Electrodeposition of Nanocrystalline Nickel - Cobalt Binary Alloy Coatings: A Review

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

REVIEW

Electrodeposition of nanocrystalline nickel–


cobalt binary alloy coatings: a review
C. Ma1, S. C. Wang*2 and F. C. Walsh2,3
The history of Ni–Co alloy deposits is summarised and recent developments are highlighted.
Electrodeposition of nanocrystalline Ni–Co binary alloys and, briefly, Co–Ni–P ternary coatings is
considered, including chemical composition, phase composition and crystalline nature.
Nanostructure, morphology, physical and mechanical properties (including corrosion resistance)
vs bath type and composition (including pH and electrolyte additives) and plating conditions,
including current density, temperature and agitation are summarised. Applications range from
wear and corrosion resistant coatings, particularly as a hard chromium replacement to speciality
hydrogen evolution electrodes in water electrolysis. Following this extensive review, future
research needs are briefly listed.
Keywords: Alloy deposits, Corrosion resistance, Electrocrystallisation, Electron diffraction, Electroplating, Hardness, Nanostructure, Transmission electron
microscopy

Introduction nanostructured alloy deposition plating by


selecting and optimising bath additives and
Part 1 of this series of reviews considered the electro- plating conditions
deposition of nanocrystalline nickel and cobalt as single (iii) despite the studies on coarse-grained deposits,
metals.1 In this paper, the authors focus on the there have been few comparable investigations
importance of nanocrystalline Ni–Co alloy deposits. of nanocrystalline Ni–Co coatings. Numerous
The advantages of electroplating compared to other mechanisms have been proposed to explain the
techniques such as physical vapour deposition, chemical deviation from the conventional Hall–Petch
vapour deposition, high velocity oxygen fed flame relationship for nanocrystalline materials when
spraying, plasma spraying, laser cladding and weld the grain size is reduced to a threshold value.
facing, etc., were summarised in Part 1. In summary, However, none of these is focused on the
direct electroplating offers versatility, reasonable speed, influence of surface porosity, which is common
relatively low costs together with the ability to control in thin electrodeposited coatings. Deviation
thickness and structure of the deposits, even on complex from the Hall–Petch relationship may partly
shapes. Further research is needed to better understand derive from the porosity
the microstructure and properties of Ni–Co coatings and (iv) few studies on the tribological properties of Co–
to develop novel Co–Ni–P coatings (to be discussed in Ni–P coatings have been reported in the
Part 3 of this series) with improved wear and corrosion literature, despite their attraction for anti-wear
resistance to replace hard chromium. The key require- applications. The novel ternary alloys are
ments of research in this area involve: expected to combine the lubricity of cobalt-rich
(i) the role of tribofilms and debris on the Co–Ni coatings and the precipitation hardening
tribological behaviour of Ni–Co coatings of achieved by heat treatment of Ni–P alloys
different cobalt contents needs clarification (forming nickel–phosphide phases) to achieve
(ii) current methods such as compositionally comparable or superior wear resistance to hard
graded coatings and application of pulsed chromium. The composition gradient and
current have limitations including complex microstructure changes from nanocrystalline
equipment and high costs. Moreover, the to amorphous structure with thickness of Co–
microhardness needs to be further increased. Ni–P coatings also require study. Electro-
There is a clear case to realise simple, direct deposited Co–Ni–P coatings will be considered
in the third of this series of reviews.2
1 To provide more competitive coatings and improve the
Crown Packaging UK plc, Downsview Road, Wantage, Oxon OX12 9BP,
UK technology of Ni–Co alloy coatings, it is important to
2
National Centre for Advanced Tribology at Southampton and Materials consider the electrodeposition of Ni–Co nanocrystalline
Engineering Research Group, University of Southampton, Highfield,
Southampton SO17 1BJ, UK
deposits in some detail together with structure–property–
3
Electrochemical Engineering Laboratory, University of Southampton, plating conditions interrelationships. Hence this review
Highfield, Southampton SO17 1BJ, UK considers structure and crystal phase composition in detail
*Corresponding author, email wangs@soton.ac.uk then considers the effect of electroplating conditions on

ß 2015 Institute of Materials Finishing


Published by Maney on behalf of the Institute
Received 12 December 2013; accepted 3 October 2014
104 DOI 10.1179/0020296714Z.000000000218 Transactions of the IMF 2015 VOL 93 NO 2
Ma et al. Electrodeposition of nanocrystalline nickel–cobalt binary alloy coatings

the structural and physical properties along with the against a stainless steel ball under dry sliding conditions
corrosion resistance and tribology of the deposits. compared to nickel-rich alloys. As shown in Fig. 5, the
Nanocrystalline Ni–Co electrodeposits have the coefficient of friction of Ni–Co alloys with cobalt content
potential to be used in aerospace, automobile and lower than 49 wt-% was higher than 0.6, which is similar to
general industries as an alternative to hard chromium that of the pure nickel coating. It started to decrease when
coatings due to their high hardness, wear-resistance and the cobalt content was 66 wt-% and the 81 wt-% Co–Ni
anti-corrosion resistance.3 As with nickel or cobalt exhibited a dramatic reduction of coefficient of friction,
coatings, the Ni–Co alloys can be deposited from which was approximately 0.25 (less than half the value of
sulphamate baths, Watts type baths and chloride baths. Ni and Ni-rich alloys). The friction reduction seen using
Tury et al.4 reported that the coatings plated from the cobalt-rich alloys was attributed to the change of crystal
former two baths were homogeneous in the whole cross- structure from fcc to hcp. Figure 6 illustrates the improve-
section, but had much lower microhardness than those ment of wear resistance with increasing cobalt peak current
produced from chloride baths. In the past few years, the density9 content in the Ni–Co deposits.
preparation and optimisation of the coatings have When the cobalt content was lower than 49 wt-%, the
drawn extensive attention, especially the relationship wear rate gradually decreased due to the increase in
among experimental parameters, microstructure and microhardness. In spite of its lower hardness, the wear
properties. rate of 81 wt-% Co–Ni was over one order of magnitude
lower than that of pure Ni and Ni-rich alloys. The
Deposit properties difference in the wear behaviour of Ni–Co alloys can be
verified by the worn surface morphologies shown in
Microstructure Fig. 7. The wear tracks on deposits of pure nickel
The surface morphology of Ni–Co alloys strongly (Fig. 7a) and 27 wt-% Co–Ni (Fig. 7b) with fcc structure
depends on their cobalt content as shown in Fig. 1.2 It exhibited the larger extent of adhesion wear and severe
changed from large polyhedral crystallites to small deformation in the sliding direction under the combined
spherical clusters as the cobalt content increased from stresses of compression and shear. However, smooth
7 to 66 wt-%Co. The 81 wt-%Co–Ni coating exhibited surfaces with smaller damaged regions, and only light
the branched structure as shown in Fig. 1f. The phase grooves and scars, were observed on the worn surface of
structure gradually changed from fcc to hcp as the the 81 wt-% Co–Ni with hcp structure (Fig. 7c). The hcp
cobalt content increased.2 As shown in Fig. 2e, the structure of cobalt-based alloys (Co–Re, Co–Mo and
66 wt-% Co–Ni coating contained the mixture structure Co–Cr) leads to a lower coefficient of friction and wear
of fcc and hcp, while the 81 wt-% Co–Ni in Fig. 2f had a rate than their counterparts with fcc structure.10,11
strong hcp (002) texture with pronounced (100) and The tribological behaviour of electrodeposited nano-
(101) peaks, similar to the structure of the pure cobalt crystalline Ni–Co coatings having controlled cobalt
coating shown in the first of this review series.1 Myung content has been considered in a recent paper.12 The
and Nobe5 reported that the transition from fcc to hcp worn surfaces and wear debris were characterised by
structure occurred when the cobalt content was in the surface analysis techniques following dry sliding tests
range of 72–78 wt-%. against a stainless steel ball. The tribofilms containing
Microhardness iron from the stainless steel counterpart were formed on
the worn surface of the coatings (less than 60 at.-%Co),
Generally, the microhardness of nanocrystalline Ni–Co
coating with different grain sizes electrodeposited by which exhibited high coefficients of friction and wear
direct current was in the range of 300–500 HV depend- rates. No tribofilm or iron transfer from the pin were
ing on the cobalt content as shown in Fig. 3.2 It was found on the Co-rich coatings (more than 70 at.-%Co),
found that the microhardness of Ni–Co alloys increased but there was a dramatic reduction of 50% in friction
initially with Co content from 0 to approximately 49 wt- and an improved wear resistance was reported.
%, then gradually decreased with a further increase the
Corrosion behaviour
cobalt content. The similar evolution of microhardness
as the function of cobalt content was also reported by Myung and Nobe5 investigated the corrosion resistance
Golodnitsky et al.6 and Srivastava et al.7 Wang et al.3 of Ni–Co alloys plated from chloride baths having
claimed that the Ni–Co alloys exhibited a nearly different cobalt contents in 0.5M NaCl after 1 h
constant Hall–Petch gradient. It is inappropriate to immersion. As shown in Fig. 8, the corrosion resistance
apply the Hall–Petch relationship to different materials, increased slightly with increasing cobalt content until it
however, since k in equation (1) is a material dependant reached the maximum at 70%Co and then sharply
constant.8 Srivastava et al.7 attributed the decrease in decreased. The nickel-rich deposits with fcc structure
hardness with the cobalt content higher than 50 wt-% to had an order of magnitude higher corrosion resistance
the formation of fibrous structure and the transforma- than cobalt-rich coatings with hcp structure. Srivastava
tion from fcc to hcp crystal structure. Li et al.9 reported et al.7 found that Ni–20%Co alloy plated from
that the Ni–Co coatings produced by pulse current sulphamate baths exhibited better corrosion resistance
plating had higher hardness (400–600 HV) than the in 3.5% NaCl in comparison to other Ni–Co alloys with
coatings plated by direct current depending on the higher cobalt content, which was attributed to its dense
composition and grain size due to different peak current surface and fcc structure. Hassani et al.13 reported that
densities, as shown in Fig. 4. the nanocrystalline 5 wt-% Co–Ni alloys with smoother
surface deposited from a bath containing saccharin had
Tribological behaviour higher corrosion resistance in 10 wt-% NaOH compared
Research conducted by Wang et al.3 showed that the to coatings with the same composition plated from
cobalt-rich alloy has excellent friction–reduction behaviour additive-free baths.

Transactions of the IMF 2015 VOL 93 NO 2 105


Ma et al. Electrodeposition of nanocrystalline nickel–cobalt binary alloy coatings

1 Surface morphologies of Ni–Co alloy deposits with Co contents of a 0 wt-%, b 7 wt-%, c 27 wt-%, d 49 wt-%, e 66 wt-%, f
81 wt-%, and g high magnification image of Ni–49 wt-%Co alloy3

106 Transactions of the IMF 2015 VOL 93 NO 2


Ma et al. Electrodeposition of nanocrystalline nickel–cobalt binary alloy coatings

4 Microhardness of Ni–Co coatings as a function of peak


current density during pulsed current plating9

2 XRD patterns of Ni–Co alloy deposits with Co contents


of a 0 wt-%, b 7 wt-%, c 27 wt-%, d 49 wt-%, e 66 wt-%
and f 81 wt-%3

Importance of experimental parameters


In order to optimise the properties of nanocrystalline
Ni–Co electrodeposition, the electroplating variables, 5 Coefficient of friction as a function of Co content in
such as cobalt ion concentration, current density, Ni–Co alloys3
additive, bath temperature and bath agitation, should
be appropriately adjusted. Previous studies showed that in Fig. 9.3 It was found that the percentage of cobalt in
these factors had different effects on the microstructure alloys was higher than the Co2z/(Co2zzNi2z) molar
and composition of Ni–Co coatings and consequently ratio in the electrolyte.3,14 The less noble metal (Co) was
led to different hardness, tribological and electrochemi-
preferentially deposited. Bai and Hu15 considered the
cal properties, which are now discussed.
anomalous deposition to be attributable to the higher
Effect of cobalt ion concentration in electrolyte adsorption ability of Co(OH)z compared to Ni(OH)z
on the cathode surface. The suggested reaction mechan-
With a fixed concentration of nickel ions, the cobalt
ism of Ni–Co electrodeposition is15
content in alloy deposit increased gradually with the
increasing Co2z concentration in electrolytes as shown

3 Microhardness as a function of Co content3 6 Wear rate as a function of Co content in Ni–Co alloys3

Transactions of the IMF 2015 VOL 93 NO 2 107


Ma et al. Electrodeposition of nanocrystalline nickel–cobalt binary alloy coatings

8 Corrosion resistance of Ni–Co alloys in 0.5M NaCl after


1 h immersion5

has been discussed in the section on ‘Corrosion


behaviour’.
Effect of current density
Fan and Piron16 reported that the current efficiency for
Ni–Co codeposition in chloride baths increased with
increasing current density, and reached 96–98% at
current densities from 0.5 to 10 A dm22. It decreased
with the further increase in current density. The coating
deposited with the current density of l00 A dm22 was
porous and fragile, and detached partially from the
electrode due to vigorous hydrogen evolution during
deposition. High current density can be applied to
produce nanocrystalline Ni–Co alloys by jet electro-
deposition14 and pulsed current deposition.9 The cobalt
content decreased as the peak current density increased
as shown in Fig. 1014 and Fig. 11,9 respectively. The
increase in cathodic current density leads to the cathodic
overpotential which favours the activation-controlled
nickel deposition, whereas diffusion-controlled cobalt
deposition has fewer responses. Therefore, higher
current density results in lower cobalt content in the
deposits.16 As shown in Fig. 12, the grain size of Ni–Co
coatings electroplated by pulsed current deposition
decreased rapidly with the increasing peak current
density to around 20 nm at l00 A dm22 then slowly

7 Worn surface of a pure nickel and Ni–Co alloy deposits


containing b 27 wt-% Co–Ni and c 81 wt-% Co–Ni3

{
2H2 Oz2e{ ~H2 z2OH (2)

M2z zOH{ ~M(OH)z (3)

M(OH)z ~M(OH)z
ads (4)

M(OH)z {
ads z2e ~MzOH
{
(5)
where M indicates a nickel or cobalt atom. The
deposition of cobalt is promoted by the surface
enrichment of adsorbed metal hydroxide cations
(Co(OH)zads). The cobalt content has a significant
influence on surface morphology, phase structure, 9 Composition of Ni–Co alloys as a function of ratio of
microhardness, wear and corrosion resistance, which Co2z/(Co2zzNi2z) in electrolytes3

108 Transactions of the IMF 2015 VOL 93 NO 2


Ma et al. Electrodeposition of nanocrystalline nickel–cobalt binary alloy coatings

12 Effect of peak current density on grain size of Ni–Co


coatings deposited by pulsed current plating9

10 Effect of cathodic current density on composition of with the coating obtained from additive-free baths, the
Ni–Co deposits produced by jet electrodeposition a addition of saccharin resulted in a slight reduction of
without saccharin and b with 2.5 g dm23 saccharin14 cobalt content due to the increasing deposition over-
potential as shown in Fig. 10, which is in favour of
reduced to 10 nm at l60 A dm22. The reduction of grain activation-controlled Ni deposition.3 Li et al.19 also
size can be attributed to the higher overpotential gk, reported that by increasing the concentration of
which is related to the probability of crystalline nucleus saccharin from 3 to 5 g dm23 in electrolytes containing
formation, W expressed by17 6 g dm23 cobalt sulphate, the cobalt content slightly
decreased from 8.5 to 6.1 wt-%.
b The deposits from saccharin-containing baths inevi-
W ~B exp{ (6)
gk tably contain certain amounts of sulphur impurities,19,20
where B and b are constants. The overpotential increases which may lead to poor ductility21 and corrosion
with the increasing current density as described by the resistance.22 Saccharin has a significant effect on the
cathodic Tafel equation9 reduction of grain size. The grain size of 5 wt-% Co–Ni
alloys decreased from 44 nm for the coating plated from
gk ~a{blog j (7) the additive-free bath to 22 nm by adding 1 g dm23
saccharin.13 At the same time, the texture changed from
where a and b are constants and j is the current density.
(200) to (111) and the surface was much smoother.
Increasing current density leads to increasing over-
These changes led to higher microhardness and
potential, which promotes the formation of a crystalline
improved anti-corrosion properties of coatings. The
nucleus and reduction in grain size.
addition of sodium dodecyl sulphate led to a compact
Electrolyte additives surface morphology, which improved the corrosion
Organic compounds (e.g. saccharin, sodium dodecyl resistance. Marikkannu et al.18 investigated the role of
sulphate and 2-butin-1,4-diol in small concentrations additives in Ni–Co deposits from acetate baths contain-
can be added to Ni–Co plating baths to modify the ing single additives, including saccharin, dextrin, cou-
process of electrocrystallisation for a variety of pur- marin, formaldehyde, glycine and crotonaldehyde.
poses, including obtaining a mirror-like surface, increas- Some corrosion characteristics of Ni–Co deposits in
ing hardness and improving corrosion resistance.18 The 5 wt/vol.-% NaCl at 25uC are shown in Table 1.18 The
effects of peak current density on the grain size of Ni–Co corrosion resistance of Ni–Co deposits with a cobalt
deposits produced by pulsed electrodeposition9 on the content of 42–47% in 5 wt./vol.-% NaCl solution was
composition, grain size, microstructure and properties improved by adding these additives. Furthermore, the
of Ni–Co deposits have been investigated. Compared microhardness of these coatings was higher than that of
deposits from the additive-free bath. Saccharin and
formaldehyde produced a mirror finish deposit.
Recent studies have examined saccharin and 2-butin-
1,4-diol as bath additives, in a modified Watts solution,
to electrodeposited high-quality nickel–cobalt alloys
(78¡2 at.-%Co) as a protective coating on steel
substrates for tribological applications.23 These addi-
tives allow controlled electrodeposition of nanocrystal-
line coatings. The surface morphology, grain size,
crystalline texture, hardness and performance against
stainless steel in a reciprocating ball-on-disc tribometer
were studied. Microstress in the coating could be
manipulated from tensile to compressive and the texture
could be modified from hexagonal closed-packed
structure (1010)
- structure to a face centred cubic (111)
11 Effect of peak current density on composition of Ni– or a hexagonal closed-packed (0002) structure. The
Co deposits produced by pulsed current plating9 effect of absorbed species on the nanocrystalline coating

Transactions of the IMF 2015 VOL 93 NO 2 109


Ma et al. Electrodeposition of nanocrystalline nickel–cobalt binary alloy coatings

13 Variation of internal stress with cobalt content in Ni–


14 Distribution of cobalt content with thickness of graded
Co alloys26
Ni–Co nanocrystalline alloys26

is explained via grain size and texture analyses. The Approaches to reduce internal stress
coating from the bath with an optimised additive Wang et al.26 reported that the internal stress of Ni–Co
content showed a high hardness (500 HV) due to its electrodeposits depended on the cobalt content. As
reduced grain size (11¡2 nm) and improved tribological shown in Fig. 13, the internal stress gradually decreased
properties due to the high proportion of hcp structure. with increasing cobalt content. It changed from com-
pressive to tensile as the cobalt content exceeded 49 wt-%.
Bath agitation and temperature A very high tensile stress in the range of 40–70 MPa was
The common methods used for bath agitation include obtained for the cobalt-rich coatings. The high internal
air agitation, overhead impeller,24 magnetic stirring,3 stress can result in lower ductility and adhesion between
high speed jet3 and peristaltic pumping25 in order to coatings and substrates,27 which restricts their applica-
maintain the flow of electrolyte at a constant and stable tions for anti-wear and anti-corrosion purposes. In
level. An increase in bath agitation results in a higher order to decrease the internal stress, the Ni–Co coating
cobalt content in deposits due to the reduction of the with continuously graded composition and structure
diffusion layer thickness accompanied by the rising were produced by controlling the cobalt ion concentra-
metal ion concentration in it. As discussed above, tion in baths during the electrodeposition process.26 As
diffusion controlled cobalt is preferentially deposited.14 shown in Fig. 14, the composition gradually changes
In addition, the agitation can increase the limiting from the Ni-rich region to the Co-rich region as a
current density jL by reducing the thickness of the function of distance from the interface between the
Nernst diffusion layer, d in the equation17 coating and the AISI-1045 steel substrate. The phase
jL ~zFD(cb {cs )=d (8) structure gradually changed from fcc to hcp with
thickness due to the increase in cobalt content. As
where F is the Faraday constant, D is the diffusion shown in Fig. 13, the internal stress of the graded Ni–Co
coefficient of the metal ion, cb is the metal ion coating was reduced to a much lower level, around
concentration in the bulk solution and cs is the metal 5 MPa. The graded coating with the microhardness of
ion concentration at the electrode surface. The increas- 550 HV exhibited a much lower coefficient of friction
ing current density results in higher overpotential to and improved wear resistance compared with ungraded
produce coatings with finer grain size. Qiao et al.14 nickel-rich coatings.
reported that with the increasing temperature from 30 to Based on this, Wang et al.28 developed nanocrystalline
50uC, the cobalt content in Ni–Co deposits decreased Ni–Co/CoO graded coatings by a subsequent cyclic
from 77 to 69 wt-%, since a higher temperature favoured thermal oxidation at 300uC and quenching. As shown in
activation-controlled Ni deposition. Fig. 15, the graded Ni–Co/CoO coating consisted of six

Table 1 Corrosion characteristics of Ni–Co deposits plated from acetate baths having different additives in 5% NaCl at
25uC18

Additives Concentration/g dm23 Ecorr/mV(SCE) jcorr/61024 A cm22 Microhardness/HV

(None) 0 2449 3.5 349


Saccharin 2 2501 1.7 385
Dextrin 3 2648 2.5 365
Coumarin 1 2684 2.3 375
Formaldehyde 2 mL dm23 2460 1.9 379
Glycine 2 2593 3.1 371
Crotonaldehyde 2 mL dm23 2585 2.9 367
1,4-butyne diol 1–2 mL dm23 2668 49.8 372
1,3-naphthalene sulphonic acid 3 2457 2.0 360

110 Transactions of the IMF 2015 VOL 93 NO 2


Ma et al. Electrodeposition of nanocrystalline nickel–cobalt binary alloy coatings

15 a SEM image of cross-section of graded Co–Ni/CoO coating. Corresponding b nickel and c cobalt element distribution
maps across coating28

sublayers of controlled composition varying from 0 to Among the many electrode materials suggested for
81 wt-%Co. The Co level in the coating was controlled advanced batteries and supercapacitors, a recent paper
by careful control of post-plating temperature and time from the authors’ laboratory has demonstrated the
during heating then quenching. The coating was electrodeposition of nanostructured Ni–Co layers.35
approximately 100 mm thick with an oxide layer of These studies used an electrolyte of an aqueous solution
3–4 mm. Compared to as-deposited Ni–Co graded coat- of 0.84M NiCl2 and 0.20M CoCl2, at 317 K with
ings, the graded Ni–Co/CoO composite coating exhib- templated deposition through a mesoporous, lyotropic
ited significantly enhanced wear resistance under dry liquid crystal phase surfactant (cetyltrimethylammo-
sliding conditions and improved corrosion resistance in nium bromide) layer having ‘toothpaste’ consistency
both 10 wt-% NaOH and 3.5 wt-% NaCl solutions due at a typical current density of 3 mA cm22. Energy
to the formation of the dense and protective oxide layer. dispersive X-ray spectroscopy analysis showed that the
atomic ratio of Ni/Co (typically 0.20 to 0.26) varied only
Electrochemical applications of slightly with current density. The high surface area Ni/
Co alloy layers with a thickness of 2 mm in 1 mol dm23
electrodeposited Ni–Co alloy NaOH were imaged by helium ion microscopy and
nanostructured coatings could store y1 C cm22 via the Ni(OH)2/NiOOH
As indicated in previous sections, major driving forces couple. This charge could be cycled .20 times.
for the development of Ni–Co electrodeposits have been
increasingly arduous service requirements for corrosion Summary
resistant (the section on ‘Corrosion behaviour’) and
wear resistant (the section on ‘Tribological behaviour’) Ni–Co electrodeposition is a preferred substitute for
coatings, often as an alternative to nickel or a possible coatings including hard chromium and nickel from the
replacement for chromium coatings. The attractions of a point of view of the balance between cost, speed and
directly produced nanostructured alloy coating having a coating service performance. Recent studies on electro-
high surface area and catalytic activity for oxygen and deposited nickel, cobalt, Ni–Co binary alloys and,
hydrogen electrodes has additionally given rise to use in briefly, Co–Ni–P ternary coatings have been sum-
water electrolysis and energy conversion in advanced marised. Further research is needed in order to better
batteries and supercapacitors, which are briefly consid- understand the microstructure and properties of Ni–Co
ered in this section. coatings and to develop further coatings with improved
Electrodeposited Ni–Co coatings may find use as wear and corrosion resistance to replace hard chro-
electrocatalytic coatings for hydrogen or oxygen evolu- mium. Rising environmental pressures on nickel solu-
tion in advanced water electrolysers.29 Hydrogen evolu- tions and wastes (and to cobalt in the longer term) mean
tion catalysts are commonly ‘activated’ nickel, which that Ni–Co coatings may receive more limitations on
consists of finely divided nickel, or Ni-alloy deposits;30 their use, despite their tribological and corrosion
nanostructured Ni–Co is a possible material although prevention advantages.
Ni–Mn is often superior. In the case of oxygen electrodes, Some key remaining questions concerning nanostruc-
a facile route to such coatings is to carefully anodise tured Ni–Co alloy electrodeposits can be summarised.
electrodeposited Ni–Co alloys to form mixed metal 1. Recent findings have shown that electrodeposited
oxides, such as NiCo2O4 spinels which are established nanocrystalline cobalt-rich Ni–Co alloys having an hcp
oxygen evolution catalysts.31 High surface area, nickel- structure exhibited better tribological properties than Ni
based oxygen evolution electrodes, including Ni and nickel-rich Ni–Co coatings.12 This may be asso-
Co(OH)2, have been examined at high current densities ciated with the participation of tribofilms and debris.
and elevated temperature, 353 K in 4M KOH.32 However, no studies on the role of tribofilms and the
NiCo2O4 spinels have also been suggested as bifunctional microstructure of debris have been reported. More
(oxygen evolution and oxygen reduction catalysts in 4M experimental work is needed to fully understand the
KOH electrolytes at 333 K for secondary Zn-air flow tribological behaviour of Ni–Co coatings of controlled
batteries33 and as electrode surfaces for supercapacitors.34 cobalt content. In addition, the corrosion resistance of

Transactions of the IMF 2015 VOL 93 NO 2 111


Ma et al. Electrodeposition of nanocrystalline nickel–cobalt binary alloy coatings

Ni–Co alloys in 3.5% NaCl solutions needs to be closely 12. C. Ma, S. C. Wang, L. P. Wang, F. C. Walsh and R. J. K. Wood:
‘The role of a tribofilm and wear debris in the tribological
compared to that of hard chromium, particularly as
behaviour of nanocrystalline Ni-Co electrodeposits’, Wear, 2013,
nanostructured Ni–Co deposits can be considered as a 306, 296–303.
possible replacement for hard chromium layers in some 13. S. Hassani, K. Raeissi, M. Azzi, D. Li, M. Golozar and J. Szpunar:
applications. ‘Improving the corrosion and tribocorrosion resistance of Ni–Co
2. It has been reported that cobalt-rich nanocrystal- nanocrystalline coatings in NaOH solution’, Corros. Sci., 2009, 51,
2371–2379.
line coatings with improved tribological properties were 14. G. Qiao, T. Jing, N.Wang, Y. Gao, X. Zhao, J. Zhou and W.
difficult to electrodeposit onto steel due to high internal Wang: ‘High-speed jet electrodeposition and microstructure of
stress. Current methods (such as graded coatings and nanocrystalline Ni–Co alloys’, Electrochim. Acta, 2005, 51, 85–92.
applying pulse current waveforms) have limitations 15. A. Bai and C. Hu: ‘Composition controlling of Co-Ni and Fe-Co
including complex equipment needs and high cost. alloys using pulse reverse electroplating through means of
experimental strategies’, Electrochim. Acta, 2005, 50, 1335–1345.
Moreover, the microhardness needs to be further 16. C. Fan and D. Piron: ‘Study of anomalous nickel-cobalt electro-
increased. In recent work, Ni–Co coatings having a deposition with different electrolytes and current densities’,
high cobalt content have been realised on steel sub- Electrochim. Acta, 1996, 41, 1713–1719.
strates by selecting and optimising electrolyte additives.23 17. C. Karakus and D.-T. Chin: ‘Metal distribution in jet plating’, J.
The effect of additives on the microstructure and pro- Electrochem. Soc., 1994, 141, 691–697.
18. K. Marikkannu, G. P. Kalaignan and T. Vasudevan: ‘The role of
perties deserves extensive study. additives in the electrodeposition of nickel–cobalt alloy from
3. Although the effect of baths with different nickel acetate electrolyte’, J. Alloys Compd, 2007, 438, 332–336.
salts on mechanical properties has been studied on 19. Y. Li, H. Jiang, D. Wang and H. Ge: ‘Effects of saccharin and cobalt
coarse-grained (grain size of the order of microns) nickel concentration in electrolytic solution on microhardness of nanocrys-
talline Ni–Co alloys’, Surf. Coat. Technol., 2008, 202, 4952–4956.
coatings, few comparable investigations on nanocrystal-
20. A. El-Sherik and U. Erb: ‘Synthesis of bulk nanocrystalline nickel
line Ni–Co coatings have been carried out. Numerous by pulsed electrodeposition’, J. Mater. Sci, 1995, 30, 5743–5749.
mechanisms have been proposed to explain the deviation 21. W. Yin, S. Whang and R. Mirshams: ‘Effect of interstitials on
from the conventional Hall–Petch relationship for tensile strength and creep in nanostructured Ni’, Acta Mater., 2005,
nanocrystalline materials when the grain size is reduced 53, 383–392.
22. R. Mishra and R. Balasubramaniam: ‘Effect of nanocrystalline
to a threshold value. However, none of them focused on grain size on the electrochemical and corrosion behaviour of
the influence of surface pores, which are common for nickel’, Corros. Sci., 2004, 46, 3019–3029.
electrodeposited nanocrystalline coatings. Deviation 23. C. Ma, S. C. Wang, C. T. J. Low, L. P. Wang and F. C. Walsh:
from the Hall–Petch relationship may derive from the ‘The effects of additives on the microstructure and properties of
porosity36 and this remains to be clarified. electrodeposited nanocrystalline Ni-Co coatings with high cobalt
content’, Trans. IMF, 2014, 92, 189–195.
24. M. Schlesinger: ‘Modern electroplating’, 5th edn, 513–516; 2010,
Acknowledgement Hoboken, NJ, John Wiley & Sons. Inc.
25. L. Burzynska and E. Rudnik: ‘The influence of electrolysis
The authors gratefully acknowledge financial support parameters on the composition and morphology of Co–Ni alloys’,
from ICUK which contributed to the PhD studies of CM. Hydrometallurgy, 2000, 54, 133–149.
26. L. Wang, Y. Gao, Q. Xue, H. Liu and T. Xu: ‘Graded composition
and structure in nanocrystalline Ni–Co alloys for decreasing
References internal stress and improving tribological properties’, J. Phys. D:
1. C. Ma, S. C. Wang and F. C. Walsh: ‘The electrodeposition of Appl. Phys., 2005, 38, 1318–1325.
nanocrystalline nickel and cobalt coatings: A review’, Trans. IMF, 27. J. Deng and M. Braun: ‘DLC multilayer coatings for wear
2015, 93, (1), in press. protection’, Diam. Relat. Mater., 1995, 4, 936–943.
2. C. Ma, S. C. Wang and F. C. Walsh: ‘The electrodeposition of 28. L. Wang, J. Zhang, Z. Zeng, Y. Lin, L. Hu and Q. Xue:
nanocrystalline nickel-cobalt-phosphorus ternary coatings: a ‘Fabrication of a nanocrystalline Ni–Co/CoO functionally graded
review’, submitted to Trans. IMF, 2014. layer with excellent electrochemical corrosion and tribological
3. L. Wang, Y. Gao, Q. Xue, H. Liu and T. Xu: ‘Microstructure and performance’, Nanotechnology, 2006, 17, 4614–4623.
tribological properties of electrodeposited Ni–Co alloy deposits’, 29. D. Pletcher and X. Li: ‘Prospects for alkaline zero gap water
Appl. Surf. Sci., 2005, 242, 326–332. electrolysers for hydrogen production’, Int. J. Hydrogen Energy,
4. B. Tury, M. Lakatos-Varsányi and S. Roy: ‘Ni–Co alloys plated by 2011, 36, 15089–15104.
pulse currents’, Surf. Coat. Technol., 2006, 200, 6713–6717. 30. D. Pletcher, X. Li and S. Wang: ‘A comparison of cathodes for
5. N. V. Myung and K. Nobe: ‘Electrodeposited iron group thin-film zero gap alkaline water electrolysers for hydrogen production’, Int.
alloys: structure property relationships’, J. Electrochem. Soc., 2001, J. Hydrogen Energy, 2012, 37, 7429–7435.
148, C136–C144. 31. M. R. Tarasevich and B. N. Efremov. ‘Properties of spinel-type
6. D. Golodnitsky, Y. Rosenberg and A. Ulus: ‘The role of anion oxide electrodes’, in ‘Electrodes of conductive metallic oxides’,
additives in the electrodeposition of nickel–cobalt alloys from (ed. S. Trasatti), 221–259; 1980, Elsevier, Amsterdam, New York.
sulfamate electrolyte’, Electrochim. Acta, 2002, 47, 2707–2714. 32. X. Li, F. C. Walsh and D. Pletcher: ‘Nickel based electrocatalysts
7. M. Srivastava, V. Ezhil Selvi, V. William Grips and K. Rajam: for oxygen evolution in high current density alkaline water
‘Corrosion resistance and microstructure of electrodeposited electrolysers’, Phys. Chem. Chem. Phys., 2013, 13, 1162–1167.
nickel–cobalt alloy coatings’, Surf. Coat. Technol., 2006, 201, 33. S. W. Price, S. J. Thompson, X. Li, S. F. Gorman, D. Pletcher,
3051–3060. A. E. Russell, F. C. Walsh and R. G. A. Wills: ‘The fabrication of a
8. M. Gutkin and I. Ovid’ko: ‘Plastic deformation in nanocrystalline bifunctional oxygen electrode without carbon components for
materials’, 7; 2004, Berlin, Springer. alkaline secondary batteries’, J. Power Sources, 2014, 259, 43–49.
9. Y. Li, H. Jiang, W. Huang and H. Tian: ‘Effects of peak current 34. M. Liu, L. Kong, C. Lu, X. Li, Y. Luo, L. Kang, X. Li and F. C.
density on the mechanical properties of nanocrystalline Ni–Co Walsh: ‘A sol-gel process for the synthesis of NiCo2O4 having
alloys produced by pulse electrodeposition’, Appl. Surf. Sci., 2008, improved specific capacitance and cycle stability for electrochemi-
254, 6865–6869. cal capacitors’, J. Electrochem. Soc., 2012, 159, A1262–A1266.
10. W. Brainard and D. Buckley: ‘Preliminary friction and wear studies 35. P. N. Bartlett, D. Pletcher, T. F. Esterle and C. T. J. Low: ‘The
of cobalt rhenium solid solution alloy in air and in vacuum’, NASA deposition of mesoporous Ni/Co alloy using cetyltrimethylammo-
Report NASA-TN-D-6165, 1971. nium bromide as the surfactant in the lyotropic liquid crystalline
11. I. Inman: ‘Compacted oxide layer formation under conditions of phase bath’, J. Electroanal. Chem., 2013, 688, 232–236.
limited debris retention at the wear interface during high 36. C. Ma, S. C. Wang, R. J. K. Wood, Q. Luo, C. T. J. Low, F. C.
temperature sliding wear of superalloy’, PhD dissertation, Walsh: ‘The hardness of porous nanocrystalline Co-Ni electro-
University of Northumbria at Newcastle, 2003. deposits’, Met. Mater. Int., 2013, 19, 1187–1192.

112 Transactions of the IMF 2015 VOL 93 NO 2

You might also like