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

Ceramics International xxx (xxxx) xxx

Contents lists available at ScienceDirect

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

SiC nanoparticles incorporation in electroless NiP-Graphene oxide


nanocomposite coatings
M. Khodaei *, A. Mohammad Gholizadeh
Faculty of Materials Science and Engineering, K. N. Toosi University of Technology, Tehran, Iran

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

Keywords: The presence of SiC nanoparticles within the Graphene oxide (GO) incorporated electroless deposited NiP layers
Electroless Ni plating (NiP-GO) on carbon steel substrate was studied in this work. The effect of co-deposition of GO nanoplatelets and/
Nanocomposite coatings or SiC nanoparticles on the morphology and structure of the heat-treated NiP coatings were investigated by
Graphene oxide nanoplatelets
scanning electron microscope and X-ray diffraction, respectively. The results revealed that the heat-treated NiP
SiC nanoparticles
Wear
and NiP–SiC coatings consisted of Ni and Ni3P phases, whereas the NiP-GO also contains the intermediated Ni2P
Corrosion and Ni12P5 metastable phases due to the incomplete precipitation of Ni3P. Such metastable phases are signifi­
cantly decreased by the incorporation of SiC nanoparticles in NiP-GO coatings. The mechanical properties of the
coatings were characterized by microhardness measurement and “pin on disk” wear test. The corrosion tests were
conducted in aqueous 3.5 %wt NaCl using electrochemical measurement for Ni–P, NiP-GO, NiP–SiC, and NiP-GO-
xSiC coatings. By co-deposition of SiC nanoparticles, the hardness of NiP-GO is significantly increased and the
wear loss is reduced, especially at a high sliding distance during the wear test. The corrosion behavior of the NiP-
GO coatings containing different amounts of SiC nanoparticles has been investigated.

1. Introduction Ni coatings have been fabricated by incorporating micron-size particles.


The possibility of the incorporation of nanomaterials (0D, 1D, and 2D)
Plated Ni based coatings prepared by both electrolytic and electro­ has introduced a new generation of composite coatings (nanocomposite
less deposition have been utilized for corrosion and oxidation as well as coatings), which demonstrate enhanced mechanical, physical and
wear protection of plane steels used in different industries such as oil functional properties as a result of nanoscale effects [3]. Selection of
and gas industries. Among them, the autocatalytic Ni electroless second-phase has been performed according to the expected improve­
method, in which the Ni atoms deposit on the surface by autocatalytic ment in the performance of the Ni layer such as providing the higher
reduction of Ni2+ ions, possesses the advantages to the electrolytic Ni corrosion resistance and/or improved tribological behavior (high
deposition [1]. These advantages are not only the non-consuming of the hardness, low wear loss, and lower/higher coefficient of friction).
electrical current but also the ability to the formation of the uniform Hence, various type of second-phase (soft/hard, micron-size/nano-size,
coatings on the parts with complex geometric (deep hole and sharp and particle/fiber) such as Al2O3 [4], TiO2 [5,6], SiO2 [7], Si3N4 [8],
edges). The non-uniformity of the thickness of the electrolytically ZrO2 [9,10], TiC [11–13], diamond [14], SiC [15,16], TiN–SiC [17]
deposited layer on the complex geometric is due to the variation of (hard second-phase) and polyfluorotetraethylene (PTFE) [18], MoS2
current density along the surface [1,2]. Hence, the electroless deposition [19], WS2 [20], carbon based such as graphite [21], carbon black [22],
is interesting especially in industrial point of view. In the last decades, carbon nano-tubes (CNTs) [23], Graphene nanoplatelets and its de­
the enhancement in performance of plated Ni coatings has been pursued rivatives [24–27] (soft second-phase) have been selected especially for
by the development of advanced coatings such as composite Ni coatings electroless Ni coatings to deposit composite/nanocomposite coatings.
by both electrolytic and electroless methods. Co-deposition of Generally, it has been revealed that the presence of the second-phase in
second-phase particles along with Ni leads to the formation of Ni com­ the Ni layer can be effective in increasing the corrosion resistance [26,
posite coatings which have been reported to have superior properties in 28–30]. The hard materials, which have higher hardness than Ni, can act
comparison to the pure Ni coatings [1,3]. Earlier works in the composite as a load-bearing part leading to the improvement of mechanical

* Corresponding author.
E-mail address: khodaei@kntu.ac.ir (M. Khodaei).

https://doi.org/10.1016/j.ceramint.2021.05.250
Received 2 April 2021; Received in revised form 3 May 2021; Accepted 24 May 2021
Available online 27 May 2021
0272-8842/© 2021 Published by Elsevier Ltd.

Please cite this article as: M. Khodaei, A. Mohammad Gholizadeh, Ceramics International, https://doi.org/10.1016/j.ceramint.2021.05.250
M. Khodaei and A.M. Gholizadeh Ceramics International xxx (xxxx) xxx

properties (increase in the surface hardness as well as a decrease in the


wear loss). On the other hand, soft materials, which are typically solid
lubricant materials, lead to the enhancement of tribological perfor­
mance (decrease in the coefficient of friction and wear loss). Generally,
the hardness and load-bearing capacity of Ni–P coatings decrease by the
presence of soft second-phase. Hence, the co-deposition of hard and soft
second-phases (creating the hybrid composite/nanocomposite coatings)
has been investigated such as NiP–WS2–AlN [31], NiP–MoS2–SiO2 [32],
NiP–MoS2–Al2O3 [33], NiP-PTFE-Al2O3 [34] to achieve the self-lubricity
along with load-bearing properties within the Ni–P coatings,
simultaneously.
Graphene family nanoplatelets, 2D nanomaterials with layered
structure, are emerging solid lubricant materials that can be incorpo­
rated in the metallic matrix to create self-lubricant coatings. In addition
to their solid lubricant property, other unique properties such as high
surface area, chemical and thermal stability resulted in the numerous
attentions for using as a reinforcement in composite coatings. To fabri­
cate the composite coatings by plating processes, the graphene oxide
(GO) nanoplatelets or reduced graphene oxide (rGO) nanoplatelets,
which are oxygenated derivatives of graphene, are preferred for using in
an aqueous solution because of their capability of easily dispersing in
polar solvents. Generally, the corrosion resistance of Ni coatings can be
increased by the incorporation of carbon-based nanomaterials, espe­
cially by Graphene family nanoplatelets [29]. The corrosion resistance
of NiP coating by incorporation of GO (NiP-GO) is higher than that of
NiP-rGO [26]. In addition, the tribological behavior; i.e., reduction in
coefficient of friction as well as wear loss of NiP coatings, can be
improved by incorporation of GO or rGO, which can be resulted from
their solid lubricant property [25,35]. Although such improvement has
been reported for NiP coatings, the reduction in the hardness of the NiP
coatings has been reported by incorporating the Graphene family
nanoplatelets, which can be due to the inappropriate load-bearing
capability that is needed for heavy-duty wear conditions and
tribo-systems. Hence, increasing the hardness of NiP-Graphene nano­
composite coatings is interesting in the technological point of view,
which has been investigated by co-deposition of Graphene nanoplatelets
and ceramic nanoparticles such as few reports that have been recently
published such as NiP-rGO-TiO2 [36], NiP-GO-TiO2 [37] Ni-GO-Al2O3
[38]. In this work, the effect of co-deposition of SiC nanoparticles and
GO nanoplatelets within the Ni layer is investigated on the tribological
and corrosion behavior of heat-treated nickel phosphorous electroless
coatings.

2. Experimental section

2.1. Materials

A commercial nickel electroless solution (SLOTONIP 70A (9–11 %P),


Schloter, Germany) containing nickel sulfate as nickel source, sodium
hypophosphite as reducing agent, and suitable amounts of additive and
stabilizer was used for electroless Ni–P plating. A commercially avail­
able plane steel (St37) rod was used to prepare the substrates. Graphene
oxide (GO) nanoplatelets with a thickness of 3.4–7 nm and lateral
dimension of 10–50 μm (US Research Nanomaterials Inc.) and β-SiC
nanoparticles (US Research Nanomaterials Inc.) with the average size of
45–60 nm were used as soft and hard second-phase, respectively. Elec­ Fig. 1. SEM micrograph of surface of the (a) NiP, (b) NiP-GO, (c) NiP–2SiC, (d)
tron microscopy images of GO nanoplatelets and SiC nanoparticles are NiP-GO-1SiC, (e) NiP-GO-2SiC, and (f) NiP-GO-3SiC samples at two different
presented in Fig. 1S of supplementary data. Reagent grade NaOH, magnifications.
H2SO4, HCl, HNO3, ammonia 25% solution, acetone (Merck; Germany)
were used as received. ultrasonic bath. Then the substrates were subjected to pickling treat­
ment before coating process as following steps; 20 min in 20 wt% NaOH
2.2. Preparation of samples aqueous solution at 70 ◦ C, 3 min in 10 vol% H2SO4 solution, and 1 min in
15 vol% HCl solution. After each step, the substrates were washed with
The disc-shaped substrates (5 cm diameter and 0.7 cm thickness) distilled water. After pickling treatment, the substrates were transferred
were cut from the St37 rod. The substrates were polished with sandpa­ to the plating solution (400 ml) which was in a 600 ml glass beaker on
pers of #400 to #1200 and sequentially cleaned by acetone using an the magnetic stirrer hotplate. The temperature of the plating solution

2
M. Khodaei and A.M. Gholizadeh Ceramics International xxx (xxxx) xxx

was kept at 90 ± 5 ◦ C and it was stirred at 150 rpm to generate a constant preciously using a balance with an accuracy of ± 5 μg after every 100 m
agitation for easy scape of hydration as well as suspending of GO sliding distance. Corrosion behavior of the samples was studied by po­
nanoplatelets and/or SiC nanoparticles in the composite plating solu­ larization tests using EG&G Model 273A according to the ASTM G59-97.
tion. Ammonia 25% solution and HNO3 were used to maintain the pH The tests were performed after 30 min of immersion in 3.5 wt% NaCl
value of the plating solution around 4.7 during deposition. Similar solution with a scan rate of 10 mV/s within the potential range of − 250
procedures were performed to deposit the composite coating, except as to +1000 mV with respect to the open circuit potential (OCP). The
various amounts of GO nanoplatelets and/or SiC nanoparticles as second electrochemical set-up consists of a saturated calomel electrode (SCE)
phases are added to the plating bath. To ensure the complete deag­ reference electrode, a platinum counter electrode, and the samples as a
glomeration and dispersion of second phases, the plating solutions were working electrode. The Tafel extrapolation of polarization curves using
subjected to the high-intensity probe ultrasonic for 30 min. Further­ Zview software was used to determine the corrosion potential (Ecorr),
more, the electroless bath containing particles was magnetically stirred corrosion current density (Jcorr), and polarization resistance (Rp) of the
at 75 ± 5 ◦ C for 30 min without putting the substrate in order to increase samples.
the wettability of the GO nanoplatelets and/or SiC nanoparticles. All
samples were deposited for a time of 90 min. After the deposition, the 3. Results and discussion
samples were washed by distilled water, dried and heat-treated under Ar
atmosphere at temperature of 400 ◦ C for 1 h, which is the optimum To investigate the effect of incorporation of GO nanoplatelets and/or
condition for NiP electroless coatings [1,16,39]. The name of the sam­ SiC nanoparticles into the electroless Ni–P coating, the surface
ples and a short description of the preparation are presented in Table 1. morphology of the samples was observed by scanning electron micro­
The samples were used as prepared for morphology and structural in­ scope as shown in Fig. 1. As can be seen, a significant difference can be
vestigations, hardness measurements, corrosion and wear tests, whereas observed between the surface morphology of NiP and NiP-
the samples were cut with a diamond saw and mounted in epoxy resin nanocomposite coatings. The incorporation of GO nanoplatelets has no
for grinding using sandpapers of #800 to #1200 and sequent polishing significant effect on the typical nodular structure of Ni–P coatings
with alumina slurries (α-alumina (0.3 μm size) and gamma γ-alumina (Fig. 1b), which can be due to micron-size 2D dimensional GO nano­
(0.05 μm size) slurries) for cross-section observation. In addition, the 30 platelets. On the other hand, the ultrafine nodular structure was ach­
vol% HNO3 aqueous solution was used to etch the surface of the samples ieved by the incorporation of SiC nanoparticles (Fig. 1c), which can be
for structural characterization. resulted from the presence of SiC nanoparticles as nucleating sites. Also,
by co-deposition of GO nanoplatelets and SiC nanoparticles, the surface
2.3. Characterization morphology of the composite samples (Fig. 1d–f) shows the significant
nodular structure, resembling the cauliflower morphology. As can be
An optical microscope (Leica) was used to observe the cross-section seen in high magnification images at Fig. 1d–f, the ultrafine features on
of coatings. Surface morphology of samples, as well as wear track of the such cauliflower morphology of coatings can be recognized, which is
samples, was observed by scanning electron microscope (SEM; VEGAII because of the presence of SiC nanoparticles within the coatings. In
TESCAN). The field emission scanning electron microscope (FESEM; other words, in addition to the presence of GO nanoplatelets, the pres­
TESCAN-MIRA3) equipped with an attachment for energy dispersive ence of SiC nanoparticles can facilitate the nucleation. On the other
spectrometry (EDS) was used to determine the microstructure as well as hand, the nodular structure becomes irregular by increasing the amount
elemental analysis of the etched coatings. X-ray diffraction (XRD) by of SiC nanoparticles (NiP-GO-3SiC sample). In general, for all samples,
CuKα radiation with a scanning rate of 0.2◦ /min was performed using a the substrates were fully covered with a uniform and dense layer,
PHILIPS PW1730 diffractometer for the identification of phases in the whereas, the significant porosities, micro-voids, and cracks have been
coatings. reported in the previous works such as NiP–MoS2/Al2O3 [33] and
The hardness of the samples was measured by a Vickers micro- NiP-GO-TiO2 [37] coatings.
hardness tester according to the ASTM E384-11 with load of 100 g The cross-section of samples was observed by optical microscope as
(HV100) for a dwell time of 10 s. The hardness value was reported as an shown in Fig. 2S of supplementary data for the NiP, NiP-GO, NiP–2SiC,
average of five measurements after removing the highest and lowest and NiP-GO-2SiC samples. As can be seen, the thickness of the NiP
values. The wear behavior of the coatings was determined according to sample is about 11.2 μm and by incorporation of GO nanoplatelets and/
the ASTM G99-17 standard by pin-on-disc wear test. The wear tests were or SiC nanoparticles in the plating bath the thickness of the NiP-GO,
carried out at ambient conditions (temperature of about 25 ◦ C and NiP–2SiC, and NiP-GO-2SiC samples is around 14.2, 8.7, and 18.9 μm,
relative humidity of about 35%) by applying a load of 20 N against the respectively. Different thickness of NiP and NiP-nanocomposite samples
AISI-52100 steel pin with the hardness of 60 HRC under sliding speed of prepared by same plating time shows the incorporation of second-phase
1 m/s. The sliding distance was chosen 500 m and in order to determine within the coatings.
the weight loss of the sample during the test, the samples were weighted To confirm the presence of the GO nanoplatelets and/or SiC nano­
particles in the Ni–P electroless composite coatings, the surface of
Table 1 etched samples using 30 vol% HNO3 aqueous solution for 5 min was
The name of the samples and description of the fabrication routes of samples in investigated. Fig. 2 shows the FESEM images of the corroded surface of
this work. NiP–2SiC and NiP-GO-2SiC samples at different magnification along
Sample Description with EDS spectra as well as elemental EDS maps. For NiP–2SiC sample,
name as can be seen in Fig. 2a–c, the nanoparticles can be identified and
NiP Electroless deposition of pure Ni–P solution
further recognized at higher magnification. The EDS spectrum confirms
NiP-GO Electroless deposition of Ni–P solution containing of graphene oxide the presence of Si and C along with Ni and P elements (Fig. 2d) and the
nanoplatelets (30 mg/L) atomic and weight percentage of elements are presented in Fig. 2d. It is
NiP–2SiC Electroless deposition of Ni–P solution containing of SiC worth mentioning that the presence of Fe is related to the steel substrate
nanoparticles (2 g/L)
and oxygen can be due to the oxidation during chemical etching of the
NiP-GO- Electroless deposition of Ni–P solution containing of graphene oxide
1SiC nanoplatelets (30 mg/L) + SiC nanoparticles (1 g/L) coatings. Fig. 2e–h, the elemental mapping images corresponding to
NiP-GO- Electroless deposition of Ni–P solution containing of graphene oxide Fig. 2a, revealed that the Ni and P atoms are entirely distributed over the
2SiC nanoplatelets (30 mg/L) + SiC nanoparticles (2 g/L) whole surface and the Si and C atoms are also presented in the surface of
NiP-GO- Electroless deposition of Ni–P solution containing of graphene oxide the NiP–2SiC sample, indicating the relatively uniform distribution of
3SiC nanoplatelets (30 mg/L) + SiC nanoparticles (3 g/L)
the embedded SiC nanoparticles. In addition, the etched surface of the

3
M. Khodaei and A.M. Gholizadeh Ceramics International xxx (xxxx) xxx

Fig. 2. FESEM images, EDS spectra, and EDS elemental maps of the HNO3 etched surface (a–h) of NiP–2SiC sample and (i–p) of NiP-GO-2SiC sample. (a–c) and (i–k)
FESEM images at different magnification; (d) and (l) EDS spectra and element weight and atomic percentage; (e–h) and (m–p) EDS elemental map corresponding to
the FESEM image at (a) and (i), respectively.

NiP-GO-2SiC sample (Fig. 3i–k) shows the presence of particulate sub­ similar observation was reported elsewhere for NiP-GO coatings [24]. In
stances resembling the embedded micron-size 2D dimensional GO addition, the EDS analysis of the NiP and NiP-GO samples are also
nanoplatelets. In order to confirm the presence of GO as well as SiC performed and their EDS spectra along with elemental composition were
nanoparticles within the Ni–P matrix, the EDS analysis was also per­ presented in Fig. 3S. The EDS spectrum of the NiP sample showed no
formed, which reveals the presence of C, Si, Ni, and P elements (Fig. 2l), peak related to the Si and C, whereas that of the NiP-GO showed the
as tabulated at Fig. 2l. A higher weight percent of C in the NiP-GO-2SiC peak of C. Along with the peak of Ni and P, the peak of Fe and O are
sample in comparison to that in the NiP–2SiC sample demonstrates the presented for both samples. As mentioned before, the presence of the
presence of GO. Furthermore, the elemental mapping images peak of Fe can be due to steel substrate and the peak of O can be due to
(Fig. 2m–p) corresponding to Fig. 2i revealed the presence of GO the oxidation of the coatings during the sample preparation (acidic
nanoplatelets in the Ni matrix, which the area with lack of Ni (Fig. 2m) etching process).
as well as area with concentrated C (Fig. 2p) can be identified, as indi­ To investigate the effect of the presence of the GO nanoplatelets and/
cated by the dashed line in the Ni and C elemental mapping images. It is or SiC nanoparticles on the crystallite structure of Ni–P coatings, the X-
worth mentioning again that the GO nanoplatelets has a micron-size ray diffraction (XRD) analysis has been performed for the NiP, NiP-GO,
lateral dimension. In addition, the Si elemental map shows the pres­ NiP–2SiC, and NiP-GO-2SiC samples, and the corresponding XRD pat­
ence of SiC within the Ni matrix as well as on the GO nanoplatelets. A terns are presented in Fig. 3a. It can be seen that the heat-treated pure

4
M. Khodaei and A.M. Gholizadeh Ceramics International xxx (xxxx) xxx

Fig. 3. XRD pattern of the heat-treated NiP, NiP–2SiC, NiP-GO, NiP-GO-2SiC samples.

Ni–P layer (NiP sample) composed of the Ni phase with face-centered continued upon the achieving the convergence and reaching to the Rwp
cubic (FCC) crystal structure as well as precipitated nickel phosphide factor of about 20%. The refined plots for all samples are presented in
(Ni3P) with body-centered tetragonal (BCT) crystal, which is precipi­ Fig. 4S of supplementary data and the calculated contents of phases are
tated during heat treatment of Ni(P) solid solution. The Ni and Ni3P presented in Table 2. Using similar method, it was shown [41] that the
phases were identified by JCPDS no. 00-004-0850 and 01-074-1384, nucleation of Ni3P will start upon heat treatment above 300 ◦ C and the
respectively. The optimum heat treatment condition for the commer­ content of Ni3P reach to about 50–60 wt% as calculated based of XRD
cial electroless Ni–P has been reported to be 400 ◦ C for 1 h [39] for analysis.
complete crystallization of Ni matrix and precipitation of Ni3P phases Furthermore, it is worth mentioning that the position of diffraction
within the matrix. To have a better comparison, the heat treatment for peaks in the XRD pattern of the samples are different to some extent
the other samples was also selected same as the heat treatment condition which is indicating of different stress state of constituent phases. The
of pure NiP samples, which is also reported previously as an optimum peak position of Ni (111) and Ni (200), which are around diffraction
condition for NiP–SiC [39] and NiP-GO [25] composite electroless angle of 2θ ≈ 44.6◦ and 2θ ≈ 51.9◦ , are presented in Fig. 3b and c,
coatings. The formation of intermetallic phases between Ni and Si respectively. As can be seen, the peaks have been shifted to the higher
(NixSiy) by heat treatment at a higher temperature than 400 ◦ C and value by the incorporation of SiC nanoparticles (NiP–2SiC sample)
longer time than 1 h has been reported in NiP–SiC [39]. As can be seen, relative to the peaks of pure NiP sample indicating the decrease in the d-
the heat-treated NiP–2SiC sample has also consisted of Ni and Ni3P spacing of Ni, which could be resulted from the more residual stress due
phases, similar to the NiP sample, except as representing the higher to the presence of SiC nanoparticles within the Ni matrix. On the other
intensity for the diffraction peaks of Ni3P in the XRD pattern of the hand, the peak position of Ni in the NiP-GO sample is lower than that of
NiP–2SiC sample in comparison to those of the NiP sample. This in­ NiP and NiP–2SiC samples showing the higher d-spacing, i.e., lower
dicates the higher precipitation of Ni3P phases during heat treatment as residual stress, which can be resulted from the presence of GO nano­
results of presence of SiC nanoparticles which can promote the hetero­ platelets acting as a stress releasing media. Furthermore, the presence of
geneous nucleation of Ni3P. On the other hand, Ni2P and Ni12P5 meta­ SiC nanoparticles along with GO nanoplatelets in the Ni matrix (NiP-GO-
stable phases are presented in NiP-GO sample after heat treatment at 2SiC sample) resulted in an increase in the XRD peak position of Ni in
temperature of 400 ◦ C for 1 h, which were identified by JCPDS no. comparison to the NiP-GO sample. Hence, it can be concluded that the
01-074-1385 and 01-074-1381, respectively. Presence of such meta­
stable phases have been also reported in the heat-treated NiP even at
temperature of 400 ◦ C for 1 h [40] which is due to the incomplete
transformation of the amorphous plated Ni matrix to the mixture of
crystalline Ni and the Ni3P final phases. Hence, the incomplete precip­
itation of Ni3P (presence of intermediated Ni2P and Ni12P5 metastable
phases) may be due to the presence of GO nanoplatelets that suppress
the transformation process. In NiP-GO-2SiC sample, the amount of
intermediated metastable phases is significantly reduced as a result of
the presence of SiC nanoparticles, i.e., the presence of nucleation sites,
within the coating. As can be seen in the XRD pattern of NiP-GO-2SiC
sample, the peaks related to the Ni12P5 metastable phase have been
vanished and the peak of Ni2P phase with low intensity has only been
remained.
To identify the content of phases in different samples by assuming
the diffraction intensities, the refinement of XRD pattern has been per­
formed. In order to refine the diffraction patterns, Rietveld method was
employed using the Maud program. The modified pseudo-Voigt profile
function was used to model the peak profiles. Refinements were Fig. 4. Microharness of the heat-treated samples (vertical error bars. represent
standard deviations).

5
M. Khodaei and A.M. Gholizadeh Ceramics International xxx (xxxx) xxx

Table 2 hardened steel pin as a counter-face. The amount of applied load during
The phase content of samples calculated based on XRD pattern refinement by the wear test was chosen as high as 20 N to achieve comparable weight
Rietveld method using the Maud program. loss during abrasion. The lower applied loads resulted in no significant
Sample Ni (wt%) Ni3P (wt%) Ni2P (wt%) Ni12P5 (wt%) effect on some samples. The variation of the average coefficient of
NiP 41.8 58.2 – –
friction and the weight loss (measured every 100 m sliding distance)
NiP-GO 45.9 46.2 6.1 1.8 during the sliding for all samples are demonstrated in Fig. 5a and b,
NiP–2SiC 37.8 62.2 – – respectively. In addition, SEM images of the worn surface of the samples
NiP-GO-2SiC 42.9 57.1 2.1 – after the first 100 m sliding distance are shown in Fig. 6. As can be seen
in Fig. 5a, the coefficient of friction of the samples reaches their rela­
tively stable values after the initial sliding distance (less than the first
presence of SiC nanoparticles in the Ni matrix leads to the nucleation of
100 m sliding distance), which can be due to the relatively high applied
Ni crystalline structure with lower d-spacing after heat treatment,
load (20 N) in these investigations.
whereas the presence of GO nanoplatelets leads to the higher d-spacing
The average coefficient of friction of the NiP sample (μNiP = 0.702) is
in comparison to the pure NiP sample.
relatively high, whereas the incorporation of GO nanoplatelets leads to
Microhardness measurement was performed to investigate the effect
the significant reduction in coefficient of friction to μNiP-GO = 0.417,
of the presence of the GO nanoplatelets and/or SiC nanoparticles in the
which can be attributed to the formation of an interfacial worn surface
Ni–P coatings. It is worth mentioning that microhardness measurements
containing the GO nanoplatelets acting as a solid lubricant film between
were also used to check the repeatability of the fabrication of the sam­
the surface and counter pin. It is worth mentioning that the NiP-GO
ples. Fig. 4 compares the microhardness of the samples, which are re­
samples have been fabricated using NiP bathes with different concen­
ported as an average of 5 measurements. It worth mentioning that the
trations of GO nanoplatelets (20, 30, and, 40 mg/L). Among them, the
hardness of the NiP sample before heat treatment was 582 ± 10 HV100
sample prepared by 30 mg/L GO nanoplatelets showed the lowest wear
whereas, the hardness reaches 1007 ± 12 HV100 after heat treating.
loss and coefficient of friction, which is selected as an optimum condi­
Generally, the hardness of the heat-treated Ni–P electroless coatings is a
tion for preparing the NiP-GO sample in this work. The variation of the
result of crystallization of amorphous Ni–P metallic matrix as well as
average coefficient of friction and the weight loss (measured every 100
Ni3P precipitation in the matrix (i.e. the precipitation hardening
m sliding distance) during the sliding for the NiP-xGO (x = 20, 30, 40)
mechanism) [1]. On the other hand, the hardness of the heat-treated
samples are demonstrated in Fig. 5S of supplementary data.
NiP-GO sample (893 ± 21 HV100) is lower than that of the NiP sample,
This reduction in the coefficient of friction of metallic matrix con­
which can be due to the presence of GO nanoplatelets acting as a soft
taining the graphene especially for Ni based coatings has been exten­
counterpart within the matrix. Both higher [25] and lower [26, 42]
sively reported elsewhere [1,25]. This difference in coefficient of
values for the hardness of heat-treated NiP-Graphene nanocomposite
friction can be easily recognized by comparison of the morphology of the
coatings in comparison to the pure NiP coatings, which were prepared in
worn surface for the NiP (Fig. 6a) and NiP-GO (Fig. 6b). The worn sur­
the same conditions, have been reported. A decrease in the hardness of
face of NiP sample after the first 100 m sliding distance (Fig. 6a) contains
NiP coatings by incorporation of soft counterparts such as WS2 [20],
the micro-cracks, irregular pits, deep grooves cross to the sliding di­
PTFE [18], Graphite [21] has been reported elsewhere. Furthermore,
rection resembling the severe plastic deformed areas showing the typical
lower hardness of the NiP-GO sample can be due to the incomplete
precipitation of Ni3P, which has been detected by the presence of
intermediated Ni2P and Ni12P5 metastable phases confirming by XRD
(Fig. 3). On the other hand, the incorporation of hard ceramic nano­
particles (SiC nanoparticles) within the Ni matrix (NiP–2SiC sample)
leads to the significant increase in the hardness to 1237 ± 27 HV100,
which has been reported because of retarding the plastic deformation of
the Ni matrix, i.e. dispersion hardening mechanism [1,3]. Grain
refinement of the Ni matrix as a result of nucleation agents (SiC nano­
particles), which can be easily recognized by comparison of Fig. 1a and
c, could be another reason for increasing the hardness. The reduction of
grain size, i.e., higher grain boundaries, leads to the existence of higher
barriers to the dislocation motion. Such increase in the microhardness of
NiP nanocomposite coating by incorporation of ceramic nanoparticles
has been numerously reported elsewhere [7,15].
In addition, presence of ceramic nanoparticles within the NiP/soft-
particle composite coatings (such as heat-treated NiP/PTFE-Al2O3 [34]
as well as NiP/GO-TiO2 nanocomposite [37]) can increase the hardness.
In this work, the hardness of NiP-GO sample is increased by the addition
of SiC nanoparticles (NiP-GO-xSiC samples). As can be seen in Fig. 4, the
hardness of NiP-GO-xSiC samples is higher than that of NiP-GO sample,
which can be realized as different effects of SiC nanoparticles including
the dispersion hardening mechanism, grain refining (as can be seen in
Fig. 1d–f), and complete precipitation of Ni3P (significant reduction in
the presence of intermediated Ni2P and Ni12P5 metastable phases) as
confirming by XRD (Fig. 3). It worth to mention that the lower hardness
of NiP-GO-3SiC sample can be due to the increase in the volume fraction
of reinforcements (GO nanoplatelets and SiC nanoparticles), which leads
to create a discontinuity in the grain structure of Ni matrix as can be
detected by non-uniform surface morphology (Fig. 1f).
The tribological behavior of the samples was investigated using an
unlubricated “pin on disc” wear test at room temperature using a Fig. 5. (a) coefficient of friction and (b) weight loss as function of the sliding
distance for samples.

6
M. Khodaei and A.M. Gholizadeh Ceramics International xxx (xxxx) xxx

addition of SiC nanoparticles in the Ni matrix (NiP–2SiC sample) is also


resulted in the reduction of coefficient of friction (μNiP-2SiC = 0.611), but
not as much as the addition of GO nanoplatelets (NiP-GO sample). Such
reduction in coefficient of friction can result from the increase in the
hardness of coatings by the presence of SiC nanoparticles leading to the
lowering the formation of wear asperities as well as the lowering the
contact area between the surface of the sample and pin. Reduction of
coefficient of friction of Ni based coatings by incorporation of ceramic
nanoparticles such as TiO2 [5], SiO2 [7], and SiC [15] and been previ­
ously reported. In addition, the weight loss of the NiP–2SiC during
sliding up to 300 m sliding distance is to some extent similar to that of
the NiP sample and the weight loss is reduced significantly at a higher
sliding distance (Fig. 5b). As can be seen in Fig. 6, the presence of SiC
nanoparticles in the Ni matrix (NiP–2SiC sample) leads to the formation
of a relatively narrower wear track in comparison to that of NiP sample
as a result of higher hardness. In addition to this, the micro-cracks as
well as deep grooves cross to the sliding direction can be also detected
showing the same wear mechanism of the NiP–2SiC sample with NiP
sample. Although the same wear mechanism is existing for both the NiP
and NiP–2SiC, the higher hardness of NiP–2SiC leads to the lower weight
loss during the pin on disc sliding at a high distance (higher than 300 m).
The effect of addition of SiC nanoparticles into the NiP-GO matrix on
its tribological behavior is also investigated. The coefficient of friction is
remained unchanged to some extend for the NiP-GO-1SiC and NiP-GO-
2SiC samples (μNiP-GO-1SiC = 0.423 and μNiP-GO-2SiC = 0.459) in com­
parison to that for NiP-GO (μNiP-GO = 0.417), whereas the coefficient of
friction is significantly increased by a further increase in SiC content
(μNiP-GO-3SiC = 0.719). As can be seen in Fig. 6d and e, after the 100 m of
sliding distance, the worn surface of NiP-GO-1SiC and NiP-GO-2SiC
samples is relatively smooth, whereas the worn surface of NiP-GO-
3SiC has a higher roughness. After the 100 m of sliding distance, the
weight loss of NiP, NiP–2SiC, and NiP-GO-3SiC samples (which is similar
to each other) is higher than that of NiP-GO, NiP-GO-1SiC, and NiP-GO-
2SiC samples. After increase in the sliding distance, the rate of weight
loss for NiP, NiP–2SiC and NiP-GO-3SiC samples is increased, whereas
the wear rate of NiP-GO, NiP-GO-1SiC, and NiP-GO-2SiC samples are to
some extent remained constant and even it is reduced for the NiP-GO-
2SiC sample. In the other words, since the general reduction of the
weight loss during the sliding can be obtained by the presence of GO
nanoplatelets within the NiP matrix, the wear rate of NiP-GO-2SiC at
high sliding distance remains unchanged. Furthermore, the wear rate at
high sliding distance can be even decreased by co-deposition of the GO
nanoplatelets and SiC nanoparticles (NiP-GO-2SiC sample). The lower
wear rate by addition of SiC nanoparticles to the NiP-GO, can be resulted
from the increasing of the hardness as well as the presence of GO
nanoplatelets as a solid lubricant. Further increase in SiC content to the
NiP-GO (NiP-GO-3SiC sample) leads to the increase in the weight loss
during sliding, which can be due to the effect of high-volume fraction of
GO nanoplatelets and SiC nanoparticles resulting in the discontinuity in
the grain structure of Ni matrix. It is worth mentioning that by
increasing the volume fraction of SiC nanoparticles in the plating bath to
4 g/L along with the GO nanoplatelets (30 mg/L), the surface of coating
was non-uniform as can be seen in the surface morphology image, which
is presented in Fig. 6S of supplementary data. This observation can
reveal that the higher volume fraction of additives leads to lower
Fig. 6. SEM micrograph of the worn surface of the (a) NiP, (b) NiP-GO, (c) reduction of Ni2+ ions during the coating and can create the more
NiP–2SiC, (d) NiP-GO-1SiC, (e) NiP-GO-2SiC, and (f) NiP-GO-3SiC samples at discontinuity in the grain structure of Ni matrix.
two different magnifications.
The effect of incorporation of the GO nanoplatelets and/or SiC
nanoparticles within the Ni matrix on the corrosion resistance of the
characteristics of severe abrasive wear, whereas the worn surface of electroless Ni–P coating was studied by the electrochemical method. The
NiP-GO is rather smooth with shallow longitudinal grooves along the potentiodynamic polarization curves of the samples in the 3.5 wt% NaCl
sliding direction indicating mild abrasive condition. The lower hardness are shown in Fig. 7.
of NiP-GO (higher plastic deformation) and the presence of GO nano­ The corrosion current density (Jcorr) and corrosion potential (Ecorr) of
platelets (acting as a solid lubricant) are the main reasons for smooth the samples were determined extrapolating of the linear sections of the
wear track. In addition, the weight loss during sliding for the NiP-GO is anodic and cathodic Tafel lines. These results along with the values of
significantly lower than that for the NiP sample (Fig. 5b). In contrast, the cathodic (βc) and anodic (βa) Tafel slopes are presented in Table 3. In

7
M. Khodaei and A.M. Gholizadeh Ceramics International xxx (xxxx) xxx

(H2PO−2 ) layer [28]. This layer can prevent the penetration of the water
and corrosive ions into the metal layer. In addition, the Ecorr of nano­
composite samples shifted to the positive direction. Incorporation of the
GO nanoplatelets and/or SiC nanoparticles in the heat-treated Ni–P
coating (containing the Ni matrix and Ni3P dispersing phase) can form
the Ni/GO-SiC and/or Ni/Ni3P-GO-SiC corrosion microcells, which the
formation of such microcells has been also revealed elsewhere for other
composite metallic coatings during corrosion [14]. In these galvanic
microcells, nickel would act as an anode and consequent dissolution of
nickel is preferred leading to move corrosion potential of nanocomposite
coating toward the noble direction. Hence, in the nanocomposite sam­
ples, the accelerated formation of a nickel passive film at the surface of
the coating, i.e., the thicker phosphorus-rich film, could be a reason of
their higher corrosion resistance in comparison to the NiP sample.
Furthermore, incorporated reinforcements act as grain refiner during
the plating of Ni matrix as well as during the subsequent heat treatment
leading to the creation of more grain boundaries, which promote the
formation of passive layer. Formation of such passive layers on the
surface resulted in decrease in oxygen permeability and corrosive ion
diffusion through the coatings. In addition, the presence of re­
inforcements within the coating, especially GO nanoplatelets, can act as
a physical barrier to the penetration of corrosive ions leading to retard
the corrosion process. It is reported that the presence of rGO [36] or GO
[37] nanoplatelets in the NiP–TiO2 nanocomposite coating leads to the
Fig. 7. Potentiodynamic polarization curves of the heat-treated samples in 3.5 significant reduction in the corrosion current density as a result of their
wt% NaCl solution. barrier effect. As can be seen in Table 3, the Jcorr of NiP-GO sample is
lower than that of NiP–2SiC sample and co-deposition of GO nano­
platelets and SiC nanoparticles resulted in to some extent similar value
Table 3
in Jcorr as well as corrosion inhibition efficiency. It can be concluded that
Anodic (βa) and cathodic (βc) Tafel slopes, corrosion potential (Ecorr), corrosion
the incorporation of GO nanoplatelets in the NiP and
current density (Jcorr) of the samples calculated by Tafel extrapolation of
NiP/ceramic-nanoparticle coatings can be significantly effective in
Potentiodynamic polarization curves along with polarization resistance (Rp) and
corrosion inhibition efficiency (IE %) of the samples. corrosion mitigation.

Sample βa (mV/ βc (mV/ Ecorr (mV Jcorr Rp IE%


decade) decade) vs. SCE) (μAcm− 2) (kΩ.
4. Conclusions
cm)
The effects of co-deposition of GO nanoplatelets and/or SiC nano­
NiP 97.5 87.3 − 389.6 2.97 6.7 0
NiP-GO 104.5 100.7 − 320.0 1.04 21.4 64.9 particles within the electroless deposited Ni layers (NiP, NiP-GO,
NiP–2SiC 92.8 90.5 − 371.6 2.14 9.3 27.9 NiP–SiC, and NiP-GO-xSiC samples) were investigated on the struc­
NiP-GO- 83.4 85.3 − 334.2 1.85 9.9 37.7 ture, morphology, hardness, tribological and corrosion behavior of the
1SiC heat-treated coatings. The results revealed that the heat-treated NiP and
NiP-GO- 98.8 98.1 − 248.2 1.18 18.1 60.3
2SiC
NiP–SiC coatings consisted of Ni and Ni3P phases, whereas the NiP-GO
NiP-GO- 101.3 100.5 − 277.4 0.98 22.4 67.0 also contains the intermediated Ni2P and Ni12P5 metastable phases
3SiC due to the incomplete precipitation of Ni3P. Such metastable phases are
significantly decreased by the incorporation of SiC nanoparticles in NiP-
GO coatings (in NiP-GO-xSiC). By co-deposition of SiC nanoparticles, the
addition, the value of the polarization resistance (Rp) and the corrosion
hardness of NiP-GO (893 ± 21 HV100) is significantly increased to 1237
inhibition efficiency (IE %) of the presence of GO nanoplatelets and/or
± 27 HV100. The average friction coefficient of the NiP sample (μNiP =
SiC nanoparticles within the coatings were calculated using the Stren-
0.702) is reduced significantly by incorporation of GO nanoplatelets
Geary equation (Eq. (1)) and Eq. (2), respectively, and summarized in
(μNiP-GO = 0.417), which can be attributed to the role of GO nano­
Table 3.
platelets as a solid lubricant part. By incorporation of SiC nanoparticles,
βa .βc the coefficient of friction is remained unchanged to some extend at
Rp = (1)
2.303Jcorr (βa + βc ) lower concentrations (μNiP-GO-1SiC = 0.423 and μNiP-GO-2SiC = 0.459),
whereas the coefficient of friction is significantly increased by further
( )
Jcorr(x) increase in the SiC content (μNiP-GO-3SiC = 0.719). Although the wear loss
IE% = 1 − × 100 (2)
Jcorr(0) of NiP-GO is lower than that of NiP sample, it is reduced further espe­
cially at high sliding distance during wear test by incorporation of op­
where the Jcorr(x) and Jcorr(0) are the corrosion current densities for the timum content of SiC nanoparticles. Furthermore, the corrosion
nanocomposite samples and NiP sample, respectively. resistance of nanocomposite samples is higher than that of pure NiP
According to Table 3, all nanocomposite samples showed higher sample (28–67% corrosion inhibition efficiency). The co-deposition of
polarization resistance and lower corrosion current densities than NiP GO nanoplatelets and SiC nanoparticles within the NiP layer resulted in
sample, which demonstrates the significant role of incorporation GO efficient corrosion mitigations along with improvement in tribological
nanoplatelets and/or SiC nanoparticles on the electrochemical behavior behavior.
of the Ni–P coating. Generally, the main reported mechanism of corro­
sion resistance of electroless Ni–P coatings is the formation of the Declaration of competing interest
enriched phosphorus layer as a result of preferential dissolution of nickel
that can react by water to form an adsorbed hypophosphite anions The authors declare that they have no known competing financial

8
M. Khodaei and A.M. Gholizadeh Ceramics International xxx (xxxx) xxx

interests or personal relationships that could have appeared to influence [19] X.G. Hu, et al., Electroless Ni–P–(nano-MoS2) composite coatings and their
corrosion properties, Surf. Eng. 25 (5) (2009) 361–366.
the work reported in this paper.
[20] I. Sivandipoor, F. Ashrafizadeh, Synthesis and tribological behaviour of electroless
Ni–P-WS2 composite coatings, Appl. Surf. Sci. 263 (2012) 314–319.
Appendix A. Supplementary data [21] M. Ananth Kumar, R.C. Agarwala, V. Agarwala, Synthesis and characterization of
electroless Ni-P coated graphite particles, Bull. Mater. Sci. 31 (5) (2008) 819–824.
[22] M.C.L. de Oliveira, et al., Structural, adhesion and electrochemical characterization
Supplementary data to this article can be found online at https://doi. of electroless plated Ni-P-carbon black composite films on API 5L X80 steel,
org/10.1016/j.ceramint.2021.05.250. J. Mater. Eng. Perform. 28 (8) (2019) 4751–4761.
[23] Z. Gao, et al., Corrosion behavior and wear resistance characteristics of electroless
Ni–P–CNTs plating on carbon steel, Int. J. Electrochem. Sci 10 (15) (2015)
References 637–648.
[24] H. Wu, et al., Preparation of Ni–P–GO composite coatings and its mechanical
[1] J.K. Pancrecious, et al., Metallic composite coatings by electroless technique – a properties, Surf. Coating. Technol. 272 (2015) 25–32.
critical review, Int. Mater. Rev. 63 (8) (2018) 488–512. [25] J. Jiang, et al., Effect of heat treatment on structures and mechanical properties of
[2] W. Sha, X. Wu, K.G. Keong, 1 - introduction to electroless copper and electroless Ni–P–GO composite coatings, RSC Adv. 6 (110) (2016)
nickel–phosphorus (Ni–P) depositions, in: W. Sha, X. Wu, K.G. Keong (Eds.), 109001–109008.
Electroless Copper and Nickel–Phosphorus Plating, Woodhead Publishing, 2011, [26] M.H. Sadhir, et al., Comparison of in situ and ex situ reduced graphene oxide
pp. 1–12. reinforced electroless nickel phosphorus nanocomposite coating, Appl. Surf. Sci.
[3] J. Anupam, et al., Electroless coating on non-conductive materials: a review, in: 320 (2014) 171–176.
R. Supriyo, B. Goutam Kumar (Eds.), Advanced Surface Coating Techniques for [27] T.R. Tamilarasan, et al., Effect of reduced graphene oxide (rGO) on corrosion and
Modern Industrial Applications, IGI Global, Hershey, PA, USA, 2021, pp. 188–208. erosion-corrosion behaviour of electroless Ni-P coatings, Wear 390–391 (2017)
[4] C.Y. Ma, et al., Ultrasonic-assisted electrodeposition of Ni-Al2O3 nanocomposites 385–391.
at various ultrasonic powers, Ceram. Int. 46 (5) (2020) 6115–6123. [28] B. Elsener, et al., Electroless deposited Ni–P alloys: corrosion resistance
[5] S. Amjad-Iranagh, M. Zarif, TiO2 nano-particle effect on the chemical and physical mechanism, J. Appl. Electrochem. 38 (7) (2008) 1053.
properties of Ni-P-TiO2 nanocomposite electroless coatings, Journal of [29] A. Gergely, A review on corrosion protection with single-layer, multilayer, and
Nanostructures 10 (2) (2020) 415–423. composites of graphene, Corrosion Rev. 36 (2) (2018) 155–225.
[6] P. Makkar, R.C. Agarwala, V. Agarwala, Chemical synthesis of TiO2 nanoparticles [30] A. Radhamani, H.C. Lau, S. Ramakrishna, Nanocomposite coatings on steel for
and their inclusion in Ni–P electroless coatings, Ceram. Int. 39 (8) (2013) enhancing the corrosion resistance: a review, J. Compos. Mater. 54 (5) (2020)
9003–9008. 681–701.
[7] A. Sadeghzadeh-Attar, G. AyubiKia, M. Ehteshamzadeh, Improvement in [31] M. Ni, et al., A novel self-lubricating Ni-P-AlN-WS2 nanocomposite coating, Mater.
tribological behavior of novel sol-enhanced electroless Ni-P-SiO2 nanocomposite Res. Express 6 (11) (2019) 116413.
coatings, Surf. Coating. Technol. 307 (2016) 837–848. [32] E. Freshteh Amjadi, M. Azadi, H. Tavakoli, Effect of SiO2 nanoparticles addition on
[8] D.R. Dhakal, et al., Understanding the effect of Si3N4 nanoparticles on wear tribological and electrochemical behaviors of Ni–P–MoS2 multi-component
resistance behavior of electroless Nickel-Phosphorus coating through structural coatings after heat treatment, Surf. Eng. Appl. Electrochem. 56 (2) (2020)
investigation, Appl. Surf. Sci. 541 (2021) 148403. 171–183.
[9] M.H. Sliem, et al., Enhanced mechanical and corrosion protection properties of [33] P. Liu, et al., Surfactant-free electroless codeposition of Ni–P–MoS2/Al2O3
pulse electrodeposited NiP-ZrO2 nanocomposite coatings, Surf. Coating. Technol. composite coatings, Coatings 9 (2) (2019) 116.
403 (2020) 126340. [34] C. Suiyuan, et al., Synthesis of Ni-P-PTFE-nano-Al2O3 composite plating coating on
[10] P. Makkar, et al., A novel electroless plating of Ni–P–Al–ZrO2 nanocomposite 45 steel by electroless plating, J. Compos. Mater. 46 (12) (2012) 1405–1416.
coatings and their properties, Ceram. Int. 40 (8, Part A) (2014) 12013–12021. [35] T.R. Tamilarasan, et al., Effect of reduced graphene oxide reinforcement on the
[11] M. Czagány, P. Baumli, Effect of surfactants on the behavior of the Ni-P bath and wear characteristics of electroless Ni-P coatings, J. Mater. Eng. Perform. 27 (6)
on the formation of electroless Ni-P-TiC composite coatings, Surf. Coating. (2018) 3044–3053.
Technol. 361 (2019) 42–49. [36] N. Promphet, P. Rattanawaleedirojn, N. Rodthongkum, Electroless NiP-TiO2 sol-
[12] C. Ma, et al., Magnetic assisted pulse electrodeposition and characterization of RGO: a smart coating for enhanced corrosion resistance and conductivity of steel,
Ni–TiC nanocomposites, Ceram. Int. 46 (11, Part A) (2020) 17631–17639. Surf. Coating. Technol. 325 (2017) 604–610.
[13] K. Shahzad, et al., Effect of concentration of TiC on the properties of pulse [37] M. Uysal, Electroless codeposition of Ni-P composite coatings: effects of graphene
electrodeposited Ni-P-TiC nanocomposite coatings, Ceram. Int. (2021). and TiO2 on the morphology, corrosion, and tribological properties, Metall. Mater.
[14] H. Ashassi-Sorkhabi, M. Es′ haghi, Corrosion resistance enhancement of electroless Trans. 50 (5) (2019) 2331–2341.
Ni–P coating by incorporation of ultrasonically dispersed diamond nanoparticles, [38] H. Yin, et al., Preparation and tribological properties of graphene oxide doped
Corrosion Sci. 77 (2013) 185–193. alumina composite coatings, Surf. Coating. Technol. 352 (2018) 411–419.
[15] N. Ghavidel, et al., Corrosion and wear behavior of an electroless Ni-P/nano-SiC [39] G. Jiaqiang, et al., Electroless Ni–P–SiC composite coatings with superfine
coating on AZ31 Mg alloy obtained through environmentally-friendly conversion particles, Surf. Coating. Technol. 200 (20–21) (2006) 5836–5842.
coating, Surf. Coating. Technol. 382 (2020) 125156. [40] M. Buchtík, et al., The effect of heat treatment on properties of Ni–P coatings
[16] C. Ma, et al., Effect of heat treatment on structures and corrosion characteristics of deposited on a AZ91 magnesium alloy, Coatings 9 (7) (2019) 461.
electroless Ni–P–SiC nanocomposite coatings, Ceram. Int. 40 (7, Part A) (2014) [41] Y. Tan, et al., Crystallization mechanism analysis of noncrystalline Ni–P
9279–9284. nanoparticles through XRD, HRTEM and XAFS, CrystEngComm 16 (41) (2014)
[17] F. Xia, et al., Preparation and wear properties of Ni/TiN–SiC nanocoatings obtained 9657–9668.
by pulse current electrodeposition, Ceram. Int. 46 (6) (2020) 7961–7969.
[18] Z.K. Karaguiozova, Characterisation of electroless Ni-P and electroless composite
coatings Ni-P/Ni-PTFE, Int. J. Surf. Sci. Eng. 12 (5–6) (2018) 496–506.

You might also like