You are on page 1of 8

Ceramics International 44 (2018) 766–773

Contents lists available at ScienceDirect

Ceramics International
journal homepage: www.elsevier.com/locate/ceramint

Synthesis of Ni–TiN composites through ultrasonic pulse electrodeposition T


with excellent corrosion and wear resistance

Fafeng Xiaa, Wanchun Jiaa, Chunyang Maa, , Jeremy Wangb
a
School of Mechanical Science and Engineering, Northeast Petroleum University, Daqing 163318, PR China
b
Department of Mechanical Engineering, University of Washington Bothell, Seattle, WA 98195, USA

A R T I C L E I N F O A B S T R A C T

Keywords: In this study, Ni–TiN composites with excellent corrosion and wear resistance were successfully prefabricated
Ni-TiN composite through the ultrasonic pulse electrodeposition (UPED) technique. The microstructure, cross-section composition
Ultrasonic pulse electrodeposition distribution, microhardness, corrosion and wear resistance of the composites were investigated through atomic
Synthesis force microscopy (AFM), X-ray photoelectron spectroscopy (XPS), scanning electron microscopy (SEM), Vickers
Microstructure
microhardness testing, electrochemical workstation utilization and wear testing. Both AFM and HRTEM results
Corrosion resistance
indicated that the Ni–TiN composite prepared with the UPED method displayed a fine and compact surface
Wear resistance
structure, whereas the average diameters of Ni grains and TiN nanoparticles in the composite were 68.4 nm and
26.8 nm, respectively. The XPS results demonstrated that the concentrations of Ti and Ni in the Ni–TiN com-
posite synthetized through the UPED method were approximately 21.7 at% and 47.2 at%, respectively. The
UPED-deposited Ni–TiN composite displayed the highest microhardness than the other coatings, whereas the
average microhardness of the nickel coating was only approximately 486.1 HV. The UPED-deposited Ni–TiN
composite displayed the best corrosion resistance, due to the introduction of moderate ultrasonication. The worn
surface morphologies of the UPED-deposited Ni–TiN composite were smooth, whereas only a few low-sized
scratches appeared on the surface, thereby displaying the best wear resistance in this wear test.

1. Introduction nanocomposite layers on steel substrates and observed that the TiN
nanoparticles could be produced on the surface of common metal
Pulse electrodeposition (PED) is a simple and effective technique for substrates through optimum conditions. Góral [23] demonstrated that
the synthesis of metal matrix composites with excellent physical and the Ni and Ni/Al2O3 composite coatings were obtained through elec-
mechanical properties [1–4]. The deposited composites are often used trodeposition technology. Dehgahi et al. [24] investigated the Ni–A-
to increase the lifetime of mechanical components, electronic compo- l2O3–SiC nanocoatings prepared through electrodeposition. It was dis-
nents or cutting tools exposed to corrosion or wear conditions [5–9]. covered that the micro-hardness, the wear resistance, the corrosion
Compared to the PED technology, the ultrasonic pulse electrodeposition performance and the passivation behaviors of the coatings were sig-
(UPED) technology, which utilizes the ultrasonic pulses as a means of nificantly improved through the incorporation of the SiC and Al2O3
dispersing bath, can achieve the co-deposition of the matrix metal and nanoparticles into the nickel matrix.
ceramic micro/nanoparticles for prefabricating certain composites with The aim of the current research was to synthesize Ni–TiN compo-
specific function [10,11]. During UPED, a method involving electronic sites with excellent corrosion and wear resistance through an UPED
agitation is utilized to inhibit agglomeration of micro/nanoparticles technique, as well as to compare the microstructures and performances
and to distribute metal ions uniformly throughout the plating bath. of PED-deposited Ni–TiN composites and nickel coatings. The micro-
In recent decades, ceramic nanoparticles (such as SiC, TiN and structure, the cross-section composition distribution, the microhard-
Al2O3) are generally utilized to be co-deposited on metals or metal alloy ness, the corrosion and the wear resistance of the composites were in-
composites to improve the corresponding properties [12–20]. Srivas- vestigated through atomic force microscopy (AFM), X-ray
tava et al. [21] reported the preparation of Ni and Ni–Co metal matrix photoelectron spectroscopy (XPS), scanning electron microscopy
composites with SiC nanoparticles through electroplating from an ad- (SEM), Vickers microhardness testing, electrochemical workstation
ditive free sulphamate bath. Wu et al. [22] electrodeposited Ni-TiN utilization and wear testing.


Corresponding author.
E-mail address: chunyangandma1@163.com (C. Ma).

http://dx.doi.org/10.1016/j.ceramint.2017.09.245
Received 13 September 2017; Received in revised form 25 September 2017; Accepted 30 September 2017
Available online 03 October 2017
0272-8842/ © 2017 Elsevier Ltd and Techna Group S.r.l. All rights reserved.
F. Xia et al. Ceramics International 44 (2018) 766–773

Table 1 generator and the heating solution unit. The pulse current was gener-
Electrolyte composition and plating conditions for preparing Ni-TiN composites. ated by a pulse power supply (SMD-30, from Handan Dashun Electro-
plating Source Co., Ltd, Handan Hebei, PR China) and the ultrasonic
Sample Ni coating Ni–TiN Ni–TiN
composite composite vibration was produced by an ultrasonic generator (TUS-50, Beijng
(denoted as (denoted as (denoted as Tianyi Ultrasonic Equipment Co., Ltd, Beijng, PR China). The heating
PED-1) PED-2) UPED-1) solution unit was applied to retain the plating bath temperature at
45 °C. During electroplating, all samples were retained at the identical
NiSO4·6H2O (g/L) 200 200 200
NiCl2·H2O (g/L) 30 30 30 thickness of ~ 80 µm.
H3BO3 (g/L) 30 30 30
TiN nanoparticles (g/L) 8 8 8 2.2. Characterization of Ni–TiN composites
CTAB (mg/L) 60 60 60
Plating temperature (°C) 45 45 45
The cross-sectional distributions and structural morphologies of the
Current density (A/dm2) 6 6 6
Duty cycle (%) 30 30 30 Ni–TiN composites were observed through X-ray photoelectron spec-
pH 4.5 4.5 4.5 troscopy (XPS, Vg-Escalab-200iXL), atomic force microscopy (AFM,
Ultrasonic power (W) 0 0 200 NanoScope Ⅲ), high-resolution transmission electron microscopy
(HRTEM, Tecnai-G2-20-S-Twin), and energy disperse spectroscopy
(EDS, Inca X-Max), respectively. The XPS analysis was conducted
2. Experimental
through an Al excitation source operated at 1.5 keV and X-ray spot size
of 200 µm. The microhardness values of the composites were examined
2.1. Preparation of Ni–TiN composites
through a VTD512 model microhardness tester (Beijing Guangce mi-
crohardness tester Co., Ltd, PR China) with the applied load of 0.98 N
In this investigation, Ni–TiN composites were prepared through
for 10 s. Prior to each investigation, the samples were polished with
UPED techniques in a 1 L glass beaker. The plating electrolyte was
800, 1000 and 1400 grit metallographic sandpapers. Consequently, the
produced through a Watts-type nickel bath, whereas the composition
microhardness values of the samples were obtained along the profile of
and plating conditions are provided in Table 1. All chemicals in this
the coatings at an interval of 20 µm.
experiment were of analytical grade and utilized without further pur-
ification. The pH value of the plating electrolyte was 4.5, through the
2.3. Electrochemical analysis
hydrochloric acid solution (10 vol%) or the sodium hydroxide solution
(10 vol%). The total deposition time was set at 50 min. The TiN na-
Electrochemical tests of the Ni–TiN composites and nickel coatings
noparticles utilized in this research were purchased from Shanghai
were carried out in a 0.5 M NaCl neutral solution for 1 h at room
Jingcai Nanometer Technology Co., Ltd. The average diameter was
temperature of 25 °C through an electrochemical apparatus (CS350,
20 nm, the surface-to-volume ratio was 100 m2/g and the density was
Hubei Wuhan Corrtest Instrument Co., Ltd, China). A conventional
0.05 g/cm3.
three electrode cell was applied in the electrochemical tests with the
The working anode was a pure nickel plate of 30 cm2 in area and the
deposited samples as working electrodes, a saturated calomel electrode
working cathode was a mild steel sheet of 20 mm × 60 mm in di-
(SCE) was utilized as the reference electrode, whereas a platinum plate
mensions. The nickel plate was composed of 99.65% Ni, 0.16% Cu,
was utilized as the auxiliary electrode. The Tafel curves were recorded
0.12% Mn, 0.05% Ti and 0.02% Al, whereas the mild steel sheet was
through the potential sweep from −1.0 V to +0.5 V at a scan frequency
composed of 99.12% Fe, 0.31% Si, 0.29% Mn, 0.2% C, 0.05% S and
of 2 Hz and a scan rate of 0.5 mV/s. Also, the corrosion current (Icorr) or
0.03% P. The distance between the anode and the cathode was retained
the corrosion potential (Ecorr) of the Ni-TiN composites were obtained
at 30 mm. Prior to deposition, the mild steel sheets were firstly polished
from the curves. Following the electrochemical corrosion tests, the
to a ~ 0.12 µm surface finish in proper order through 400, 800, 1000
corrosion surfaces of the composites were investigated with a scanning
and 1400 grit metallographic sandpapers. Consequently, the sheets
electron microscope (SEM, JSM-6460LV).
were ultrasonically washed (300 W) in alcohol for 15 min, followed by
In order obtain an improved understanding of the Ni–TiN compo-
cleaning with distilled water at room temperature. Subsequently, the
sites corrosion ability, the electrochemical impedance spectroscopy
substrate was activated in a 0.5 M hydrochloric acid solution for 15 s, as
(EIS) measurements were conducted in the frequency range of 100 kHz
well as rinsed with acetone and distilled water, and immediately im-
to 100 mHz and the sinusoidal signal amplitude was 10 mV.
mersed in the plating solution. The electrodeposition setup for the
Ni–TiN composites preparation is presented in Fig. 1. The experimental
2.4. Friction measurement
installation mainly comprised the pulse power supply, the ultrasonic
The wear behaviors of the Ni–TiN composites and Ni coatings
sliding against a chromium ball of 3 mm in diameter were obtained
from a 10 N applied load on a wear testing instrument (MR-H5A,
Beijing Shidai wear tester Co., Ltd, PR China) under the following
conditions: dry sliding condition, rotation speed of 100 rpm, tempera-
ture of 25 °C and testing duration of 140 min. Subsequently to sliding,
the wear surfaces of the coatings were investigated with the JSM-
6460LV type scanning electron microscope.

3. Results and discussion

3.1. AFM observation

Fig. 2 illustrates the AFM surface morphologies of the Ni coating


and Ni–TiN composites. The Ni coating in Fig. 2a and the Ni–TiN
composite in Fig. 2b were deposited through the PED technique. The
Fig. 1. Experimental diagram for preparing Ni–TiN composites.
two coatings indicated coarse surface morphology in the micro-regions.

767
F. Xia et al. Ceramics International 44 (2018) 766–773

3.2. HRTEM detection

Fig. 3a–c display the HRTEM microstructures of the Ni coating and


Ni–TiN composites produced through PED and UPED deposition, re-
spectively. The white and black sections were the nickel grains and the
TiN nanoparticles, respectively. The nickel grains in the UPED-de-
posited Ni–TiN composite were observed to be lower-sized compared to
the Ni coating and the Ni–TiN composite deposited through PED. It was
indicated that a moderate ultrasonication led to an increase in the TiN
nanoparticle content in the Ni-TiN composite, resulting in the TiN na-
noparticles inhibition of the nickel crystals further growth.
Fig. 3a′–c′ exhibit the statistical distribution of nickel grains within
the Ni coating and the Ni–TiN composites prepared through PED and
UPED deposition, respectively. It was evident that the average diameter
of the Ni grains in the Ni coating was approximately 148.4 nm. More-
over, the average diameters of the Ni grains and the TiN nanoparticles
in the PED-deposited Ni–TiN composite were approximately 119.3 nm
and 41.1 nm, respectively. In contrast, the average diameters of the Ni
grains and the TiN nanoparticles in the UPED-deposited Ni–TiN com-
posite were 68.4 nm and 26.8 nm, respectively.

3.3. Plating mechanism analysis

Fig. 4 presents the plating process for the Ni–TiN composites pre-
paration. According to the Gugliemi's absorption model, the process
between nickel irons and TiN nanoparticles could be divided into two
steps [25]: Firstly, the TiN nanoparticles were weakly adsorbed on the
surface of the Ni–TiN composites under the electrical field forces gen-
erated by the pulse power supply. Secondly, the TiN nanoparticles were
captured by nickel irons and consequently embedded in the coating.
Fig. 5 reveals the formation and refinement mechanism of Ni–TiN
composites with different preparation methods. During PED, the addi-
tion of TiN nanoparticles limited the growth of nickel grains, resulting
in the crystallite diameters in the Ni–TiN composite becoming lower-
sized compared to the Ni coating. The explanation for this phenomenon
was that the TiN nanoparticles in the Ni–TiN composite that were
dispersed in the composite increased the number of nuclei for the nu-
cleation of nickel metal crystals as well as the grain growth inhibition
[26]. In contrast, the UPED-deposited Ni–TiN composite possessed an
optimum microstructure compared to the coatings produced through
the PED technique. The reason for this phenomenon was that the TiN
nanoparticles in the Ni–TiN composite could effectively inhibit the
growth of nickel crystals. Furthermore, the shock waves produced by
the ultrasonic generator might disrupt the higher-sized nickel grains to
produce lower-sized nuclei. In addition, the moderate ultrasonication
led to homogeneous dispersion of the TiN nanoparticles within the
composite. Consequently, these caused the formation of a fine and
compact structure on the surface of the UPED-deposited Ni–TiN com-
posites.

3.4. XPS survey

Fig. 6 shows the XPS cross-sectional profiles of the Ni coating and


Ni–TiN composites. It was indicated that Ti did not exist in the nickel
coating. In contrast, it could be observed that the TiN nanoparticles
Fig. 2. AFM images of the Ni coating and Ni–TiN composites: (a) PED-1, (b) PED-2, (c) were completely embedded in the Ni-TiN composites. The concentra-
UPED-1.
tions of Ti and Ni in the Ni–TiN composite synthetized through the PED
method were approximately 14.8 at% and 60.5 at% respectively
The crystallite diameter in the Ni coating (Fig. 2a) was higher than the (Fig. 4b). By contrast, in the composite deposited by the UPED method,
crystallite diameter in the Ni–TiN composite (Fig. 2b). In contrast, the high concentrations of Ti (21.7 at%) and Ni (47.2 at%) were detected
Ni–TiN composite prepared with the UPED method displayed a fine and throughout the corresponding thickness (Fig. 4c), due to moderate ul-
compact surface structure (Fig. 2c). This result suggested that the trasonication leading to the electrochemical reaction acceleration and
UPED-deposited Ni–TiN composite displayed an optimum micro- the co-deposition promotion of both nickel crystals and TiN nano-
structure compared to the coatings produced under other conditions. particles.

768
F. Xia et al. Ceramics International 44 (2018) 766–773

Fig. 3. HRTEM dark field images and nickel grain sizes of the Ni coating and Ni–TiN composites: (a) and (a′) PED-1, (b) and (b′) PED-2, (c) and (c′) UPED-1.

3.5. Microhardness test Ni–TiN composite, which produced a dispersion–hardening effect.


Furthermore, the TiN nanoparticles already possessed high micro-
Fig. 7 indicates the microhardness profiles of the Ni coating and the hardness, thereby enhancing the microhardness of the Ni–TiN compo-
Ni–TiN composites. The profiles could be divided into two zones: the sites [27].
composite and substrate zones. It could be clearly observed that the
UPED-deposited Ni–TiN composite possessed the highest microhardness
(average value of ~ 714.6 HV) than the other coatings, whereas the 3.6. Electrochemical analysis
average microhardness of the nickel coating was only 486.1 HV ap-
proximately. Based on the aforementioned AFM and HRTEM results, a The polarization curves of the Ni coating and Ni–TiN composites in
high number of TiN nanoparticles existed in the UPED-deposited a 0.5 M NaCl neutral solution are illustrated in Fig. 8, whereas in
Table 2 the electrochemical corrosion data collected from the Tafel

769
F. Xia et al. Ceramics International 44 (2018) 766–773

Fig. 4. Sketch map of the codeposition process between nickel irons and TiN nano-
particles.

Fig. 6. XPS cross-sectional profiles of the Ni coating and Ni–TiN composites: (a) PED-1,
(b) PED-2, (c) UPED-1.

Fig. 5. The form and refinement mechanism of Ni–TiN composite with different pre-
paration methods: (a) PED-1, (b) PED-2, (c) UPED-1.

curves of the Ni coating and the Ni–TiN composites are summarized.


Among all three coatings, the UPED-deposited Ni–TiN composite had
the minimum Icorr and Ecorr values of 1.12 × 10−3 mA/cm2 and
−0.312 V, respectively, thereby illustrating the best corrosion re-
sistance. In contrast, the Ni coating had the maximum Icorr and Ecorr
values of 23.65 × 10−3 mA/cm2 and −0.760 V, respectively, which Fig. 7. Microhardness curves of the Ni coating and Ni–TiN composites: (a) PED-1, (b)
presented the worst corrosion resistance. Fig. 9 presents the SEM cor- PED-2, (c) UPED-1.

rosion morphologies of the Ni coating and the Ni–TiN composites


subsequently to the corrosion tests. The nickel coating sustained the

770
F. Xia et al. Ceramics International 44 (2018) 766–773

Fig. 8. Tafel curves of the Ni coating and Ni–TiN composites in a 0.5 M NaCl neutral
solution: (a) PED-1, (b) PED-2, (c) UPED-1.

Table 2
Electrochemical corrosion data of the Ni coating and Ni–TiN composites.

Sample Ba (mV) Bc (mV) Icorr (mA cm−2) Ecorr (V)

PED-1 80.45 271.03 23.65 × 10−3 −0.760


PED-2 113.04 213.58 9.14 × 10−3 −0.587
UPED-1 165.32 186.41 1.12 × 10−3 −0.312

most severe corrosion damage, whereas the UPED-deposited Ni–TiN


composite displayed the best corrosion resistance in this corrosion test.
Even following the immersion in the 0.5 M NaCl neutral solution for
1 h, this composite only exhibited certain low-sized pits on the surface.
This finding supported that the moderate ultrasonication led to the
formation of lower-sized nickel grains in the Ni–TiN composite, re-
sulting in a fine and compact morphology, which could effectively re-
strict the NaCl etchant solution contact with the composite. In contrast,
the Ni coating with the roughest surface morphology could be propi-
tious to the reaction between the NaCl etching solution and the nickel
grains, resulting in a poor corrosion resistance of the coating [28].
Fig. 10 demonstrates the EIS analysis of the Ni coating and the
Ni–TiN composites. Also, the equivalent circuit parameters fitted to the
impedance curves are presented in Table 3. The Rs and Rct represented
the solution resistance and the charge transfer resistance, respectively.
The CPEdl was the electric double layer capacity [29]. The results
clearly demonstrated that the PED-deposited Ni coating presented the
minimum value of impedance (Z), which illustrated that the Ni coating
displayed inadequate corrosion resistance. In contrast, the UPED-de-
posited Ni–TiN composite displayed a drastic increase in corrosion re-
sistance, implying that it possessed the optimal anticorrosion ability. It
could also be observed from Table 3 that with the introduction of ul- Fig. 9. SEM micrographs of the Ni coating and Ni–TiN composites after the corrosion test:
trasonication, the Rct increased and the CPEdl decreased, demonstrating (a) PED-1, (b) PED-2, (c) UPED-1.
the corrosion resistance increase of the UPED-deposited Ni–TiN com-
posite. These results were consistent with the polarization curve find-
abrasion weight loss of the coatings depends on both the microstructure
ings.
of the surface and the amount of the incorporated ceramic particles
[30]. On the one hand, the UPED-deposited Ni-TiN composites with
3.7. Friction measurement
lower grain sizes were significantly denser than the PED-deposited Ni-
TiN composites and the Ni coating, which could be observed from
The abrasion weight losses as a function of the wear testing duration
Fig. 2. This may contribute to the improvement of wear resistance. On
for the Ni coating and Ni–TiN composites are demonstrated in Fig. 11.
the other hand, the TiN amount in the UPED-deposited Ni-TiN com-
During the measurements time period, the abrasion curve trend for the
posite was the highest compared to the PED-deposited Ni-TiN compo-
PED- and UPED-deposited Ni–TiN composites was the same: initially
site and the Ni coating, which could be observed in Fig. 6. This could
the curves increased rapidly, followed by a slight change. Also, the improve the wear resistance of the UPED-deposited composite [31].
abrasion weight losses of the Ni coating exhibited a continual and rapid
The surface morphologies of the Ni coating and the Ni–TiN com-
increase as the wear duration increased. Moreover, it was observed that posites subsequently to wear testing for 140 min are provided in
the Ni–TiN composite fabricated by the UPED method sustained the
Fig. 12. Certain high-sized wear debris, certain clear scratches and
least weight loss, whereas the Ni coating displayed the highest weight grooves on the worn surface of the Ni coating existed. Also, the worn
loss under the same wear testing duration. As it is well known, the

771
F. Xia et al. Ceramics International 44 (2018) 766–773

Fig. 10. Equivalent electrical circuit model for the corrosion behavior of the Ni coating
and Ni–TiN composites in a 0.5 M NaCl neutral solution: (a) PED-1, (b) PED-2, (c) UPED-
1.

Table 3
The equivalent circuit parameters of the Ni coating and Ni–TiN composites.

Sample RS (Ω cm2) Rct (Ω cm2) Qdl (F) n

−5
PED-1 10.1 5021 5.11 × 10 0.8
PED-2 12.5 9754 3.76 × 10−5 0.8
UPED-1 14.2 19,192 3.19 × 10−5 0.8

Fig. 11. Relationship between the weight losses of the Ni coating and Ni–TiN composites
and the wear time: (a) PED-1, (b) PED-2, (c) UPED-1.

surface of the Ni coating was typical of material spall, scratching, as


well as plastic deformation. The worn surface of the PED-deposited Fig. 12. SEM micrographs of the worn surface of the Ni coating and Ni–TiN composites:
Ni–TiN composite exhibited that the abrasive scratches and grooves (a) PED-1, (b) PED-2, (c) UPED-1.
were significantly decreased. In contrast, the worn surface morpholo-
gies of the UPED-deposited Ni–TiN composite were smooth, whereas 4. Conclusion
only a few low-sized scratches appeared on the surface, thereby dis-
playing the best wear resistance in this wear test. It was inferred that Ni–TiN composites with excellent corrosion and wear resistance
the co-deposited TiN led to the wear resistance increase of the com- were successfully prefabricated through the ultrasonic pulse electro-
posites, which was attributed to the dispersion strengthening effect deposition technique. Also, the microstructure, the cross-section com-
through the TiN nanoparticles addition [32]. During UPED plating, the position distribution, the microhardness, the corrosion and the wear
TiN particles with high hardness were dispersed homogeneously within resistance of the composites were investigated in detail. Both AFM and
the Ni–TiN composites, as presented in Fig. 3. The composites became HRTEM results indicated that the Ni–TiN composite prepared with the
significantly robust and stable, consequently demonstrating a high wear UPED method displayed a fine and compact surface structure, whereas
resistance. Besides, during wear testing, the co-deposited TiN could the average diameters of Ni grains and TiN nanoparticles in the com-
slightly bulge out of the nickel matrix, which retained the applied loads posite were 68.4 nm and 26.8 nm, respectively. The XPS results de-
transmitted from the matrix and caused a shear force decrease between monstrated that the concentrations of Ti and Ni in the Ni–TiN compo-
the contact interfaces [33]. Consequently, the wear resistance was ap- site synthetized through the PED method were approximately 14.8 at%
parently improved through the well distributed TiN nanoparticles ad- and 60.5 at% respectively. By contrast, in the composite deposited by
dition in Ni-TiN composites.

772
F. Xia et al. Ceramics International 44 (2018) 766–773

the UPED method, high concentrations of Ti (21.7 at%) and Ni (47.2 at Membr. Sci. 530 (2017) 65–72.
[13] M.K. Tripathi, D.K. Singh, V.B. Singh, Microstructure and properties of electro-
%) were detected throughout the corresponding thickness. The UPED- chemically deposited Ni-Fe/Si3N4 nanocomposites from a DMF Bath, J.
deposited Ni–TiN composite displayed the highest microhardness Electrochem. Soc. 162 (2015) D87–D95.
(average value of ~ 714.6 HV) than the other coatings, whereas the [14] C.H. Park, J.H. Lee, J.P. Jung, J.H. Kim, Mixed matrix membranes based on dual-
functional MgO nanosheets for olefin/paraffin separation, J. Membr. Sci. 533
average microhardness of the nickel coating was only approximately (2017) 48–56.
486.1 HV. The nickel coating sustained the most severe corrosion da- [15] R.M. Reddy, B.M. Praveen, C.M.P. Kumar, T.V. Venkatesha, Pulse electrodeposition,
mage, whereas the UPED-deposited Ni–TiN composite displayed the characterization, and corrosion behavior of Ni-Si3N4 composites, J. Mater. Eng.
Perform. 24 (2015) 1987–1994.
best corrosion resistance in this corrosion test. The worn surface [16] E. Bełtowska-Lehman, A. Goral, P. Indyka, Electrodeposition and characterization
morphologies of the UPED-deposited Ni–TiN composite were smooth, of Ni/Al2O3 nanocomposite coatings, Arch. Metall. Mater. 56 (2011) 919–931.
whereas only a few low-sized scratches appeared on the surface, [17] S. Sangeetha, G.P. Kalaignan, Tribological and electrochemical corrosion behavior
of Ni-W/BN (hexagonal) nano-composite coatings, Ceram. Int. 41 (9) (2015)
thereby displaying the best wear resistance in this wear test.
10415–10424.
[18] F. Xia, W. Jia, M. Jiang, W. Cui, J. Wang, Microstructure and corrosion properties of
Acknowledgement Ni-TiN nanocoatings prepared by jet pulse electrodeposition, Ceram. Int. (2017),
http://dx.doi.org/10.1016/j.ceramint.2017.07.117.
[19] T. Borkar, S.P. Harimkar, Effect of electrodeposition conditions and reinforcement
This work has been supported by the National Natural Science content on microstructure and tribological properties of nickel composite coatings,
Foundation of China (Granted no. 51474072). Surf. Coat. Technol. 205 (2011) 4124–4134.
[20] C. Ma, M. Jiang, F. Xia, Preparation and characterization of Ni–TiN thin films
electrodeposited with nickel baths of different TiN nanoparticle concentration, Surf.
References Rev. Lett. 24 (3) (2016) 1–7.
[21] M. Srivastava, V.K. William Grips, K.S. Rajam, Electrochemical deposition and tri-
[1] C. Ma, X. Guo, J. Leang, F. Xia, Synthesis and characterization of Ni–P-TiN nano- bological behaviour of Ni and Ni–Co metal matrix composites with SiC nano-
composites fabricated by magnetic electrodeposition technology, Ceram. Int. 42 (8) particles, Appl. Surf. Sci. 253 (2007) 3814–3824.
(2016) 10428–10432. [22] M. Wu, J. Wei, P. Lv, Electrodepositing Ni-TiN nanocomposite layers with applying
[2] F. Xia, J. Tian, W. Wang, Y. He, Effect of plating parameters on the properties of action of ultrasonic waves, Procedia Eng. 174 (2017) 717–723.
pulse electrodeposited Ni–TiN thin films, Ceram. Int. 42 (11) (2016) 13268–13272. [23] A. Góral, Nanoscale structural defects in electrodeposited Ni/Al2O3 composite
[3] X. Li, Y. Zhu, G. Xiao, Application of artificial neural networks to predict sliding coatings, Surf. Coat. Technol. 319 (2017) 23–32.
wear resistance of Ni–TiN nanocomposite coatings deposited by pulse electro- [24] S. Dehgahi, R. Amini, M. Alizadeh, Corrosion, passivation and wear behaviors of
deposition, Ceram. Int. 40 (8) (2014) 11767–11772. electrodeposited Ni–Al2O3–SiC nano-composite coatings, Surf. Coat. Technol. 304
[4] H.F. Zhou, N. Du, L.W. Zhu, J.K. Shang, Z.H. Qian, X.M. Shen, Characteristics in- (2016) 502–511.
vestigation of Ni-diamond composite electrodeposition, Electrochim. Acta 151 [25] T.S. Han, A.B. Yu, X.L. Tian, Effect of surfactants on Ni-Si3N4 composite coating,
(2015) 157–167. Chin. Surf. Eng. 6 (2012) 32–35.
[5] H. Zhao, C. Jiang, X. He, J. Ren, C. Wan, Preparation of micro-porous membrane [26] M.H. Wu, Z. Li, F.F. Xia, L.Y. Chen, Preparating technique of nanocermet composite
electrodes and their application in preparing anodes of rechargeable lithium bat- layer with method of ultrasonic-electrodepositing, Mater. Mech. Eng. 29 (8) (2005)
teries, J. Membr. Sci. 310 (1–2) (2008) 1–6. 58–61.
[6] M. Eslami, H. Saghafian, F. Golestani-fard, A. Robin, Effect of electrodeposition [27] F. Xia, J. Tian, C. Ma, M. Potts, X. Guo, Effect of pulse frequency on microstructural,
conditions on the properties of Cu–Si3N4 composite coatings, Appl. Surf. Sci. 300 nanomechanical and wear properties of electrodeposited Ni–TiN composite coat-
(2014) 129–140. ings, J. Appl. Phys. 116 (23) (2014) 234301–234306.
[7] J. Man, S. Zhang, J. Li, B. Zhao, Y. Chen, Effects of electrolyte pH on morphologies [28] S.L. Mu, N. Li, D.Y. Li, L.Y. Xu, Corrosion behavior and composition analysis of
and mechanical properties of α-Al2O3/Ni composite coatings and role of zeta po- chromate passive film on electroless Ni-P coating, Appl. Surf. Sci. 256 (13) (2010)
tentials in co-deposition process, Surf. Coat. Technol. 249 (2014) 118–124. 4089–4094.
[8] S. Sangeetha, G. Paruthimal Kalaignan, Tribological and electrochemical corrosion [29] Y.S. Huang, X.T. Zeng, X.F. Hu, Corrosion resistance properties of electroless nickel
behavior of Ni–W/BN (hexagonal) nano-composite coatings, Ceram. Int. 41 (9) composite coatings, Electrochim. Acta 49 (2004) 4313–4319.
(2015) 10415–10424. [30] S.R. Allahkaram, S. Golroh, M. Mohammadalipour, Properties of Al2O3 nanoparticle
[9] K.A. Kumar, G.P. Kalaignan, V.S. Muralidharan, Pulse electrodeposition and char- reinforced copper matrix composite coatings prepared by pulse and direct current
acterization of nano Ni–W alloy deposits, Appl. Surf. Sci. 259 (2012) 231–237. electroplating, Mater. Des. 32 (2011) 4478–4484.
[10] G.Q. Liang, L.J. Zou, Y.Y. Zhu, F.F. Xia, Preparation and corrosion behavior of [31] F. Xia, J. Tian, C. Ma, X.Y. Xu, M. Huang, Microstructural, nanomechanical and
Ni–TiN coatings deposited by ultrasonic electrodeposition, J. Funct. Mater. 45 (13) wear properties of magnetic pulse electrodeposited Ni-TiN composite coatings, Sci.
(2014) 13059–13066. Eng. Compos. Mater. 23 (5) (2016) 535–541.
[11] F.F. Xia, C. Liu, F. Wang, M.H. Wu, J.D. Wang, H.L. Fu, J.X. Wang, Preparation and [32] X.S. Peng, X.M. Zhang, C.Y. Ma, Predicative study on the wear resistances of Ni-TiN
characterization of nano Ni–TiN coatings deposited by ultrasonic electrodeposition, coatings using BP neural networks, J. Synth. Cryst. 45 (6) (2016) 1718–1721.
J. Alloy. Compd. 490 (2010) 431–435. [33] Y. Suzuki, S. Arai, M. Endo, Electrodeposition of Ni-P alloy-multiwalled carbon
[12] T. Makinouchi, M. Tanaka, H. Kawakami, Improvement in characteristics of a nanotube composite films, J. Electrochem. Soc. 157 (2010) D50–D53.
Nafion membrane by proton conductive nanofibers for fuel cell applications, J.

773

You might also like