyte without additive
A. Godon a , J. Creus a , X. Feaugas a , E. Conforto b , L. Pichon c , C. Armand d , C. Savall a,
a

Laboratoire d'Etudes des Matriaux en Milieux Agressifs, EA3167, Universit de La Rochelle, Av. Michel Crpeau, F-17042 La Rochelle,
France
b
Fdration de Recherche en Environnement pour le Dveloppement Durable (FR-EDD), FR CNRS 3097, Centre Commun Analyses,
Universit de La Rochelle, 5 Alle de l'Ocan, F-17042 La Rochelle Cedex 9, France
c
Institut Pprime, UPR 3346 CNRS, Universit de Poitiers, SP2MI, Boulevard Marie et Pierre Curie, BP 30179, 86962 Chasseneuil,
Futuroscope Cedex, France
d
INSA Toulouse, Dpartement de Physique, 135 avenue de Rangueil, 31077 Toulouse Cedex 4, France

AR TIC LE D ATA

ABSTR ACT

Article history:

In this paper, the influence of deposition current density on microstructure and purity of

Received 28 June 2010

nickel coatings was studied. Complementary characterization methods (SEM, TEM, XRD,

Received in revised form

EBSD, GDOES and SIMS) were used to investigate different scales of the microstructure and

25 October 2010

to understand the metallurgical states of the coatings. As deposition current density

Accepted 18 November 2010

decreases, grain refinement and texture modifications are observed which are linked with
the grain boundary character (disorientation angle and Coincidence Site Lattice). Moreover,

Keywords:
Electrodeposited nickel
Grain refinement

in sulphamate bath without additive, the contamination by light elements and metallic
impurities strongly depends on deposition parameters and must be taken into account to
discuss the microstructure changes.
2010 Elsevier Inc. All rights reserved.

Grain boundaries
EBSD
Chemical composition

1.

Introduction

Nanocrystalline materials have been the subject of intensive


research because of their unique properties [13]. For example,
concerning the corrosion resistance of pure metals, several
works report that the susceptibility to localized corrosion is
lower in nanocrystalline materials [46], but the mechanisms
responsible for this superior corrosion resistance are not
clearly established [5,7]. As corrosion resistance can be
affected by several metallurgical parameters (defects, grain
size, grain boundary, purity, crystallographic texture, roughness, etc.), a careful control of microstructure is necessary.

Nanocrystalline nickel with a grain size below 100 nm was


obtained by electrodeposition but deposition parameters
largely vary from one study to another. For example, in
additive-free Watts bath [8] ultra-fine-grained nickel electrodeposits (grain size down to 100 nm) were obtained by pulse
plating at very high pulse-current. By using organic additives
(especially saccharin in the case of nickel), several studies
show that it was possible to produce nanocrystalline nickel
coatings in different baths with grain sizes in the range of 6
100 nm [6,8,9]. It was shown that the use of organic additives
leads to an increase of the contamination of coatings [10,11],
which can affect both mechanical properties and corrosion

Corresponding author. Tel.: +33 5 46 45 72 93; fax: + 33 5 46 45 72 72.


E-mail address: csavall@univ-lr.fr (C. Savall).
1044-5803/$ see front matter 2010 Elsevier Inc. All rights reserved.
doi:10.1016/j.matchar.2010.11.011

M A TE RI A L S CH A RACT ER IZ A TI O N 62 ( 20 1 1 ) 1 6 4 1 7 3

resistance. Few studies have tried to explain the influence of


deposition parameters by a careful analysis of the microstructure [8,12]. In most cases, only one parameter is studied,
mainly grain size which is evaluated by analyzing the
broadening of the diffraction peaks [3,9,13] or by scanning
electron microscopy [14,15]. However, it was shown that for
the same electrodeposited nickel sample, the size of structural
elements can largely vary depending on the observation tool,
and thus the microstructure needs to be evaluated at different
scales [11].
Among the different baths, sulphamate based bath is of
particular interest as it leads to ductile deposits with low
internal stress [17,18], even without sulphur (S) containing
additive [16].
In this paper, an additive-free sulphamate bath is used in
order to limit the incorporation of impurities and especially S
because of its dramatic effect on corrosion resistance. The
influence of current density on the microstructure and on
contamination of nickel coatings is studied by using different
characterization methods. The correlation between structural
observations at different scales and chemical analysis allows
understanding the metallurgical states of the coatings.

2.

Material and Methods

Nickel coatings were prepared by direct-current galvanostatic


deposition in a three-electrode cell by using a VSP potentiostat
from Biologic. A conventional sulphamate bath (V = 400 mL)
without additive was used, and composed of 300 g/L Ni(NH2
SO3)2.4H2O, 15 g/L NiCl2.6H2O, and 30 g/L H3BO3. Solutions
were prepared by dissolving pure salts in ultrapure water
(18.2 M cm) and pH was then adjusted to 4.2 by adding nickel
carbonate. Special attention was devoted to avoid contamination of the bath. A thermostated glass reactor was used to
fix the temperature at 50 C and the solution was mechanically stirred during the deposition. The anode was of pure
nickel (99.99%) and embedded in a polypropylene anode bag.
Nickel substrates (S = 2 cm2) were polished with silicon carbide
(particle size 5 m), sonicated for 2 min, rinsed with ultrapure
water and dried before electrodeposition. Deposition current
density was varied between 1 and 50 mA/cm2. In the following,
the nomenclature of samples (Table 1) refers to this deposition
parameter (for example CD 1 refers to a deposition current
density of 1 mA/cm2). Cathodic efficiency was estimated by
weighting the samples before and after deposition. Deposition
time was adjusted to obtain thicknesses of 50 m.

Table 1 Mean sizes deduced from SEM, EBSD and TEM


for coatings elaborated at different current densities.
Name

j
mA/cm2

(SEM)
m

d (EBSD)
m

d (TEM)
m

CD1
CD5
CD10
CD20
CD50

1
5
10
20
50

0.37
0.74
1.4
3.9
4.3

0.25
0.35

1.02

0.120
0.180

165

The surface morphology was observed by scanning electron


microscopy (SEM) with a FEI Quanta 200 ESEM-FEG operating at
20 kV as acceleration voltage. Electron backscatter diffraction
(EBSD) was used to obtain grain size and to characterize
microtexture and grain boundaries. For top-view EBSD analyses, samples of 75 m thickness were electrodeposited and
then electropolished in a H2SO4/CH3OH mixture [19] in order to
remove 25 m. After electropolishing, samples were very flat,
with a roughness below 2 nm (estimated by Atomic Force
Microscopy experiments). For cross-section EBSD analyses,
samples were cut with a wire saw and cross-sections were
mechanically polished. A final polishing was performed with
OPS preparation from Struers. EBSD maps were acquired at half
of the coating thickness using an acceleration voltage of 25 kV
on SEM and the TSL OIM Data collection 5 Software, with a step
size of 30 nm or 70 nm, depending on the grain size. A clean-up
was performed on maps in order to remove points which were
not indexed or to index according to the first neighbours those
which were originally incorrectly indexed. Grain size and
orientation pictures were then calculated using TSL OIM
Analysis 5 software.
Complementary transmission electronic microscopy (TEM)
observations were carried out with a JEOL JEM 2011 electron
microscope operating at 200 kV. Foils for TEM were thinned in
double twin-jet electro-polisher using an electrolyte of 25%
nitric acid and 75% methanol at a temperature of 30 C and a
current of 150 mA. To understand the microstructure observed
at high current density, TEM observations were also performed on the cross-sections of sample CD50. For this
specimen, stereographic analyses (stereographic projection)
were established for each observed grain in order to evaluate
the orientation of each grain. Special care was taken in the
marking of TEM specimens. So, the direction of the normal of
the electrodeposited surface was identified on the stereographic map of each studied grain.
X-ray diffraction analyses in 2 mode were performed on
a Brucker apparatus (AXS D8-Advanced) with the Cu-K
radiation ( = 0.15405 nm). Spectra were acquired between 40
and 100, with a step width of 0.02 and the K2 peak and
background were removed. Composition analyses were
obtained by Glow Discharge Optical Emission Spectrometry
(GD Profiler from Horiba Jobin Yvon). Secondary Ion Mass
Spectrometry (IMS 4FE6 from CAMECA) was also used with
two ionic sources Cs+ (at 14.5 keV) and O+2 (at 5.5 keV) to obtain
the best sensitivity. Concentration profiles were acquired after
a pulverization of 5 to 10 m in order to avoid surface
contamination effects. All atomic elements were analyzed
except nitrogen. For both methods, the detection limit for this
element was too high. Calibration with bulk nickel samples of
known composition was performed for quantitative analysis.
Several profiles were obtained for each sample, leading to
reliable results. However, due to the small volumes which are
analyzed by these techniques, concentration values cannot be
given with a high accuracy.

3.

Results

The aim of this work is to use complementary analyses to obtain


an overview of the metallurgical state of electrodeposited

166

MA TE RI A L S CH A R A CT ER IZ A TI O N 62 ( 20 1 1 ) 1 6 4 1 7 3

coatings. Section 3.1 describes surface morphologies in relation


with SEM observations. Section 3.2 outlines the interest to use
X-ray diffraction analyses to study the macroscopic texture and
to extract dimensional data. The following Sections 3.3 and 3.4
deal with the opportunity to obtain spatial information using
EBSD maps and TEM analyses. Finally, chemical composition
was analyzed in connection with structural results.

3.1.

Scanning Electron Microscopy

SEM views presented in Fig. 1 show the surface morphology of


coatings prepared at different current densities. At high
current densities, large crystallites with a truncated pyramidal
shape are observed leading to bright deposits in good
agreement with previous results in sulphamate bath [12,15].
A strong hydrogen evolution leads to the formation of bubbles
and edge effects at current densities above 50 mA/cm2. As the
current density decreases, this pyramidal morphology is
replaced by a nodular morphology. The mean size of the
nodules deduced from SEM was estimated by statistical
analyses of images obtained at different magnifications
(Table 1). The values suggest a refinement at low current
density. However, the morphological features observed by

SEM cannot be directly assigned to grains and other characterization tools will be used below to clarify this point.

3.2.

X-ray Diffraction Analysis

The diffraction patterns for different deposition current


densities are plotted on Fig. 2. At high current density (above
30 mA/cm2), a strong crystallographic texture along the <100>
direction is observed, which is replaced by a <110> preferred
orientation at current densities below 20 mA/cm2. At 1 mA/cm2,
no preferred orientation is observed but the (220) line is
slightly high and the (200) one is slightly low respectively to a
non texture nickel sample (JCPDS data no. 00-004-0850).
Complementary texture analysis by using inverse pole figures
obtained by EBSD will be presented in Section 3.3, confirming
the above results. For coating CD1, a broadening of the
diffraction peaks can be noticed, suggesting a grain refinement
effect. Assuming a Cauchy-shaped profile, the full width at
half maximum (FWHM) was evaluated for each diffraction
peak, after correction by the experimental broadening estimated by using the LaB6 standard sample. The Scherrer
equation obviously led to a strong underestimation of
the grain sizes of these coatings. So, an approach based on

Fig. 1 SEM top views showing the surface morphology of the coatings deposited at different current densities. (a: CD1 (1 mA/cm2),
b: CD5 (5 mA/cm2), c: CD10 (10 mA/cm2), and d: CD50 (50 mA/cm2)).

M A TE RI A L S CH A RACT ER IZ A TI O N 62 ( 20 1 1 ) 1 6 4 1 7 3

Fig. 2 2 scans of coatings elaborated at different current


densities.

WilliamsonHall diagrams was used in order to estimate


crystallite sizes and internal stresses. The approach developed
by Reimann [20] and used by Thiele [11] in electrodeposited
nickel was followed, which takes into account the elastic
anisotropy of nickel.
The WilliamsonHall plots obtained for coatings deposited
at 1 mA/cm2 (CD1) led to a mean internal stress (<2>1/2)
of 300 MPa. This value is in the range of those previously
reported in electrodeposited nickel [11] which showed an
increase of mean internal stress as the grain size decreases.
The mean size of coherent scattering regions for sample CD1
deduced from this analysis is around 130 nm. For coatings
deposited at higher current densities, the broadening of the
diffraction peaks is smaller. Moreover, for these coatings, the
presence of a crystallographic texture does not allow this
approach.

3.3.

Electron Backscatter Diffraction

Top-view orientation maps for coatings elaborated at different


current densities are presented on Fig. 3. Inverse pole figures
were calculated from these orientation maps, showing the
orientation densities for the different crystallographic directions parallel to the sample normal direction. The preferred
orientation along the <100> direction suggested by 2 XRD
scans for coatings prepared at 50 mA/cm2 (CD50) is confirmed.
Comparison with SEM views shows that the large truncated
pyramidal structures are mainly oriented with their <100> axis
perpendicular to the substrate surface. Between these pyramidal grains, much smaller grains are found, with different
crystallographic orientations. Even if a preferred orientation
along the <110> direction is found for the coating prepared at
5 mA/cm2 (CD5), the texture is less marked (as the proportion
of pixels which <110> crystal direction is disoriented versus the
sample normal direction is higher). For the coating deposited
at the lowest current density (CD1), the crystallographic
texture along the <110> direction is very weak, in accordance
with 2 XRD scans. Grain boundary position is superimposed as grey lines to the orientation maps of Fig. 3.
Neighbouring pixels in the map with disorientation smaller
than 5 are associated with the same grain. According to this

167

disorientation angle, the grain size distribution can be


measured and a mean grain size (dEBSD) can be evaluated.
For each sample, the analyzed area was large enough to take
into account more than 5000 grains. The results are given in
Table 1, and in accordance with SEM observations, the grain
size decreases and the grain distribution becomes narrower
when the deposition current density is reduced.
EBSD orientation maps obtained on cross-sections of
different deposits are given on Fig. 4. The growth direction
which is perpendicular to the surface of the substrate is also
shown on this figure. For the CD50 sample, fibers (whose axis
is perpendicular to the substrate surface) characterized by a
dominant colour are observed. These fibers are formed by
grains slightly disoriented with regard to the neighbouring
grains, but with the (100) direction mainly parallel to the
growth direction. Between these fibers, some less oriented
regions are found. The thickness of these fibers (around 5 m)
is quite similar to the size of large crystallites with a truncated
pyramidal shape, which are observed on the surface (4.3 m,
Table 1). As the deposition current density decreases, these
fibers are no longer observed and the mean size of the grains
decreases. It can be noticed that the grains do not show any
elongation along the growth direction whatever the deposition current density.
Two parameters are mainly used to describe the nature of
grain boundaries: the disorientation angle and the factor,
which denotes the fraction of atoms in the grain boundary
plane which are coincident to both lattices. These parameters
were evaluated by using EBSD [21] and are given in Table 2
and Fig. 5. An increase of the fraction of high angle grain
boundaries (HAGB) is observed as the grain size decreases and
as the marked texture along the <100> direction is replaced by
a weak texture along the <110> direction (Table 2). The amount
of coincidence site lattice (CSL) is also strongly modified,
showing a decrease of the abundance of 1 boundaries and an
increase of the number of 3 and 9 boundaries when the
grain size decreases (Fig. 5).

3.4.

Transmission Electronic Microscopy

Grain size was evaluated using TEM observation on a


population around 150 grains and the mean values are given
for CD1 and CD5 in Table 1. These values are lower than the
ones obtained by EBSD, but for the CD1 sample, the value is in
agreement with XRD analysis (130 nm). As a strong heterogeneity of grain sizes was observed for CD50, the mean value is
not relevant for this sample. TEM observations were also
performed on cross sections for this sample to evaluate the
crystallographic orientation of different grains. These analyses are time consuming, thus only a semi-statistical study
on 56 grains at different locations inside the sample was
performed. However, 56 grains seemed to be sufficient to
reflect the heterogeneity of the sample, as the results were not
significantly modified when this number was increased.
Different populations of grains were identified, characterized
by three angles (100), (111) and (110) (Fig. 6). (hkl) relates
the angle between (hkl) plane and the normal to the coating
surface. The first one (V1) corresponds to the largest grains
(>700 nm) and exhibits an angle (100) near 0. This means that
this crystallographic population mainly contributes to the

168

MA TE RI A L S CH A R A CT ER IZ A TI O N 62 ( 20 1 1 ) 1 6 4 1 7 3

Fig. 3 Left: Top view orientation maps obtained by EBSD for coatings elaborated at different current densities: CD1 (a), CD5 (b),
and CD50 (c). Right: Inverse pole figures of the normal direction for the three coatings.

macroscopic texture observed by XRD. In a random zone (cf.


EBSD analyses), three other kinds of crystallographic populations were identified (Fig. 6), which do not correspond to
macroscopic texture obtained by XRD. The size of these grains
is generally lower (130 to 250 nm) than the grain with V1
variant. The correlations of these observations with SEM and
EBSD results show that two kinds of regions can be distinguished in the CD50 coating: the first one corresponds to large

grains with a <100> preferred orientation and the second one


is associated with random regions, with a much lower grain
size and weaker texture.

3.5.

Composition Analysis

Table 3 lists the different elements detected in the coatings


and their contents in weight ppm obtained by SIMS and

M A TE RI A L S CH A RACT ER IZ A TI O N 62 ( 20 1 1 ) 1 6 4 1 7 3

169

Fig. 4 Cross-section orientation maps obtained by EBSD for coatings elaborated at different current densities: CD1 (a), CD5 (b),
and CD50 (c). The substrate surface normal is given by an arrow.

GDOES. For the coating CD50, impurity amounts are very low,
leading to a purity around 99.99%. However, for the coatings
prepared at lower current density the contamination drastically increases especially for light elements (H, O, C, etc.) and
for Cl and Cu. For these coatings, some impurity contents are
given with a large inaccuracy, and the purity of the coating
could not be evaluated. In these cases and especially for
chloride for which the concentration in ppm was not given,
the quantification was not reliable as the reference samples
contained much lower amounts of these elements. Concentration profiles and cartographies were obtained for each
atomic element, showing that the impurities were homoge-

neously distributed laterally and through the thickness of the


coatings.

4.

Discussion

Electrodeposited layers often exhibit a fiber texture, i.e.


preferred crystallographic orientation of their crystallites
along the growth direction, which is the case for deposits
CD5 and CD50. Our results are in good agreement with
published results for sulphamate bath which report a strong
crystallographic texture along the <100> axis associated

170

MA TE RI A L S CH A R A CT ER IZ A TI O N 62 ( 20 1 1 ) 1 6 4 1 7 3

Table 2 Classification of the grain boundaries character


for coatings deposited at different current densities. Lowangle grain boundaries (LAGB) are characterized by a
disorientation angle below 15 and high-angle grain
boundaries are characterized by a disorientation angle
above 15.
CD1
LAGB/HAGB
(%)
LCSL/HCSL
(%)

CD5

CD50

93

91

42

58

42

58

57

43

57

43

CSL denotes coincidence site lattice (CSL) with low sigma (LCSL,
< 29) or high sigma (HCSL, > 29).

with large grains for deposition currents between 30 and


200 mA/cm2 [12,13]. Based on cross-section observations by
optical microscopy or SEM after chemical etching, the <100>
texture has been associated with the formation of long
columnar grains, some of them extending across the whole
thickness of the coatings (10 to 50 m) [15,18]. Cross-section
observations of the CD50 coating by optical microscopy after
acidic etching show kinds of columns, parallel to the growth
direction, with a width of few micrometers (Fig. 7a). EBSD
and TEM observations on cross sections allow to distinguish
unambiguously the grains and to evaluate their orientations.
The results obtained by EBSD clearly show that these columns
are formed by grains weakly disoriented with regard to their
neighbouring, with the <100> direction perpendicular to the
substrate surface. This microstructure, associated with a quite
high amount of low angle grain boundaries and particularly of
1 boundaries explains the large disagreement between the
structure size elements deduced by EBSD and SEM observations
in these coating (Table 1). Results obtained by TEM confirm that
the largest grains (and the more numerous) are oriented with
one direction <100> parallel to the growth direction. However, a
significant amount of grains, much smaller (<250 nm), is
differently oriented.
As the current densities decrease, grain refinement is
observed associated with the evolution of the <100> texture
towards a less marked <110> texture, in good agreement with
published results in similar bath [18,23]. In the literature, such

Fig. 5 Amounts of LCSL plotted versus value for coatings


deposited at different current densities.

texture changes are associated with a grain size decrease,


obtained by using pulse plating [2,22] or organic additives
[23,24]. For coatings deposited at low current densities, very
thin fibers parallel to the growth direction are observed on
cross-sections after chemical etching (Fig. 7b). However, EBSD
observations clearly show that the grains are not elongated
along the growth direction. In agreement with the results
obtained for the coating CD50, EBSD is a powerful tool to
observe the microstructure of electrodeposited coatings and to
avoid artefacts linked with chemical etching necessary to
display grains with more conventional observation techniques. EBSD analyses presented here show that these evolutions of texture and grain size are associated with an increase
of the amount of high angle grain boundaries. Particularly, a
decrease of the fraction of 1 boundaries and an increase of
the fraction of 3 boundaries are observed as the grain size
decreases. Similar trends were reported in copper proceeded
by equal Channel Angular Extrusion [25]. Several studies suggest that the presence of low- coincidence site lattice (CSL)
boundaries could be associated with a better corrosion
resistance [7,26] and with better mechanical properties [27].
However, this can be moderated by the fact that the presence
of low- coincidence site lattice boundaries seems to be
correlated with high impurity contents.
At low current densities, equiaxed nickel with very low
grain sizes can be deposited in direct current mode without
additive. For these deposits, good correlations between grain
sizes deduced from EBSD and nodule sizes deduced from SEM
are found. At low deposition current densities, the random
orientation of grains is linked with high disorientations
between grains, which appear as distinct entities in top-view
SEM observations of the coatings. A mean grain size value
around 250 nm is found by EBSD (with a disorientation of 5)
for the coating prepared at the lowest current density. A much
smaller value is deduced from analyzing the broadening of the
diffraction pattern in accordance with the grain size measured
by TEM. Thus, the choice of the disorientation angle value
used to define grain size by EBSD needs to be validated by a
correlation with XRD or TEM analyses.
Our results show that the grain refinement and the
changes of crystallographic orientation are linked with
an increase of the amounts of several impurities. As the
nickel anode was of high purity, Co and Cu contamination
probably originates from the chemicals of the bath. Copper,
which is nobler than nickel, is preferentially deposed at
low current density (and thus at low overpotential).
Because of their position in the Mendelev table, these
metallic impurities can easily replace nickel in the coating
and should act as substitution impurities. The changes of
the microstructure are probably linked with the incorporation of light atomic elements. Inhibition phenomena
[28,29], are known to strongly influence electrocrystallisation processes. In the case of nickel electrodeposition
from Watts bath, Amblard et al. [30] show that interfacial
inhibitors (Hads, Ni(OH)2, etc.) led to several growth modes
and textures, depending on deposition parameters, but
different species including C, O, N, H or Cl atoms were
proposed [29,31]. A significant drop of the deposition efficiency
was observed as the deposition current density decreases from
98% (for coatings deposited at 50 mA/cm2) to around 84% (for

M A TE RI A L S CH A RACT ER IZ A TI O N 62 ( 20 1 1 ) 1 6 4 1 7 3

171

Fig. 6 TEM observations (CD50) and stereographic projections showing the orientation of different grains in a random
oriented region. The table gives the grain population in terms of angle between the coating normal surface and the (hkl) plane.

coatings deposited at 1 mA/cm2). Voltammograms obtained in


the plating bath with the same stirring conditions allowed us
to estimate the dioxygen reduction current density around

Table 3 Impurity content in weight ppm for coatings


deposited at different current densities. For the values in
italics, the quantification was not possible as the
reference samples contained a much lower amount of
these impurities.

CD50
CD5
CD1

Cl

Fe

Co

Cu

Mo

1
70
130

5
100
400

25
400
1000

<1
6
25

<1
230
1000

4
15
7

32
100
155

25
150
600

<40
<40
<40

0.1 mA/cm2 and thus the contribution of this reaction could


explain the decrease of deposition efficiency. At low current
densities, and thus low deposition rates, this reaction could
hinder the growth of crystallites, contributing to the refinement
effect. More generally, the adsorption of different foreign species
(including O, H, C, and Cl) at the cathode surface probably
prevents grain growth by avoiding surface diffusion of adatoms
and significant amounts of these species are incorporated into
the coatings. EBSD analyses show that, in coatings deposited at
low current density, grain boundaries are more defective (higher
disorientation angle and factor) with probably an increased
concentration of vacancies. Thus, the results are consistent
with a decrease of grain size when current density decreases,
associated with the incorporation of impurities at grain
boundaries.

172

MA TE RI A L S CH A R A CT ER IZ A TI O N 62 ( 20 1 1 ) 1 6 4 1 7 3

and higher contamination. Both light elements and substitution impurities are incorporated when grain size
decreases, which can affect mechanical properties and
corrosion resistance. So, chemical contamination of electrodeposited coatings must be carefully evaluated before
discussing the influence of their microstructure on
properties.

Acknowledgement
Thanks are due to the Agence Nationale de la Recherche (GIP
ANR Program no. ANR-06JCJC-0023-01) for the financial
support.

REFERENCES

Fig. 7 a) Cross-section view of the coating CD50 obtained by


optical microscopy after chemical etching, and b) cross-section
view of the coating CD5 obtained by SEM after chemical
etching.

5.

Conclusion

Although extensive experimental works have been published concerning characterization of nickel electrodeposited coatings, a study combining composition analyses and
multi-scale microstructural characterization is missing. In
sulphamate bath without additive, microstructure modifications are linked with the incorporation of impurities and
particularly light atomic elements whose content largely
depends on electrodeposition conditions. Deposits obtained
at current densities above 20 mA/cm2 show a strong <100>
texture along the growth direction but are characterized by
different structural heterogeneities which can be evidenced
by using complementary observation tools. TEM and EBSD
observations offer the opportunity to distinguish the different microstructural scales and to better understand the
microstructure of coatings. As the current density decreases,
grain refinement and texture modifications are observed
which are associated with more defective grain boundaries

[1] Qin L, Xu J, Lian J, Jiang Z, Jiang Q. Surf Coat Technol 2008;203:


1427.
[2] Qu NS, Zhu D, Chan KC, Lei WN. Surf Coat Technol 2003;168:
1238.
[3] Rashidi AM, Amadeh A. Surf Coat Technol 2008;202:37726.
[4] Kim SH, Erb U, Aust KT, Gonzalez F, Palumbo G. Plat Surf
Finish 2004;91:6870.
[5] Miyamoto H, Harada K, Mimaki T, Vinogradov A, Hashimoto
S. Corros Sci 2008;50:121520.
[6] Mishra R, Balasubramaniam R. Corros Sci 2004;46:301929.
[7] Roy I, Yang HW, Dinh L, Lund I, Earthman JC, Mohamed FA.
Scr Mater 2008;59:3058.
[8] El Sherik AM, Erb U, Page J. Surf Coat Technol 1996;88:
708.
[9] Ebrahimi F, Ahmed Z. J Appl Electrochem 2003;33:7339.
[10] Wang YM, Cheng S, Wei QW, Ma E, Nieh TG, Hamza A. Scr
Mater 2004;51:10238.
[11] Thiele E, Klemm R, Hollang L, Holste C, Schell N, Natter H,
et al. Mater Sci Eng A 2005;390:4251.
[12] Rasmussen AA, Moller P, Somers MAJ. Surf Coat Technol
2006;200:603746.
[13] Zhao H, Liu L, Zhu J, Tang Y, Hu W. Mater Lett 2007;61:
16058.
[14] Xuetao Y, Yu W, Dongbai S, Hongying Y. Surf Coat Technol
2008;202:1895903.
[15] Banovic SW, Barmak K, Marder AR. J Mater Sci 1998;33:
63945.
[16] Kelly JJ, Goods SH, Talin AA. Electrochem Soc Proc 2004;17:
43247.
[17] Baudrand D. Met Finish 1996;94:158.
[18] Marquis EA, Talin AA, Kelly JJ, Goods SH, Michael JR. J Appl
Electrochem 2006;36:66976.
[19] Sahal M, Creus J, Sabot R, Feaugas X. Acta Mater 2006;54:
215767.
[20] Reimann K, Wrschum R. J Appl Phys 1997;81:718692.
[21] pdf database EBSD, OIMDC Manual, Copyright 19972006,
EDAX-TSL.
[22] Fritz T, Cho HS, Hemker KJ, Mokwa W, Schnakenberg U.
Microsyst Technol 2003;9:8791.
[23] Ebrahimi F, Bourne DG, Kelly MS, Matthews TE. Nanostruct
Mater 1999;11:34350.
[24] Lin CS, Hsu PC, Chang L, Chen CH. J Appl Electrochem 2001;31:
92533.
[25] Dalla Torre FH, Gazder A, Gu CF, Davies CHJ, Pereloma EV.
Metall Mater Trans A 2003;38:108095.
[26] Palumbo G, Aust KT, Lehockey EM, Erb U, Lin P. Scr Mater
1998;38:168590.

M A TE RI A L S CH A RACT ER IZ A TI O N 62 ( 20 1 1 ) 1 6 4 1 7 3

[27] Godon A, Creus J, Cohendoz S, Conforto E, Feaugas X, Girault


P, Savall C. Scr Mater 2010;62:4036.
[28] Fischer H. Electrod Surf Treat 1972/73;1:31934.
[29] Winand R. Electrochim Acta 1994;39:1091105.

[30] Amblard J, Epelboin I, Froment M, Maurin G. J Appl


Electrochem 1979;9:23342.
[31] Natter H, Schmelzer M, Hempelmann R. Mater Res Soc
1998;13:118697.

173

You might also like