Anda di halaman 1dari 7

Applied Surface Science 242 (2005) 326–332

www.elsevier.com/locate/apsusc

Microstructure and tribological properties of


electrodeposited Ni–Co alloy deposits
Liping Wanga,b, Yan Gaoa,b, Qunji Xuea, Huiwen Liua, Tao Xua,*
a
State Key Laboratory of Solid Lubrication, Lanzhou Institute of Chemical Physics,
Chinese Academy of Sciences, Lanzhou 730000, PR China
b
Graduate School of the Chinese Academy of Sciences, Beijng 100039, PR China
Received in revised form 24 August 2004; accepted 30 August 2004
Available online 20 October 2004

Abstract

Ni–Co alloys with different compositions and microstructures were produced by electrodeposition. The effects of Co content
on the composition, surface morphology, phase structure, hardness and tribological properties of Ni–Co alloys were investigated
systemically. Results showed that the morphology and grain size of alloys are mainly influenced by the Co content and the phase
structure of Ni–Co alloys gradually changed from fcc into hcp structure with the increase of Co content. The hardness of Ni–Co
alloys with a maximum around 49 wt.% Co followed the Hall–Petch effect. It was found that the improvement of wear resistance
of Ni-rich alloys with hardness increase fits Archard’s law. In addition, the Co-rich alloys exhibited much lower friction
coefficient and higher wear resistance when compared with Ni-rich alloys. It has been concluded that hcp crystal structure in Co-
rich alloys contributed to the remarkable friction–reduction effect and better anti-wear performance under the dry sliding wear
conditions.
# 2004 Elsevier B.V. All rights reserved.

Keywords: Ni–Co alloy; Electrodeposition; Structure; Tribological properties

1. Introduction activity [1–5]. Additionally, the use of Ni–Co alloys


has been extended to the production of three-
Ni–Co alloys have been investigated as important dimensional, complex-shaped finished components
engineering materials for several decades because of by the electroforming technique [6,7]. The investiga-
their unique properties, such as high-strength, good tions on the electrodeposited Ni–Co alloys have
wear resistance, heat-conductive, electrocatalytic shown that their microstructure and properties were
found to depend strongly on the Co content, which can
* Corresponding author. Tel.: +86 931 496 8169;
be controlled by the experimental parameters, such as
fax: +86 931 496 8169. bath composition, temperature, pH value, and current
E-mail address: lpwang@lsl.ac.cn (T. Xu). density, etc. [1,3,8]. The effects of plating parameters

0169-4332/$ – see front matter # 2004 Elsevier B.V. All rights reserved.
doi:10.1016/j.apsusc.2004.08.033
L. Wang et al. / Applied Surface Science 242 (2005) 326–332 327

on the composition and morphology of Ni–Co The surface morphology and microstructure of the
deposits were compared in many literatures [4,5,9]. alloy deposits were investigated using a JSM-5600Lv
Golodnitsky et al. recently studies the effects of Co scanning electron microscopy (SEM). The composi-
content on the tensile strength, internal stress and tions of Ni–Co alloys were determined with energy
high-temperature oxidation of Ni–Co alloys [3]. Their dispersive X-ray spectroscopy (EDS) analysis tool
activities for the oxygen evolution reaction and attached to SEM. The crystal structure and phase
hydrogen evolution reaction were also studied on composition of alloy deposits were studied by X-ray
electrodeposited Ni–Co ultramicroelectrodes [10,11]. diffraction (XRD). Microhardness of the deposits was
Moreover, much interest is focused on the magnetic determined using a Vicker’s microhardness indenter
properties of Ni–Co alloys due to the application of with a load of 50 g for 10 s, indentations were made on
these alloys in various magnetic devices, especially in the 50 mm thick deposits. The final value quoted for
microsystem technology for manufacture of sensors, the hardness of a deposit was the average of 10
actuators and inductors [12,13]. It is reported that the measurements.
magnetic properties of Ni–Co alloy are greatly The tribological behavior was tested on a
influenced by the composition and phase structure reciprocating ball-on-disk UMT-2MT tribometer
of Ni–Co alloy [14]. Unfortunately, there are very (Center for tribology, Inc., California, USA) at room
limited studies focused on the friction and wear temperature with a relative humidity of 45–55% under
properties of Ni–Co alloys as a function of their dry sliding conditions. AISI-52100 stainless steel ball
microstructure and composition. (diameter 4 mm with hardness of RC 62) was used
In the present paper, Ni–Co alloys with different Co as the counter body; all tests were performed under a
content were electrodeposited on AISI-1045 steel load of 3 N with a sliding speed of 55 mm s1. The
substrates. The composition, microstructure, mechan- friction coefficient and sliding time were recorded
ical, and tribological properties of Ni–Co alloys were automatically during the test. The wear volume loss
compared systemically in order to specifically was measured using a surface profilometer, wear rates
correlate the structure and tribological properties of of all the alloy deposits were calculated using the
Ni–Co alloys. equation of K ¼ V=SF, where V is the wear volume
loss in mm3, S the total sliding distance in m and F the
normal load in N.
2. Experimental

Ni–Co alloys were electrodeposited from a typical 3. Results and discussion


Watts-type electrolyte, containing Nickel sulfate
(200 g/l), sodium chloride (20 g/l), boric-acid (30 g/ 3.1. Composition of Ni–Co alloys
l), sodium lauryl sulfate (0.1 g/l) and cobalt sulfate (0–
80 g/l). In addition, pure Ni was also produced for The dependence of the composition of Ni–Co alloys
comparison purpose. The Ni–Co alloys were depos- on the concentration of Co2+ ions in the electrolyte at a
ited on AISI-1045 steel substrates by choosing a fixed concentration of Ni2+ ions is presented in Fig. 1. It
current density of 3 A/dm2 at a bath temperature of is clearly observed that the Co content in alloy deposits
45 8C. The anode was a pure Ni plate. The pH of the increased gradually with the increase of Co2+ con-
bath was kept at 4.0 adjusted by ammonia water or centration in the electrolyte. Note that the percentage of
dilute sulfuric-acid. Before deposition, the substrates Co in the alloys was always higher than in the
were mechanically polished to a 0.10–0.12 mm electrolyte in agreement with [3,5], which is confirmed
surface finish, the substrate was then degreased in by the anomalous codeposition of Ni–Co alloy. Namely,
acetone with ultrasonic cleaning for 5 min, rinsed in the less noble metal (Co) is preferentially deposited. A
the running water to remove contamination on the generally accepted explanation for these anomalous
substrate surface. After than, the steel substrates were phenomena was the change of the near-electrode pH,
activated for 20 s in the 20 vol.% HCl solution, and the formation of metal hydroxyl and their competitive
finally rinsed with distilled water. adsorption [15,16].
328 L. Wang et al. / Applied Surface Science 242 (2005) 326–332

Fig. 1. The alloy compositions as a function of Co2+ concentrations


in the baths.

3.2. Morphology and phase structure of Ni–Co alloys

Typical surface morphologies of Ni–Co alloys with


different Co content are shown in Fig. 2b–f,
respectively. Fig. 2a shows a typical morphology of
a Watt Ni deposit, which has relatively large grain size
(3–10 mm) and showed polyhedral crystallites.
Sequentially increasing Co content from 7 to
49 wt.% (Fig. 2b–d) results in a gradual decrease in
the grain size of the Ni–Co alloy down to a sub-micron Fig. 2. SEM morphologies of Ni–Co alloy deposits with their Co
grain size. When the Co content reached the 49 wt.%, contents of (a) 0 wt.%, (b) 7 wt.%, (c) 27 wt.%, (d) 49 wt.%, (e)
close observation of SEM morphology at high- 66 wt.%, (f) 81 wt.%, (g) high-magnification of Ni–49 wt.% Co
magnification (Fig. 2g) revealed that the Ni–Co alloys alloy.
have spherical cluster surface piled with a large
number of equally sized grains with spherical-shape.
At above 49 wt.% Co, the grain size of Ni–Co largest grain size of pure Ni. With the codeposition of
deposits, however, increased with the increase of Co Co, the Ni–Co solid solution was formed. As can be
content in alloys. When increasing Co content up to seen in Fig. 3b–f, both the crystal structure and phase
81 wt.%, the morphology of the Ni–Co alloys changes composition are mainly dependent on the Co contents
dramatically, and with less compact structure, the in alloys. For the Ni-rich alloys with Co content lower
Ni–Co alloy showed a rather regularly branched than 49 wt.%, the Ni–Co alloys show complete fcc
structure with extended acicular 3–6 mm length phase structure, which is in agreement with previously
crystallites (Fig. 2f). reported results [2,3]. Furthermore, the (1 1 1) growth
The phase composition and structure of pure Ni and orientation gradually increased with the increase of Co
Ni–Co alloys with different Co contents were content, and the FWHM of the Bragg line for the
investigated using XRD shown in Fig. 3. As can be (1 1 1) peak also increased correspondingly, which is
seen from Fig. 3a, the pure Ni deposit exhibits face- in accordance with the gradual reduction of grain size
centered cubic (fcc) lattice with remarkable (2 0 0) when increasing the Co content from 0 to 49 wt.% in
growth orientation, which can be attributed to the Ni–Co alloys as shown in Fig. 2b–d. Moreover, when
L. Wang et al. / Applied Surface Science 242 (2005) 326–332 329

81 wt.%, as shown in Fig. 3f that a very strong hcp


(0 0 2) texture with pronounced (1 0 0) and (1 1 0)
peaks were observed, which is commonly observed in
both conventionally electrodeposited Co and nano-
crystalline Co [17,18]. Therefore, it can be concluded
that the phase structure of Ni–Co alloys gradually
changed from fcc into hcp with the increase of Co
content as shown in Fig. 3.

3.3. Microhardness of Ni–Co alloy deposits

Fig. 4a presented the microhardness of Ni–Co


alloys as a function of Co content in alloys. It is clearly
that microhardness of Ni–Co alloys increased initially
with Co content varying from 0 to approximately
49 wt.%, and then gradually decreased as Co content
increased further above 49 wt.%. The explanation to
this gradual reduction of microhardness is the gradual
increase of grain size with the increase of Co content
in Co-rich alloys as shown in Fig. 2e–f. Note that the
microcrystalline Ni–Co alloys show the maximum
hardness at approximately 49 wt.% Co, which can be
associated with the smallest grain size as mentioned in
Fig. 3. XRD patterns of Ni–Co alloy deposits with their Co contents
microstructure analysis sector.
of (a) 0 wt.%, (b) 7 wt.%, (c) 27 wt.%, (d) 49 wt.%, (e) 66 wt.%, Normally, strengthening of polycrystalline materi-
(f) 81 wt.%. als by grain size refinement is technologically
attractive because it generally does not adversely
affect ductility and toughness [19], which can be
the Co content was increased to 66 wt.%, the presence represented by the classical Hall–Petch effect:
of (1 0 0) peak demonstrated the initial formation of a
hexagonal close packed (hcp) lattice, indicating that H ¼ H0 þ kd 0:5 (1)
the crystal structure of the Ni–Co alloy changed from where H0 is hardness constant, k constant, and d
complete fcc lattice into a mixed (majority of fcc) + diameter of grain. The hardness change with average
(minority of hcp) phase as shown in Fig. 3e. At above grain size (d) of the Ni–Co deposits is shown in Fig. 4b

Fig. 4. Microhardness as function of Co content (a) and d0.5 (b) of the Ni–Co deposits.
330 L. Wang et al. / Applied Surface Science 242 (2005) 326–332

in the form of a Hall–Petch plot. It is obvious that the


Ni–Co alloy deposits exhibit a nearly constant Hall–
Petch gradient; such a relationship has also been
observed on pure Ni, pure Co and pure Zn from other
studies [20,21].

3.4. Friction and wear properties

The effect of Co content on friction coefficient of Ni–


Co alloys were shown in Fig. 5. It is observed that the
friction coefficient of pure Ni and Ni–Co alloys with Co
content lower than 49 wt.% (Ni-rich alloys) were quite Fig. 6. The comparison of friction coefficients vs. sliding time
close. With the further increase of Co content, the Co- between Ni-rich and Co-rich alloy deposits.
rich alloys showed excellent friction–reduction beha-
vior. The Co-rich alloy deposit with Co content higher the increase of Co content can be associated with the
than 81 wt.% exhibited the smallest friction coefficient change of crystal structure from fcc to hcp crystal phase.
(more than two times lower than Ni and Ni-rich alloys), The variation of the wear rates of Ni–Co alloys as a
followed by Ni–66 wt.% Co alloy (a litter lower than Ni function of Co content and microhardness of alloys are
and Ni-rich alloys) under identical wear test conditions. shown in Fig. 7. It is observed that all Ni–Co alloy
In addition, the friction coefficients of Co-rich alloys deposits in this study have lower wear rates when
were much more stable than that of Ni-rich alloy compared with pure Ni deposit. Moreover, the wear
deposits (see Fig. 6). Combined with the XRD analysis, rate of Ni–Co alloys slowly decreased with the
the close friction coefficient for Ni and Ni-rich alloy can increase of Co content from 6 to 49 wt.%. It is clear
be attributed to the same fcc crystal structure they have. that when the Co content is lower than 49 wt.%, the
In case of Ni–66 wt.% Co alloy, a mixed fcc/hcp phase gradual decrease of wear rates with the increase of Co
with smaller ratio of hcp phase structure led to the content was attributed to the microhardness increase
gradual reduction of friction coefficient. Furthermore, as from 315 to 462 HV. Above improvement of wear
for the Ni–81 wt.% Co alloy, dramatic reduction of resistance with hardness increase, in this study due to
friction coefficient was observed due to the higher ratio the grain size reduction, could be expressed using
of hcp phase structure. Hence, we can conclude that the Archard’s law mostly used in adhesive wear condi-
reduction in friction coefficient of Co-rich alloys with tions [22,23], since the wear mechanism of Ni and

Fig. 5. Friction coefficient as function of Co content in the Ni–Co Fig. 7. Wear rates as function of Co content in the Ni–Co alloy
alloy deposits. deposits.
L. Wang et al. / Applied Surface Science 242 (2005) 326–332 331

Ni-rich deposits is mostly the adhesive wear as higher friction coefficient of Ni-rich alloys due to
evidenced by SEM morphology of worn surface in fcc phase structure led to the more wear loss. More
Fig. 8a and b. Thus, the Archard’s law can be important is the fact that the Co-rich alloys exhibited
expressed as: excellent wear resistance and anti-friction behavior.
LN The difference in the wear behavior of Ni–Co
Q¼K (2) alloys can be further verified by the worn surface
H
morphologies of Ni-rich and Co-rich deposit as shown
where Q is the volumetric wear loss, N the applied in Fig. 8a–c. For the pure Ni deposit and Ni-rich alloy
load, L the total sliding distance, K the wear coefficient with completely fcc crystal structure, the wear track
and H the hardness of the wear surface. Under the (Fig. 8a and b) shows the larger extent of adhesion
same wear conditions, the wear rate is proportional to wear and severe deformation in the sliding direction
the inverse microhardness of materials. The data of under the combined stresses of compression and shear,
wear rate for microcrystalline Ni–Co alloys with Co which results in larger wear rate of pure Ni and Ni-rich
content lower than 49 wt.% fit Archard’s law very alloys. Furthermore, larger tendency for plastic
well. However, with further increase in Co content deformation, this in turn increased the probability
above 49 wt.%, the wear rates of Co-rich alloys of formation of asperity junctions resulting in higher
decreased rapidly in spite of the fact that the hardness and unstable friction coefficient for Ni and Ni-rich
also decreased. The wear rate of Co-rich alloy with alloys. Compared with pure Ni and Ni-rich alloys, a
approximately 81 wt.% Co content is more than one densification of the worn surface of Co-rich alloy
order of magnitude lower than that of pure Ni and Ni- seems to take place, the worn surface of Co-rich alloy
rich alloys. This reverse-Archard law may be caused with hcp crystal structure revealed slight adhesion
by special hcp crystal structure of Co-rich alloys. This wear and rather smooth surface with smaller damaged
agree well with the reduction in friction coefficient for regions, only some light grooves and scars are noted
Co-rich alloys, namely, the lower and stable friction on the worn surface (Fig. 8c). This resulted in the
coefficient of Co-rich alloys caused by hcp phase better wear resistance of Co-rich alloy than Ni-rich
structure resulted in the less wear loss, while the alloys. That is also the reason why the friction

Fig. 8. Worn surface of Ni–Co alloy deposits: (a) 0% Co; (b) 27% Co; (c) 81% Co.
332 L. Wang et al. / Applied Surface Science 242 (2005) 326–332

coefficients of Co-rich alloys were much more stable Acknowledgements


and more than two times lower than that of Ni and
Ni-rich alloys. It is evident that the high the amount of The authors gratefully acknowledge the National
hcp phase structure, the better the friction and wear Natural Science Foundation of China (Grant No
behavior will be. Above evidence suggests that the 50172052, 50271080 and 50323007), the 863 Program
crystal structure is indeed a dominant factor, which of China (No. 2003AA305670), and ‘Top Hundred
influences the friction and wear behavior of Ni–Co Talents Program’ of Chinese Academy of Sciences for
alloys. Hence, it clearly demonstrates that hcp crystal financial support of this research work.
structure in Ni–Co alloys contributed to the remark-
able friction–reduction effect and better anti-wear
performance of Co-rich alloys. References

[1] A.N. Correia, S.A.S. Machado, Electrochim. Acta 45 (2000)


4. Conclusions 1733.
[2] V.B. Singh, V.N. Singh, Plat. Surf. Finish. 7 (1976) 34.
[3] D. Golodnitsky, Yu. Rosenberg, A. Ulus, Electrochim. Acta 47
(1) The Co content in Ni–Co alloys increased (2002) 2707.
gradually with the increase of Co2+ concentration [4] A. Bai, C.-C. Hu, Electrochim. Acta 47 (2002) 3447.
in the electrolyte, which is confirmed by the [5] L. Burzynska, E. Rudnik, Hydrometallurgy 54 (2000) 133.
anomalous codeposition of iron group metals. [6] D. Golodnitsky, N.V. Gudin, G.A. Volyanuk, Plat. Surf. Finish.
85 (1998) 65.
(2) Surface morphology of Ni–Co alloys changed [7] H.R. Johnson, J.W. Dini, Plat. Surf. Finish. 70 (1983) 47.
from regularly polyhedral crystallites into sphe- [8] E. Gómes, J. Ramirez, E. Vallés, J. Appl. Electrochem. 28
rical cluster surface when increasing Co content (1998) 71.
from 7 to 49 wt.% and the morphology of the [9] W.H. Safranek, Properties of Electrodeposited Metals and
Ni–Co alloys with 81 wt.% Co showed a rather Alloys, New York London, 1974.
[10] C.-C. Hu, Y.-S. Lee, T.-C. Wen, Mater. Chem. Phys. 48 (1997)
regularly branched structure. Both the crystal 246.
structure and phase composition are mainly [11] A.N. Correia, S.A.S. Machado, L.A. Avaca, Electrochem.
dependent on the Co content in alloys. The phase Commun. 1 (1996) 600.
structure of Ni–Co alloys gradually changed from [12] E. Gómes, E. Valles, J. Appl. Electrochem. 29 (1999) 805.
fcc into hcp with the increase of Co content. [13] S. Armyanov, Electrochim. Acta 45 (2000) 3323.
[14] E. Gómes, E. Vallés, J. Appl. Electrochem. 32 (2002)
(3) Microhardness of Ni–Co alloys increased initially 693.
with Co content increasing from 0 to approxi- [15] D. Golodnitsky, N.V. Gudin, G.A. Volyanuk, J. Electrochem.
mately 49 wt.%, and then gradually decreased as Soc. 147 (2000) 4156.
Co content increased further above 49 wt.%. The [16] N. Zech, E.J. Poklaha, D. Landolt, J. Electrochem. Soc. 146
hardness change of Ni–Co with grain size follows (1999) 2886.
[17] F.R. Morral, Met. Finish. 62 (1964) 82.
Hall–Petch effect. [18] G. Hibbard, K.T. Aust, G. Palumbo, et al. Scripta Mater. 44
(4) For the Ni-rich alloys, the improvement of wear (2001) 513.
resistance with hardness increase fits Archard’s [19] E. Arzt, Acta Mater. 46 (1998) 5611.
law. In addition, the Co-rich alloys exhibited [20] F. Dalla Torre, H. Van Swygenhoven, M. Victoria, Acta Mater.
much lower friction coefficient and higher wear 50 (2002) 3957.
[21] Kh. Saber, C.C. Koch, P.S. Fedkiw, Mater. Sci. Eng., A 341
resistance than Ni-rich alloys. It has been (2003) 174.
suggested that hcp crystal structure in Co-rich [22] J.F. Archard, J. Appl. Phys. 24 (1953) 981.
alloys contributed to the remarkable friction– [23] D.H. Jeong, F. Gonzalez, G. Palumbo, et al. Scripta Mater. 44
reduction effect and better anti-wear performance. (2001) 493.

Anda mungkin juga menyukai