You are on page 1of 97

Materials Chemistry and Physics

CeO2 modified reduced graphene oxide nanocomposites for tribology, anti-corrosion


and photocatalysis applications
--Manuscript Draft--

Manuscript Number: MATCHEMPHYS-D-22-05852R1

Article Type: Full Length Article

Keywords: Graphene; rare earth metals; Lubrication; corrosion resistance; photocatalysis

Abstract: CeO2 modified reduced graphene oxide (CeO2@rGO) nanocomposites synthesized


using a hydrothermal method were studied for their performance and mechanisms for
tribology, anti-corrosion and photocatalysis applications. Results showed that ceria
nanoparticles with an average diameter of ~12.24 nm were uniformly distributed and
covalently bonded onto the surface of rGO. When the nanocomposites were added
into PAO5w-40 lubricating oil, the coefficient of friction was 31.9% lower than that of
pure lubricating oil. The corresponding wear mechanism was changed from oil-film
lubrication to reaction film lubrication, which is mainly attributed to the dense filling
effect of the nanostructured CeO2 and the lubrication effect of rGO. The
nanocomposites were also added into chromium-free Dacromet coating and the
corrosion current density in the corrosion test was decreased by an order of
magnitude. This is mainly due to the formation of various bonds and passivation layers
in the modified coating, which improves its density. It is also attributed to the uniformly
dispersed rGO, which forms a conductive channel, resulting in the preferred corrosion
of the Zn-Al alloy in the coating thus achieving an effective protection effect to the
substrate. When the composites were applied as the photocatalysts in an RhB solution
under the irradiation of ultraviolet light, the RhB dye was effectively removed with an
efficiency up to 90.5%, which is mainly because the rGO can extend the life time of
active electron-hole pairs and prevent CeO2 nanoparticles from agglomeration.

Powered by Editorial Manager® and ProduXion Manager® from Aries Systems Corporation
Title Page

Title Page

Authors
Bo Sun1, Wenge Chen1,*, Hui Zhang1, Tao Feng1, Wanqing Xing1, Ahmed Elmarakbi2,
Yong-Qing Fu2,*

Author affiliations
1
School of Materials Science and Engineering, Xi’an University of Technology, Xi’an,
Shaanxi, 710048, P.R. China
2
Faculty of Engineering and Environment, Northumbria University, Newcastle upon
Tyne, NE1 8ST, UK.
*Corresponding authors: Prof. Wenge Chen; Prof. Richard Yongqing Fu.
E-mail: wgchen001@263.net (W.G. Chen), richard.fu@northumbria.ac.uk (Y.Q. Fu)

Acknowledgement
The authors would like to acknowledge the financial supports from the Science
and Technology Project of Shaanxi Province of China (No.2020ZDLGY12-06), Xi’an
Science research project of China (No.2021XJZZ0042) and International Exchange
Grant (IEC/NSFC/201078) through Royal Society and National Science Foundation of
China (NSFC).
Revised Manuscript (clean copy) Click here to view linked References

CeO2 modified reduced graphene oxide nanocomposites for

tribology, anti-corrosion and photocatalysis applications

Bo Sun1, Wenge Chen1,*, Hui Zhang1, Tao Feng1, Wanqing Xing1, Ahmed Elmarakbi2,
Yong-Qing Fu2,*

1
School of Materials Science and Engineering, Xi’an University of Technology, Xi’an,

Shaanxi, 710048, P.R. China

2
Faculty of Engineering and Environment, Northumbria University, Newcastle upon

Tyne, NE1 8ST, UK.


*Corresponding authors: Prof. Wenge Chen; Prof. Richard Yongqing Fu.

E-mail: wgchen001@263.net (W.G. Chen), richard.fu@northumbria.ac.uk (Y.Q. Fu)

Abstract

CeO2 modified reduced graphene oxide (CeO2@rGO) nanocomposites

synthesized using a hydrothermal method were studied for their performance and

mechanisms for tribology, anti-corrosion and photocatalysis applications. Results

showed that ceria nanoparticles with an average diameter of ~12.24 nm were uniformly

distributed and covalently bonded onto the surface of rGO. When the nanocomposites

were added into PAO5w-40 lubricating oil, the coefficient of friction was 31.9% lower

than that of pure lubricating oil. The corresponding wear mechanism was changed from

oil-film lubrication to reaction film lubrication, which is mainly attributed to the dense

filling effect of the nanostructured CeO2 and the lubrication effect of rGO. The

nanocomposites were also added into chromium-free Dacromet coating and the

corrosion current density in the corrosion test was decreased by an order of magnitude.
This is mainly due to the formation of various bonds and passivation layers in the

modified coating, which improves its density. It is also attributed to the uniformly

dispersed rGO, which forms a conductive channel, resulting in the preferred corrosion

of the Zn-Al alloy in the coating thus achieving an effective protection effect to the

substrate. When the composites were applied as the photocatalysts in an RhB solution

under the irradiation of ultraviolet light, the RhB dye was effectively removed with an

efficiency up to 90.5%, which is mainly because the rGO can extend the life time of

active electron-hole pairs and prevent CeO2 nanoparticles from agglomeration.

Keywords: Graphene; Rare earth metals; Lubrication; Corrosion resistance;

Photocatalysis

1. Introduction
Graphene is considered as a promising nanomaterial with its unique two-dimensional
layered structure, low shear strength, good chemical stability and carbon-only
composition [1]. Graphene and its derivatives, such as graphene oxide (GO), reduced
graphene oxide (rGO), graphene nanoribbon and their nanocomposites, have
considerable application potentials in the fields of photocatalysis, lubrication and
corrosion resistance [2, 3]. rGO is commonly obtained by reducing GO by chemical
and thermal reduction methods. The reduction of oxygen atoms makes the rGO more
stable than GO, so that it can provide unique lubrication and wear properties, as ideal
nano-additives [4, 5]. It is also widely used in the fields of metal corrosion protection
due to its high chemical and thermal stabilities and excellent chemical barrier properties
[6, 7]. If rGO is added into a protective coating, the layered graphene will be dispersed
in the coating and effectively delay the penetration of corrosive media, form a good
physical barrier to corrosive media and improve the corrosion resistance of the coating
[8]. In addition, attributed to its large specific surface areas, high electron mobility and
a large number of reaction sites, graphene has been used as a perfect substrate for
photocatalytic materials [9, 10].
However, rGO is easily agglomerated due to its large specific surface area, high-
aspect-ratio and strong van der Waals effect, which hinders its wide-range applications
[11]. For this reason, many studies aimed to improve its dispersion in the composites,
using various methods such as molecular level mixing, additions of metal nanoparticles,
and in-situ growth of rGO [12-14]. It is worthwhile to note that doping of various
inorganic metal oxides has also been proposed as a simple but effective method to
improve the physical and chemical properties of graphene-based materials, such as
stability and compatibility [15, 16].
Being one of the most abundant rare earth elements in the earth's crust, cerium is
chemically active and has strong affinities with C, H, O, N, and other non-metallic
elements. It is often used as the additives in tribology, corrosion resistance, catalysis
and other fields [17-19]. For examples, Chen et al. [20] prepared stearic acid-capped
cerium borate composite nanoparticles using a hydrothermal method, which was used
as an additive in rape-seed oil, showing an excellent dispersing stability and outstanding
tribological properties. Mohamed et al. [21] used bis(2-ethylhexyl) phosphate to
disperse CeO2 in synthetic engine oil, and reported that CeO2 nano-lubricant enhanced
the antifriction and anti-wear characteristics by 12-21% and 62-80%, respectively.
Valentine et al. [22] synthesized Ce based metal organic framework (MOF) using a wet
chemical synthesis, and effectively inhibited the cathodic and anodic corrosion reaction
of X65 steel in CO2 solution. Sujay et al. [23] synthesized CeO2 with rich oxygen
vacancies by ultrasound irradiation and used it as an efficient photocatalyst of CO2.
However, rare earth oxides also face the same problem as graphene of severe
agglomeration in their applications.
Chromium-containing Dacromet coating is widely used as a corrosion inhibitor
for iron-based structural components in automotive, power, and marine industries.
However, there are various issues such as the toxic chromium solutions, poor bonding
strength and corrosion resistance with the substrate, and adding modified graphene
additives into the coatings has become one of the solutions [24]. Rare earth elements
can react with oxygen-containing functional groups of graphene oxide to produce new
functional groups, and reduce the interface energy of graphene oxides, thereby
improving the dispersion of graphene oxides. If the rare earth oxides are strongly
bonded with the rGO nanosheets, it will not only solve the problem of rGO
agglomeration, but also promote their binding abilities with other substances. For
example, Zhang et al. [25] loaded lanthanum oxides onto the surface of rGO using a
reflux reduction method, and prevented the aggregation of graphene nanosheets.
Thangavelu et al. [26] prepared Sm2O3/RGO composites using a sonochemical method
and showed that the addition of samarium can effectively avoid the agglomeration of
graphene. Yu et al. [27] synthesized CeO2/rGO with rich oxygen vacancies, showing its
excellent catalytic activity and oxygen reduction durability.
Although there are a lot of studies on the rare earth elements modified rGO
composites, the structures and properties of rGO are easily changed during its
modification process and the rare earth elements easily become agglomerated. So far,
there are few studies to use the rare earth element modified rGO to improve the anti-
wear performance of lubricating oil and as an additive of chromium-free Dacromet
coating to improve the corrosion resistance. In addition, when rare earth elements
modified rGO is used as a photocatalyst, it still has issues such as agglomeration and
poor stability, resulting in low utilization of light energy.
To solve these issues, in this study, CeO2 modified reduced graphene oxide
nanocomposite was prepared using a one-step hydrothermal method. The
nanocomposites were firstly investigated as the additives in PAO5w-40 lubricating oil
for improving its tribological properties, and also in the chromium-free Dacromet
coating for enhancing anti-corrosion properties of coatings. It has further been applied
as a photocatalyst for degrading the organic dye RhB in water solutions. The
enhancement mechanisms of CeO2 modified reduced graphene oxides for these
applications were systematically studied.
2. Experimental
2.1. Preparation and Characterization of CeO2@rGO nanocomposite
CeO2 modified reduced graphene oxide nanocomposite was prepared using a one-
step hydrothermal method as shown in Fig. 2. Graphite oxide powder (80 mg) was
mixed into 40 mL of deionized water, and the mixture was ultrasonically dispersed for
1 h to obtain a homogeneous graphene oxide (GO) solution (2 mg/mL). Then
Ce(NO3)·6H2O was added and continuously stirred for 1 h. The pH value of the mixture
was adjusted to 10 with the diluted ammonia solution. After continually stirred for 2 h,
the above mixture was then transferred into a 100 ml Teflon-lined and tightly sealed
autoclave and maintained at 180℃ for 12 h. After the autoclave was cooled down to
room temperature, the obtained product was dialyzed into neutrality with the deionized
water and then dried using a vacuum freeze-drying method to obtain CeO2 modified
reduced graphene oxide nanocomposite, namely CeO2@rGO (CeO2 wt.%:rGO
wt.%=3:1).

Fig. 1. Illustration of preparation processes of CeO2@rGO.


X-ray diffraction (XRD) patterns of the nanocomposite was obtained using an X-
ray diffractometer (XRD, Bruker D8 ADVANCE) with Cu kα radiation (40 kV,
λ=0.154 nm) at a scan rate of 2°/min within two theta range from 5 to 90°. Raman
spectra were obtained using a Raman spectrometer (inVia reflex, Renishaw, UK) in the
range from 400 cm-1 to 3000 cm-1 with a resolution of 4 cm-1. Microstructures of the
composite were characterized using a transmission electron microscope (TEM, FEI-
Talos, F200X) equipped with the selected area electron diffraction (SAED).
2.2. Tribology, corrosion resistance and photocatalytic tests
The tribological performance of pure lubricating oil (PAO5w-40, Netherlands) and
lubricating oil mixed with 3% concentration of rGO, CeO2 and CeO2@rGO was
evaluated using a ball and disc friction and wear tester (MRS-10G, China). The applied
load was 10 N, and the rotating speed of the steel disk was 300 r/min. The rotating
radius of the ball on the steel disk was 3 mm, and the duration was 30 min. The material
of the grinding ball used in the experiment was GCr15 bearing steel, with a hardness of
62 HRC and a diameter of 5 mm. The material of the steel disk was quenched and
tempered 45 steel, with a surface roughness of 0.8-1.6 microns. The wear scar and
sliding tracks were observed using the SEM and EDX.
In terms of anti-corrosion coating, 1.8 wt.% sodium molybdate (Na2MoO4), 7.2
wt.% Tween-20, 4.8 wt.% acrylic resin lotion, 8.4 wt.% ethylene glycol, and 24 wt.%
zinc aluminum alloy powder (ZnAl20) were mixed thoroughly, and then 0.18 wt.% of
the CeO2@rGO nanocomposite was added under an ultrasonic agitation. The thickener
and defoamer (<1 wt.%) were added in a sequence and stirred for about 5~8 h to obtain
a Dacromet coating solution added with different contents of CeO2@rGO. The coating
was cold sprayed onto the surface of Q235 steel substrate, and then heated at 85℃ for
15 min, then further sintered/solidified at 280℃ for 25 min. Electrochemical tests were
carried out using a P4000 electrochemical workstation with a three-electrode system. A
platinum electrode and saturated calomel electrode were served as the counter and
reference electrodes, respectively, while the sample was used as the working electrode
with an active working area of 1 cm2. The electrochemical tests were performed after
the open circuit potential (OCP) was stabilized in a 3.5% NaCl solution. The
polarization curves were recorded at a scanning speed of 0.5 mV/s within the range
from -250 mV to +250 mV. To study the electrochemical performances of the coatings
with CeO2@rGO, the EIS tests were conducted using the sample with its working area
of 1 cm2, the frequency ranges from 10-2 Hz to 105 Hz and a sine wave amplitude of 10
mV. During the electrochemical impedance spectroscopy (EIS) tests, the samples were
immersed in a 3.5% NaCl solution for different durations, and the impedance and
corrosion resistance of the coatings were evaluated. The salt spray tests were conducted
continuously for 720 h in a NaCl solution of 50 g/L, pH=6.5-7.2.
CeO2@rGO was directly used as a photocatalyst to degrade rhodamine B (RhB)
in a water solution to explore its photocatalytic performance. The nanocomposite was
dispersed into 15 mg/L RhB aqueous solution with a mass ratio (e.g., CeO2@rGO:RhB)
of 20:3 under continuously magnetic stirring. Meanwhile, the mixture was exposed to
UV irradiation (with a UV source power of 30 W and a wavelength of 254 nm). The
samples were taken out periodically and then centrifuged to study the variations of
absorbance of the RhB using a UV-5100 UV-Vis spectrophotometer with a wavelength
of 554 nm. For comparisons, the degradation experiments using the same conditions
were also performed in a dark environment without any UV light irradiation. In the
recycle experiments, after the photocatalysts were separated from solution, they were
washed with ethanol and DI water before being re-dispersed into the RhB dye solution
for another testing cycle.
3. Results and discussion
3.1. Microstructures of CeO2@rGO nanocomposite
Fig. 2(a) shows XRD patterns of CeO2, CeO2@rGO and GO. It can be seen that
the diffraction peaks at the 2θ angles of 28.6°, 33.1°, 47.6°, 56.4°, 59.2°, 69.5°, 76.8°,
79.1° and 88.5° are related to the planes of (111), (200), (220), (311), (222), (400), (331),
(420) and (422) of CeO2, respectively (JCPDS card no. 43-1002). The (001) diffraction
peak of GO appears at 10.5°, which is corresponding to a d-spacing of 8.4 Å, due to the
presence of oxygen-containing functional groups on the surface of graphene oxide sheet
[28]. Another weak (100) diffraction peak of the GO is at 41.8°. Except for the peaks
corresponding to CeO2, there is no peak of rGO. The reason is attributed to that CeO2
was generated between the rGO layers after hydrothermal process, which destroys the
regular structure of rGO [29]. In addition, it may also be that the crystalline peaks of
CeO2 is relatively high, which covers the weak signal of rGO [30].
Fig. 2. (a) XRD patterns of CeO2, GO and CeO2@rGO; (b) Raman spectroscopy of GO and

CeO2@rGO.

Fig. 2(b) shows Raman spectra of GO and CeO2@rGO nanocomposite. The


typical D and G bands are found at ~1350 cm-1 and ~1588 cm-1 in both the GO and
CeO2@rGO. The D band is ascribed to the sp3-hybridized carbon caused by the
presence of oxygen-containing functional groups on the carbon basal plane and the G
band is associated with the in-plane stretching vibration of sp2-hybridized carbon [31,
32]. The intensity ratio of the D band to G band (ID/IG) reflects the defects density in
the graphitic materials [33]. It is calculated that the ID/IG of CeO2@rGO (1.13) is larger
than that of GO (0.98), indicating that the structure of GO has many defects which are
increased during hydrothermal reduction. In addition, the Raman spectra of
CeO2@rGO exhibit a peak at 462 cm-1, corresponding to the symmetrical stretching
mode of the Ce-O-Ce vibrational unit, assigned as F2g [34] . The peak of CeO2@rGO is
very weak, which is related to the small size of CeO2 nanocrystals [35].
Fig. 3(a) shows the FT-IR spectra of the GO, rGO and CeO2@rGO composite. All
the typical characteristic peaks of rGO are visible, and the peak at 3400 cm -1 is due to
the stretching vibration mode of the O-H on the GO surface. The peak at 1386 cm-1 is
attributed to the in-plane bending vibration mode of the C-OH on the rGO surface. The
peaks at 1720 cm-1 and 1650 cm-1 are due to the stretching vibration modes of the C=O
and C=C on the carboxyl group, respectively, the peaks at 1230 cm-1 and 1053 cm-1 are
due to the stretching vibration modes of the C-O-C [36, 37]. In comparison to that of
rGO, the CeO2@rGO exhibits a new Ce-O peak at 557 cm-1, which implies that the rare
earth oxide CeO2 has been successfully grafted onto the surface of rGO, providing an
effective foundation for its usage as a lubricant additive [38]. Comparing the FT-IR
spectra of GO and rGO, it can be observed that the peak intensity of the C=O bond
stretching vibration mode at around 1720 cm-1 is much stronger in rGO than that in GO.
Furthermore, it can be observed that the peak width of the epoxy group in the GO is
much wider than that in the rGO at the same position, which is likely due to the partial
reduction of the product during the reaction process at various reaction temperature and
reagents.
Fig. 3(b) shows the XPS survey spectrum of the CeO2@rGO composite. The major
peaks in the survey spectrum are attributed to C 1s, O 1s, Ce 3d and Ce 4d, indicating
that the main components of the composite are carbon, oxygen and cerium. Fig. 3(c)
shows the high-resolution C 1s spectrum of CeO2@rGO, which exhibits a typical C 1s
peak for the rGO. The curve fitting results reveal three characteristic peaks at 284.8 eV,
285.9 eV, and 287.8 eV, corresponding to C-C, C-O, and C=O bonds, respectively [39].
Fig. 3(d) shows the high-resolution Ce 3d spectrum of the CeO2@rGO composite,
which exhibits a typical Ce 3d peak consisting of a single peak and two multiplet peaks,
indicating the typical Ce 3d peak of CeO2. The Ce 3d peak of CeO2 can be decomposed
into ten peaks through Gaussian-Lorentz function fitting, corresponding to v0 (880.8
eV), v (882.8 eV), v' (885.2 eV), v'' (888.5 eV), v''' (898.6 eV), u0 (899.9 eV), u (901.2
eV), u' (903.6 eV), u'' (906.5 eV), and u''' (916.9 eV). Based on literature, those peaks
of v0, v', u0, and u' are corresponding to 3d5/2 and 3d3/2, which are the signature states of
Ce(III) 3d. Whereas those peaks of v, v'', v''', u, u'', and u''' are corresponding to 3d5/2
and 3d3/2, which are the signature states of Ce(IV) 3d [40]. These results clearly show
that the CeO2 of the composite is mainly in the form of Ce(IV), but there is also a certain
amount of Ce(III). This may be due to the reduction of coordination numbers of Ce
caused by oxygen vacancies in the CeO2 lattice, or it is because Ce3+ is not fully
oxidized during thermal treatments due to insufficient O sources on the rGO [41].
Fig. 3. (a) FT-IR spectra of GO, rGO and CeO2@rGO; (b) XPS survey spectra of

CeO2@rGO; (c) high resolution C1s spectrum of CeO2@rGO; (d) high resolution Ce 3d spectrum

of CeO2@rGO.

Fig. 4 shows the TEM images of rGO and CeO2@rGO. It can be observed from
Fig. 4(a1) that the rGO nanosheets is transparent and has a layered structure. Its surface
is relatively flat, although with wrinkles. From the HRTEM images shown in Fig. 4(a2),
it can be seen that the thickness of rGO layered stacking is 0.85 nm to 1.7 nm. The
attached figure shows that the thickness of rGO monolayer is 0.335 nm, and about 5~7
atomic layers can be clearly observed. As shown in Fig. 4(a3), the selected area electron
diffraction (SAED) pattern of rGO is a typical hexagonal diffraction pattern of graphene
[42]. From the TEM images (see Fig. 4(b1)), the rGO nanosheets are smooth, the
wrinkles on the translucent sheet disappear, whereas the CeO2 nanoparticles are
uniformly distributed. From the HRTEM images shown in Fig. 4(b2), the average
particle size of CeO2 nanoparticles is about 12.24 nm, and the number of rGO
nanosheets is 1~2, which indicates the effective improved dispersion of rGO after
cerium functionalization. The thickness of rGO nanosheets after functionalization is
reduced, which makes it easy for the rGO to be filled into the coating as nano-
reinforcement phase. This will form a dense and intercalated structure, which can
effectively block the channel for corrosion medium to immerse into the substrate. The
lattice fringes of CeO2 nanoparticles can be observed from the HRTEM images. Its fast
Fourier transform (FFT) image is linked with the (200) crystal plane of cubic CeO2,
indicating that CeO2 with a good crystallinity is in-situ grown on the rGO. Fig. 4(b3)
shows the SAED image of CeO2. A series of polycrystalline diffraction spots in the
diffraction pattern correspond to the crystal plane spacings of cubic CeO2 (JCPDF card
no.34-0394).

Fig. 4. (a1-a3) and (b1-b3) are TEM and SAED images of rGO and CeO2@rGO, respectively.

According to the characteristic peaks in the XRD spectrum of CeO2@rGO, the


Debye-Scherrer equation (Eq. 1) was used to estimate the crystallite size of CeO2 [43].
The average crystallite size was calculated to be ~11.02 nm, which is very close to the
result obtained by TEM.

D= βcosθ (1)

where D (nm) is crystallites size. K is Scherrer constant. λ (nm) is wavelength of the X-


ray sources. β (radians) is the value of full width at the half maximum (FWHM). θ
(radians) is diffraction angle positions.
3.2. Enhanced tribological performance of lubricating oil by CeO2@rGO
Fig. 5(a) shows the variation curves of the friction coefficient of lubricating oil
added with different additives tested under the same conditions. It can be seen that the
average values of friction coefficients of pure lubricating oil and lubricating oil added
with rGO, CeO2@rGO and CeO2 are 0.141, 0.107, 0.096 and 0.168, respectively. The
friction coefficient of lubricating oil added with CeO2@rGO is ~31.9% lower than that
of pure lubricating oil. Results clearly shows that CeO2@rGO nanocomposite as an
additive can effectively improve the lubricating ability. On the one hand, due to the
two-dimensional layered structure of the rGO, the atoms in the layers are bonded
through covalent bonds, which can improve the stability of the lubrication. At the same
time, the rGO layers are bonded by van der Waals force, and they are easily slide under
the action of shearing force. In addition, 2D structured rGO has a higher surface area
than that of the traditional materials. When they are existed in the friction interface,
they can cover a large area of the worn surfaces, thus reducing the probability of direct
contact of two counterfaces and decreasing the friction. On the other hand, CeO2, as a
rare earth oxide, has the characteristics of large electronegativity and high activity due
to its special 4f electronic configuration. It can not only clean the surface of graphene,
but also form C-O-Ce or mixed hybrid to make a stable rGO. At the same time, a friction
film was mainly composed of chemical reacted films of Fe2O3, which was formed
during the friction process to avoid direct contact of the friction pair, thus improving
the friction reduction and lubrication performance [44]. Therefore, CeO2 will not only
protect the structure of rGO, but also reduce the surface energy of rGO and improve its
uniform dispersion in the substrate, thus improving the lubricating ability of the pure
lubricating oil.
To confirm the effect of CeO2 modified reduced graphene oxide nanocomposites
on the contact surface, Raman spectra at any two positions on the surface of 45 steel
were obtained after the friction tests of lubricating oil containing CeO2@rGO, and the
obtained results are shown in Fig. 5(b). It can be observed that the Ce-O-Ce vibrational
mode (F2g) shows a typical band at 535.2 cm-1. Compared with the Raman spectrum of
CeO2@rGO nanocomposite, the F2g peak of Ce-O-Ce is shifted to the right-hand side
by 70.2 cm-1, which is due to the deflection of its diffraction peaks under the external
load, consistent with that reported in literature [45]. At the same time, the A1g peak of
Fe3O4 is appeared at 666.8 cm-1, which is because the Fe element in 45 steel reacts with
oxygen in the air during the friction experiment [46]. The typical D and G bands of
CeO2@rGO are appeared at ~1350 cm-1 and ~1590 cm-1, and the 2D band are appeared
at ~2700 cm-1, indicating that rGO effectively fills the pits on the surface of the friction
pairs during the friction process and forms a good lubricating film. However, results
also show that the peak intensities of D and G bands after the wear tests are significantly
lower than those of CeO2@rGO nanocomposite before the friction tests, and the
intensity ratio of D and G bands are also lower than that of the original CeO2@rGO
nanocomposite. These results indicate that the structure of CeO2@rGO nanocomposite
in the worn surface is different from that of the original nanocomposite before the wear
tests. This is because CeO2@rGO nanocomposite was damaged under the action of
friction and shear force after entering the friction contact area, resulting in more defects
and structural changes.
Fig. 5. (a) The variation curves of the friction coefficient of lubricating oil with different additives;

(b) Raman spectra of the wear scar on the surface of 45 steel after friction experiments with

lubricating oil containing CeO2@rGO; (c-e) SEM and EDX images of the wear scar on 45 steel

surface after friction experiments with different additives: (c-c1) pure lubricating oil; (d-d1)

lubricating oil containing rGO; (e-e1) lubricating oil containing CeO2@rGO.

Figs. 5(c-e) show the SEM images and EDX results of the worn track on the
surface of 45 steel after the friction tests using pure lubricating oil, lubricating oil
containing rGO and CeO2@rGO. Fig. 5(c-c1) shows that when the pure lubricating oil
was used as the lubricating medium, the worn surface of 45 steel shows deep wear
marks and gullies. The number of formed furrows is obvious, and the groove marks are
quite wide. The composition analysis of the worn surface shows that it only contains
Fe elements, indicating that no other impurity elements are introduced during the wear
tests. Fig. 5(d-d1) shows that when the lubricating oil containing the rGO nanoparticles
is used as the lubricating medium, the gullies and pear grooves on the surface of 45
steel are much shallow, although there is more abrasive wear observed. The
composition analysis of the wear mark area shows that only C and Fe elements are
detected, which indicates that rGO nanoparticles are covered on the surface of the
friction pair during the friction process.
Results showed that adding a certain amount of rGO nanoparticles improved the
lubricating performance of the lubricating oil. However, due to the poor dispersion
stability of rGO in the lubricating oil, agglomeration phenomenon was observed which
greatly affects the lubricating effect of rGO as a lubricating additive. However, after
the friction test of the lubricating oil containing CeO2@rGO, the furrow phenomenon
in the wear mark area on the surface of 45 steel became insignificant, and the surface
was relatively smooth without any protrusions and depressions (see Fig. 5(e-e1)). These
results show that the friction phenomenon in the wear mark area on the surface of 45
steel is gradually weakened with the increase of Ce3+ concentration, and the grooves
and gullies on the surface are gradually become much shallower. The composition
analysis of the wear mark area shows that the worn surface only contains C, O, Ce, and
Fe elements, which indicates that CeO2@rGO nanocomposites are successfully
deposited on the surface of the friction pair during the friction experiment. This is
consistent with the analysis in Figs. 5(a) and 5(b). To sum up, the wear mechanism is
changed from oil-film lubrication to reaction film lubrication after adding the
nanocomposite into the lubricating oil.
3.3. Enhanced corrosion resistance of coating by CeO2@rGO
Fig. 6 shows SEM images and EDX results of the coatings with 0%, 0.18%, and
0.27% CeO2@rGO after the 720 h salt spray test. It can be seen that the pore contents
on the surface of the coatings with 0.18%, 0.09%, 0%, and 0.27% CeO 2@rGO are
gradually increased. From the enlarged images, the morphology of the coating surface
is gradually changed from a metallic powder shape to a sponge-like structure, and
finally to a typical needle-shaped structure of the corrosion product. Among them, the
surface of the coating with 0.18% CeO2@rGO is still very dense, and the morphology
of the metallic powder can still be observed. Whereas for the coatings with 0% and
0.27% CeO2@rGO, the typical corrosion products of the coatings, i.e., zinc aluminum
chloride hydrates, have an obvious needle-shaped structure. EDX analysis of the
surface of each coating shows that the surface of the coating with 0.18% CeO2@rGO
is mainly composed of Zn, Al, and O elements. Compared with the uncorroded coating,
the amount of O has been increased, indicating that the metallic powder has generated
zinc aluminum oxides after the corrosion test. However, the corroded coating is still
effective for shielding. For the coatings with 0% and 0.27% CeO2@rGO, Fe elements
were detected, and there was more oxygen detected, indicating that more corrosion
products and Fe oxides were generated on the surface of the coating. This can also be
clearly seen from the image where the red rust was produced on the surface of the
substrate. The reason why the EDX results of the coating with 0.18% CeO2@rGO did
not show Fe elements is that this coating has the best corrosion resistance. It can also
be seen that there was no red rust produced on the surface of the substrate coated with
this composite layer.
Fig. 6. SEM and EDX images of surface of chromium-free Dacroment coating reinforced with

different contents of CeO2@rGO before and after corrosion: (a-a4) 0% CeO2@rGO; (b-b4) 0.09%

CeO2@rGO; (c-c4) 0.18% CeO2@rGO; (d-d4) 0.27% CeO2@rGO.

Fig. 7(a) shows the Tafel curves of the coatings with different contents of
CeO2@rGO. Except for the coating with 0.27% CeO2@rGO, the corrosion potentials
of all the other coatings are lower than that of the substrate. The potential difference is
between 0.0504 V to 0.1092 V. All these results indicate that the coatings added with
CeO2@rGO can provide a certain degree of cathodic protection for the substrate [47].
Comparing the coatings added with different contents of CeO2@rGO, the corrosion
potential of the coating with 0.18% CeO2@rGO shows the most negative value, and the
corrosion potential of the coating with 0.27% CeO2@rGO shows the most positive
value. In addition, the corrosion current densities of the coatings with 0.09% and 0.18%
CeO2@rGO are decreased by an order of magnitude if compared to that of the coating
without adding the CeO2@rGO. The corrosion resistance was significantly increased
after adding the CeO2@rGO, indicating that a certain amount of CeO2@rGO can reduce
the Icorr value of the coating and improve its corrosion resistance. However, the Icorr
value of the coating with 0.27% CeO2@rGO is increased. This is because after adding
an excessive amount of CeO2@rGO, the rGO could become contacted with each other
and can form an effective corrosion path, thus increasing the conductivity of the coating
and the corrosion current density. It is also possible that adding too much amount of
composite causes severe agglomeration, producing more pores for the corrosion
medium easier to pass through, thus reducing the corrosion resistance of the coating.

Fig. 7. (a) Tafel polarization curves of chromium-free Dacromet coatings reinforced with different

contents of CeO2@rGO; (b) Nyquist diagram of chromium-free Dacromet coatings reinforced

with different contents of CeO2@rGO; Bode diagram of chromium-free Dacromet coatings

reinforced with different contents of CeO2@rGO: (c) phase frequency diagram; (d) amplitude-

frequency diagram.

Fig. 7(b) shows the Nyquist diagram of the coatings with different contents of
CeO2@rGO in a 3.5% NaCl solution. It can be seen that the capacitive arc radii of the
coatings with 0.09% and 0.18% CeO2@rGO are much larger than that of the coating
with 0% CeO2@rGO, and the density of the coating is increased after adding a certain
amount of composite, both of which can effectively improve the corrosion resistance
of the coating [48]. This is because the graphene is uniformly dispersed in the coating.
It can effectively fill up the defects in the coating, which plays a positive role as a
physical barrier to hinder the intrusion of the corrosive medium. When the CeO2@rGO
content is 0.27%, the capacitive arc radius of the coating decreases, indicating that
adding an excessive amount of composite will reduce its dispersion in the coating,
leading to agglomeration, increased defects in the coating, and also decreased corrosion
resistance of the coating.
Figs. 7(c-d) show the Bode diagrams of coatings added with different amounts of
the composite. From the frequency response diagram, the |Z| values of the coatings with
0.09% and 0.18% CeO2@rGO are higher than those of the coatings without adding the
composite. Among them, the coating with 0.18% CeO2@rGO exhibits the maximum
low-frequency impedance value (|Z|f→0) of about 3.68×103 Ω·cm2, indicating its
minimum corrosion rate and relatively good corrosion resistance. When the added
amount of CeO2@rGO reaches 0.27%, the |Z| value of the coating becomes decreased,
and its corrosion resistance is also decreased. This again proves that adding a certain
amount of composite enhances the corrosion resistance of the coating, but adding
excessive amounts of composite will reduce the corrosion resistance of the coating.
This may be due to the van der Waals force which causes the rGO to be agglomerated
locally, thus increasing the pores in the coating and decreasing its shielding effect.
Fig. 8 shows the equivalent circuit model of the coating’s electrochemical
impedance spectrum, which is Rs(Cc(Rc(QfRf)))(CpRp). Here, Rs, Rc, Rf, and Rp are the
resistances of the solution, coating, surface film layer of zinc-aluminum metal powder,
and surface activation dissolution reaction of zinc-aluminum metal powder,
respectively. Cc, Qf, and Cp are the capacitances of the coating, surface film layer of
zinc-aluminum alloy powder, and surface activation dissolution reaction of zinc-
aluminum metal powder, respectively. Table 1 summarizes the fitting results of the
electrochemical impedance spectra of all the coatings. It can be seen that the Rc values
increase with the increase of the amount of composites. The Rc values of coatings with
0%, 0.09%, and 0.18% CeO2@rGO are relatively small, around 25 Ω·cm2, indicating
that the metallic powder is not completely covered by the passivation film, and cannot
provide effective cathodic protection to the substrate. In Table 1, Rf represents the
resistance of the corrosion product film formed on the surface of the metallic powder.
The larger the value of Rf, the denser the corrosion product layer formed on the surface
of the metallic powder, and the better its protection effect. This will make it more
difficult for the metallic powder to undergo activation dissolution, thus enhancing the
anti-corrosion resistance of the coating. In addition, the value of Rf increases firstly but
then decreases with the further increase of CeO2@rGO. Among them, the coating with
0.18% CeO2@rGO exhibits the largest Rf value of 4064 Ω·cm2, indicating that the
corrosion product film formed by this coating is the densest, thus effectively forming a
physical barrier and delaying the corrosion of the metallic powder. Rp represents the
resistance of the activation dissolution reaction of the metallic powder. The larger the
value of Rp, the more difficult the reaction of the zinc-aluminum powder can occur, and
the better the corrosion resistance of the coating. According to the Table 1, Rp value is
increased with increasing the amounts of CeO2@rGO, and the coating added with 0.27%
CeO2@rGO exhibits the largest Rp. Since both the Rf and Rp values are critical to the
corrosion resistance of the coating, the sum of Rf and Rp is used in this study to represent
the corrosion resistance of the coating, with a larger Rf+Rp indicating a better corrosion
resistance. As shown in the Table 1, the coating with 0.18% CeO2@rGO exhibits the
largest Rf+Rp value, indicating its best corrosion resistance.

Fig. 8. Equivalent circuit diagram.

Table 1. Fitting parameters of electrochemical impedance spectrum for chromium-free Dacormet

coatings reinforced with different contents of CeO2@rGO


wt.% Rs Cc Rc CPEf n Rf Cp Rp
Ω·cm2 F/cm2 Ω·cm2 S-n/cm2 Ω·cm2 F/cm2 Ω·cm2
0% 14.92 7.353×10-7 24.47 3.203×10-4 0.7175 2712 4.212×10-8 64.16
0.09% 13.44 6.706×10-7 25.47 2.784×10-4 0.6558 3985 3.994×10-8 66.31
0.18% 11.45 1.321×10-7 26.51 3.145×10-4 0.6434 4064 4.459×10-8 74
0.27% 12.74 6.024×10-8 54 5.357×10-4 0.5357 3081 8.404×10-8 608.1

Fig. 9 shows the FT-IR spectrum of the Q235 steel substrate with the chromium-
free Dacromet coating with the CeO2@rGO. It can be seen that the coating exhibits the
characteristic peaks corresponding to those of O-Ce, C-O-Si and Si-O-Si bonds [49,
50]. The hydrolyzed silane alcohol produced by the silane coupling agent in the coating
can form C-O-Si bond, and at the same time, the silane alcohol itself becomes
dehydrated and condensed to form linear or cyclic Si-O-Si network structures. The
Ce(OH)3 passivation layer is formed at the interface between the substrate and coating,
which prevents direct contact between the substrate and the external environment [51].
The formation of Si-O-Fe and Ce-Fe bonds at the interface between the coating and
substrate is also beneficial for improving the adhesion between the coating and
substrate, thereby improving its corrosion resistance [32, 52, 53]. Formation of these
bonds can also improve the density of the coating and enhance the bonding strength
between the coating and substrate.

Fig. 9. FT-IR spectrum of the coatings with CeO2@rGO.

3.4. Enhanced photocatalytic performance by CeO2@rGO


Fig. 10(a) shows the photodegradation curves of the organic dye RhB by using
CeO2, GO and CeO2@rGO in the dark and under UV irradiation. It can be seen that the
removal of RhB by GO mainly occurs in the adsorption process under dark conditions,
and the removal rate under the UV irradiation is not decreased significantly, with the
maximum removal rate of 53.2%. CeO2 also mainly removes the RhB through the
adsorption process under dark conditions. However, under the UV irradiation, it
exhibits photocatalytic activity, with the maximum removal rate of 20.3%. CeO2@rGO
achieves a high removal rate of 90.5% for RhB, of which 77.8% is removed through
the adsorption process under dark conditions. The subsequent photocatalytic
degradation process further increases the removal rate up to 90.5%. Fig. 10(b) shows
the UV-vis spectra before and after the RhB adsorption. This shows that the RhB has a
maximum absorption peak at 554 nm before adsorption, but this peak is gradually
weakened after the RhB removal. The reason for this phenomenon may be that the rGO
nanosheets are excellent electron acceptors, while the UV-excited CeO2 is an excellent
electron donor. Therefore, the interactions between CeO2 and rGO nanosheets can
cause excited electrons to be effectively transferred from CeO2 to rGO nanosheets, thus
reducing the possible recombination of electron-hole pairs. Modifications of covalent
bonding between rGO and CeO2 can also improve the dispersion of rGO in the solution.
Furthermore, the layered structure of rGO can prevent the severe aggregation of CeO2
nanoparticles, thereby exposing more active sites to enhance the photocatalytic
performance of the CeO2@rGO composite.
Adsorption and photocatalytic efficiency the CeO2@rGO nanocomposite was
evaluated by analyzing the removal of organic dye RhB in both the dark environment
and under UV irradiation. Fig. 10(c) shows the concentration changes of the RhB dye
solution with time for the CeO2@rGO at different experimental conditions. The red
curve is the degradation curve of CeO2@rGO to the RhB under UV light conditions,
and the black one is the curve of adsorption-desorption equilibrium of CeO2@rGO to
the RhB dye in the dark conditions. After 15 min testing, the concentration of RhB
shows a plateau (see the black line in Fig. 10(c)), indicating that the adsorption-
desorption equilibrium between the nanocomposite and dye molecules is reached after
15 min. It can be seen that ~88% of the RhB dye is absorbed at the equilibrium
adsorption state under dark reaction conditions, indicating that the CeO2@rGO has
shown as good physical adsorption ability. The rGO has a large specific surface area,
and CeO2 nanoparticles are uniformly dispersed on the surface of rGO, exposing
numerous active sites, so that the CeO2@rGO composite has significantly improved the
adsorption activity for an organic dye (RhB). The adsorption kinetics can be described
using Eq. 2 [53]:
ln(C/C0) = -kt (2)
in which the ratio of C (e.g., the concentration of RhB at an adsorption time of t) to C0
(e.g., the initial concentration of RhB) can be given from the relative intensity ratio of
the respective absorbance (A/A0). k is the apparent first-order rate constant (min-1) and
t is adsorption time. Fig. 10(d) shows the ln(C/C0) versus time and a good linear
relationship is observed with the correlation coefficients (R2) value of 0.991, indicating
that the adsorption process follows the pseudo-first-order kinetics. Calculated from the
slope of the straight line, the kinetic adsorption rate constants k is 0.807 min-1.

Fig. 10. (a) Photodegradation curves of different concentrations for RhB by the CeO2, GO and

CeO2@rGO under UV irradiation; (b) UV-vis spectra of RhB before and after photocatalytic
degradation; (c) Concentration changes of RhB dye solutions with time for CeO2@rGO in the

dark and under UV irradiation; (d) The pseudo-first-order reaction kinetics of the CeO2@rGO

composite for RhB degradation; (e) Comparisons of the removal efficiency for RhB dye of

different CeO2 based catalyst; (f) Recycling performance of the CeO2@rGO during adsorption-

photodegradation of the RhB solution.

Under UV irradiation conditions, it is observed that the concentration of RhB is


decreased rapidly in the first 15 min due to the superior adsorption ability of
CeO2@rGO. Compared with those in the dark conditions, the RhB concentration is
gradually decreased from 15 min to 120 min, which is attributed to the photo-
degradation of RhB enhanced by CeO2. In the initial stage, CeO2@rGO rapidly adsorbs
the RhB dye molecules due to the action of electrostatic attraction. As the adsorption
process is progressed further, the electrostatic attraction between CeO2@rGO and RhB
molecules becomes weaker, eventually reaching an adsorption-desorption equilibrium.
When exposed to the UV light, the CeO2 nanostructures on the rGO generate electrons
and holes. The rGO in the nanocomposite acts as both acceptor and transporter for the
electrons due to its good conductivity. Therefore, the photogenerated electrons from the
CeO2 is effectively transferred to the rGO nanosheets [32]. The photogenerated holes
are transferred to the surface of CeO2, and after the reaction, they will degrade the
adsorbed RhB dye molecules, so that the concentration of RhB solution will continue
to decrease. Results show that about 90.5% of RhB is degraded by CeO2@rGO after
120 min. As noted in Fig. 10(e), the removal efficiency for the RhB dye of our work is
significantly higher than those of the other CeO2 based catalyst reported in literature
[54-59]. Fig. 10(f) shows the cycling performance of the nanocomposite for the RhB
photocatalytic degradation. The nanocomposite shows only slight decrease in the
photocatalytic activity after three cycles, which is mainly due to an incomplete removal
of the RhB decomposition products during the regeneration process and the
photoetching induced by accumulation of photogenerated holes on the CeO2 surface
[60].
4. Mechanisms of CeO2@rGO in lubrication, anti-corrosion and photocatalysis
Based on the characteristics for the integrated composite of rGO and CeO2, we
proposed the mechanisms of using CeO2@rGO nanocomposites for improvement of
tribological properties, enhancement of corrosion resistance of coating, and
photocatalytic performance for degradation of dyes, as shown in Fig. 11.

Fig. 11. Schematic illustrations of mechanisms using CeO2@rGO nanocomposite for lubrication,

anti-corrosion and photocatalysis applications. (a) Lubrication mechanisms of different types of

lubricating oil (in which (a1) is for pure lubricating oil; lubricating oil with (a2) CeO2; (a3) rGO;

(a4) CeO2@rGO); (b) Anti-corrosion mechanism of chromium-free Dacromet coating added with

CeO2@rGO ((b1) before adding CeO2, (b2) after adding, (b3) interfacial structure); (c)

Photocatalytic mechanism.
When the lubricating oil is used as a lubricating medium, it will become "extruded"
under the action of large contact stress. This will cause the direct contact between the
steel ball and substrate, resulting in increased friction and wear [61], as shown in Fig.
11(a1). When CeO2 is added inside the lubricating oil, the three-dimensional structure
of CeO2 nanoparticles can effectively fill the pits on the friction pair surface and
smoothen the surface. However, the surface energy of CeO2 nanoparticles is still quite
large, and thus they are easily agglomerated in the lubricating oil, which will reduce the
lubricating performance of the lubricating oil [62], as shown in Fig. 11(a2). When
graphene nanosheets are used as the additives, the layered structure can fill the pits of
friction pairs, and graphene nanosheets are easily broken under the action of shearing
force, thus forming a good lubricating film to prevent the direct contact of friction pairs
[63], as shown in Fig. 11(a3). Similarly, graphene itself is easily agglomerated due to
its large surface energy. Using the CeO2 modified reduced graphene oxide
nanocomposites (CeO2@rGO) can solve this issue as a good lubricating oil additive, as
shown in Fig. 11(a4). Rare earth elements such as Ce easily interact with the oxygen-
containing functional groups on the surface of rGO to form O-Ce. Grafting CeO2 onto
the surface of rGO can effectively reduce the surface energy of rGO, thus dispersing
the CeO2 uniformly in the lubricating oil and avoiding the severe agglomeration. In
addition, CeO2@rGO nanocomposites have a dual lubrication mechanism from both
the CeO2 nanoparticles and rGO nanosheets. In this case, the surface of friction pairs
can be worn smoothly by filling the vacancies with a polishing effect, and a good
lubricating film can be formed between friction pairs to prevent any direct contact of
friction pairs. The PAO5w-40 lubricating oil containing CeO2@rGO can not only fill
the rough physical interface, but also form a lubricating film, greatly reducing the
friction coefficient. Accordingly, the friction and wear resistance of PAO5w-40
lubricating oil with the nanocomposites as additive has been significantly improved.
Figs. 11(b1) to 11(b3) show the mechanisms of corrosion resistance of the
CeO2@rGO composites which are dispersed uniformly in the coating. In this study of
the chromium-free Dacromet coating, we used the molybdic acid instead of the
traditionally used chromic acid. The corrosion resistance mechanism of the coating has
also been changed to the formation of insoluble molybdate on the substrate as shown
in Eq. 3. The formed oxide film on the substrate’s surface passivates it from corrosion,
which competes with Cl- during corrosion, thus reducing the corrosion of Cl-. In
addition, the corrosion resistance performance of the coating is further improved after
the addition of CeO2@rGO. The modified rGO is uniformly dispersed in the coating,
and its layered structure easily adheres to the zinc-aluminum alloy powder and plays a
key role in the solidification of the CeO2@rGO. On the other hand, the nanoflakes and
nanoparticles fill the porous structure of the coating, greatly increasing its density and
hardness. The homogeneous distribution of the CeO2@rGO in the coating promotes the
formation of a conductive channel, facilitates the electron transfer, and is thus beneficial
for sacrificing zinc-aluminum alloy with a lower potential for protecting the substrate.
The hydrolysis of the silane coupling agent in the coating produces a silanol that can
form C-O-Si, and Si-O-Fe is also formed at the interface between the coating and the
substrate [49]. At the same time, the self-dehydration and condensation of silanol form
a linear or cyclic Si-O-Si network structure. The Ce(OH)3 passivation layer is formed
at the interface between the substrate and the coating, which actively prevents direct
contact between the substrate and the external environment. This indicates that the
compound in the anti-corrosion film has barrier/active dual-functions. The presence of
Ce-Fe bonds can also improve the adhesion between the coating and substrate, thereby
improving the wear resistance and corrosion resistance of the coating [52]. Therefore,
the corrosion resistance of the chromium-free Dacromet coating with the added
CeO2@rGO has been improved.
MoO42- + Fe2+ → FeMoO4 → Fe2(MoO4)3 (3)
Fig. 11(c) shows the photocatalytic degradation mechanism of the composite for
RhB solution. The rGO acts as both acceptor and transporter for the photo-generated
electrons due to its good conductivity, thus prolonging the life of electron/hole pairs
[32]. The interactions between CeO2 and rGO can lead to the effective transfer of
excited electrons from CeO2 to the surface of rGO nanosheets, thus reducing the
electron-hole pairs recombination [54]. The 2D π-conjugation system of rGO facilitates
the easy movement of electrons towards the conduction band to participate in
degradation reaction (generating superoxide radical which in turn produces hydroxyl
radicals) rather than recombination with holes. In addition, the uniform loading of rare
earth oxides on the rGO surface without significant agglomeration helps to expose more
active sites. Due to these reasons, the CeO2@rGO nanocomposites will have good
properties for photocatalysis.
5. Conclusions
In this paper, CeO2 modified reduced graphene oxide nanocomposite (CeO2@rGO)
was successfully prepared with cerium nitrate hexahydrate and graphene nanosheets as
the raw materials using a hydrothermal method. When the nanocomposite was added
into PAO5w-40 lubricating oil, the friction coefficient was 31.9% lower than that of
pure lubricating oil. The wear mechanism was changed from oil-film lubrication to
reaction film lubrication. The corrosion current density of the chromium-free Dacromet
coating added with the nanocomposite was decreased by an order of magnitude
compared with that of the coating without adding the nanocomposite. This is mainly
due to the addition of the composites will increase the density of the coating and form
a conductive channel which easily corrode the alloy powder at a much lower potential.
When it was used as the photocatalyst, the removal rate of RhB dye can reach 90.5%,
because rGO can reduce the recombination of electron-hole pairs and prevent CeO2
from agglomeration, thereby exposing more active sites and extending the life of
electron-hole pairs.
Acknowledgement
The authors would like to acknowledge the financial supports from the Science and
Technology Project of Shaanxi Province of China (No.2020ZDLGY12-06), Xi’an
Science research project of China (No.2021XJZZ0042) and International Exchange
Grant (IEC/NSFC/201078) through Royal Society and National Science Foundation of
China (NSFC).
Conflicts of interest
We declare that we do not have any commercial or associative interest that
represents a conflict of interest in connection with the work submitted.
References
[1] Lee C, Li Q, Kalb W, Liu X, Berger H, Carpick R and Hone J 2010 Frictional
Characteristics of Atomically Thin Sheets Science 328 76
[2] Kim K, Lee H, Lee C, Lee S, Jang J, Ahn J, Kim J and Lee H 2011 Chemical Vapor
Deposition-Grown Graphene: The Thinnest Solid Lubricant ACS Nano 5 5107
[3] Berman D, Erdemir A and Sumant A 2014 Graphene: a new emerging lubricant
Materials Today 17 31.
[4] Xu H, Zang J, Yuan Y, Tian P and Wang Y 2019 Preparation of multilayer graphene
coatings with interfacial bond to mild steel via covalent bonding for high performance
anticorrosion and wear resistance Carbon 154 156.
[5] Cui G, Bi Z, Zhang R, Liu J, Yu X and Li Z 2019 A comprehensive review on
graphene-based anti-corrosive coatings Chemical Engineering Journal 373 104.
[6] Ji X, Lin Y, Zeng J, Ren Z, Lin Z, Mu Y, Qiu Y and Yu J 2021
Graphene/MoS2/FeCoNi(OH)x and Graphene/MoS2/FeCoNiPx multilayer-stacked
vertical nanosheets on carbon fibers for highly efficient overall water splitting Nature
Communication 12 1380.
[7] Dong C, Lu J, Qiu B, Shen B, Xing M and Zhang J 2018 Developing stretchable
and graphene-oxide-based hydrogel for the removal of organic pollutants and metal
ions Applied Catalysis B: Environmental 222 146.
[8] Wang J, Jin X, Li C, Wang W, Wu H and Guo S 2019 Graphene and graphene
derivatives toughening polymers: Toward high toughness and strength Chemical
Engineering Journal 370 831.
[9] Hu X, Deng F, Huang W, Zeng G, Luo X and Dionysiou D 2018 The band structure
control of visible-light-driven rGO/ZnS-MoS2 for excellent photocatalytic degradation
performance and long-term stability Chemical Engineering Journal 350 248.
[10] Ho P, Lee W, Liou Y, Chiu Y, Shih Y, Chen C, Su P, Li M, Chen H, Liang C and
Chen C 2015 Sunlight-activated graphene-heterostructure transparent cathodes:
enabling high-performance n-graphene/p-Si Schottky junction photovoltaics Energy &
Environmental Science 8 2085.
[11] Lee B, Koo M, Jin S, Kim K and Hong S 2014 Simultaneous strengthening and
toughening of reduced graphene oxide/alumina composites fabricated by molecular-
level mixing process Carbon 78 212.
[12] Peng Y, Lu B and Chen S 2018 Carbon-Supported Single Atom Catalysts for
Electrochemical Energy Conversion and Storage Advanced Materials 30 1801995.
[13] Feng Y, Zhang S, Zhu L, Li G, Zhao N, Zhang H andChen B 2022 Reduced
graphene oxide-supported ruthenium nanocatalysts for highly efficient electrocatalytic
hydrogen evolution reaction International Journal of Hydrogen Energy 47 39853.
[14] Yang Z, Liu C, Liu X, Du Y, Jin H, Ding F, Li B, Ouyang Y, Bai L and Yuan F
2022 In-situ synthesis of graphene nanosheets encapsulated silicon nanospheres by
thermal plasma for ultra-stable lithium storage Carbon 199 424.
[15] Gong C, Hinojos D, Wang W, Nijem N, Shan B, Wallace R, Cho K and Chabal Y
2012 Metal-Graphene-Metal Sandwich Contacts for Enhanced Interface Bonding and
Work Function Control ACS Nano 6 5381.
[16] Sun Y, Liu Q, Gao S, Cheng H, Lei F, Sun Z, Jiang Y, Su H, Wei S and Xie Y 2013
Pits confined in ultrathin cerium(IV) oxide for studying catalytic centers in carbon
monoxide oxidation Nature Communication 4 2899.
[17] Zhang Z, Wang Z, Liang B and La P 2006 Effects of CeO2 on friction and wear
characteristics of Fe–Ni–Cr alloy coatings Tribology International 39 971.
[18] Pereira G, Prada O, Avila P, Avila J, Pinto H, Miyazaki M, de Melo H and Bose W
2022 Cerium conversion coating and sol-gel coating for corrosion protection of the
WE43 Mg alloy Corrosion Science 206 110527.
[19] Liang X, Wang P, Gao Y, Huang H, Tong F, Zhang Q, Wang Z, Liu Y, Zheng Z,
Dai Y and Huang B 2020 Design and synthesis of porous M-ZnO/CeO2 microspheres
as efficient plasmonic photocatalysts for nonpolar gaseous molecules oxidation: Insight
into the role of oxygen vacancy defects and M=Ag, Au nanoparticles Applied Catalysis
B: Environmental 260 118151.
[20] Boshui C, Kecheng G, Jianhua F, Jiang W, Jiu W and Nan Z 2015 Tribological
characteristics of monodispersed cerium borate nanospheres in biodegradable rapeseed
oil lubricant Applied Surface Science 353 326.
[21] Ali M and Xianjun H 2022 Exploring the lubrication mechanism of CeO2
nanoparticles dispersed in engine oil by bis(2-ethylhexyl) phosphate as a novel antiwear
additive Tribology International 165 107321.
[22] Anadebe V, Chukwuike V, Ramanathan S and Barik R 2022 Cerium-based metal
organic framework (Ce-MOF) as corrosion inhibitor for API 5L X65 steel in CO2-
saturated brine solution: XPS, DFT/MD-simulation, and machine learning model
prediction Process Safety and Environmental Protection 168 499.
[23] Sujay S, Alkanad K, Thejaswini B, Alnaggar G, Zaqri N, Drmosh Q, Boshaala A
and Lokanath N 2022 Alkaline mediated sono-synthesis of surface oxygen-vacancies-
rich cerium oxide for efficient photocatalytic CO2 reduction to methanol Surfaces and
Interfaces 34 102389.
[24] Gao Z, Zhang D, Li X, Jiang S and Zhang Q 2018 Current status, opportunities
and challenges in chemical conversion coatings for zinc Colloids and Surfaces A 546
221.
[25] Zhang J, Zhang Z, Jiao Y, Yang H, Li Y, Zhang J and Gao P 2019 The
graphene/lanthanum oxide nanocomposites as electrode materials of supercapacitors
Journal of Power Sources 419 99.
[26] Sakthi T, Nataraj N, Chen T, Chen S and Kokulnathan T 2022 Synergistic
formation of samarium oxide/graphene nanocomposite: A functional electrocatalyst for
carbendazim detection Chemosphere 307 135711.
[27] Yu Y, Wang X, Gao W, Li P, Yan W, Wu S, Cui Q, Song W and Ding K 2017
Trivalent cerium-preponderant CeO2/graphene sandwich-structured nanocomposite
with greatly enhanced catalytic activity for the oxygen reduction reaction Journal of
Materials Chemistry A 5 6656.
[28] Qian X, Jin L, Zhu L, Yao S, Rao D, Shen X, Xi X, Xiao K and Qin S 2016 CeO2
nanodots decorated ketjen black for high performance lithium–sulfur batteries RSC
Advances 6 111190.
[29] Bakry A, Alamier W, Salama R, El-Shall M and Awad F 2022 Remediation of
water containing phosphate using ceria nanoparticles decorated partially reduced
graphene oxide (CeO2-PRGO) composite Surfaces and Interfaces 31 102006.
[30] Min C, He Z, Liu D, Jia W, Qian J, Jin Y and Li S 2019 Ceria/reduced Graphene
Oxide Nanocomposite: Synthesis, Characterization, and Its Lubrication Application
ChemistrySelect 4 4615.
[31] Ferrari A and Basko D 2013 Raman spectroscopy as a versatile tool for studying
the properties of graphene Nature Nanotechnology 8 235.
[32] Ferrari A, Meyer J, Scardaci V, Casiraghi C, Lazzeri M, Mauri F, Piscanec S, Jiang
D, Novoselov K, Roth S and Geim A 2006 Raman spectrum of graphene and graphene
layers Physical Review Letters 97 187401.
[33] Dong L, Chen W, Deng N and Zheng C 2016 A novel fabrication of graphene by
chemical reaction with a green reductant Chemical Engineering Journal 306 754.
[34] Verma R and Samdarshi S 2016 In Situ Decorated Optimized CeO2 on Reduced
Graphene Oxide with Enhanced Adsorptivity and Visible Light Photocatalytic Stability
and Reusability The Journal of Physical Chemistry C 120 22281.
[35] Murugan R, Vijayaprasath G and Ravi G 2015 The influence of substrate
temperature on the optical and micro structural properties of cerium oxide thin films
deposited by RF sputtering Superlattices and Microstructures 85 321.
[36] Deosarkar M, Pawar S and Bhanvase B 2014 In-situ sonochemical synthesis of
Fe3O4-graphene nanocomposite for lithium rechargeable batteries Chemical
Engineering and Processing 83 49.
[37] Gu M, Farooq U, Lu S, Zhang X, Qiu Z and Sui Q 2018 Degradation of
trichloroethylene in aqueous solution by rGO supported nZVI catalyst under several
oxic environments Journal of Hazardous Materials 349 35.
[38] Abbasi A, Amin K, Ali M, Ali Z, Atif M, Ensinger W and Khalid W 2022
Synergetic effect of adsorption-photocatalysis by GO-CeO2 nanocomposites for
photodegradation of doxorubicin Journal of Environmental Chemical Engineering 10
107078.
[39] Stankovich S, Dikin D, Piner R, Kohlhaas K, Kleinhammes A, Jia Y, Wu Y,
Nguyen S and Ruoff R 2007 Synthesis of graphene-based nanosheets via chemical
reduction of exfoliated graphite oxide Carbon 45 1558.
[40] Beche E, Charvin P, Perarnau D, Abanades S and Flamant G 2008 Ce 3d XPS
investigation of cerium oxides and mixed cerium oxide (CexTiyOz) Surface and
Interface Analysis 40 264.
[41] Yuan S, Zhang Q, Xu B, Jin Z, Zhang Y, Yang Y, Zhang M and Ohno T 2014
Porous cerium dioxide hollow spheres and their photocatalytic performance RSC
Advances 4 62255.
[42] Dreyer D, Park S, Bielawski C and Ruoff R 2010 The chemistry of graphene oxide
Chemical Society Reviews 39 228.
[43] Basak M, Rahman M, Ahmed M, Biswas B and Sharmin N 2022 The use of X-ray
diffraction peak profile analysis to determine the structural parameters of cobalt ferrite
nanoparticles using DebyeScherrer, Williamson-Hall, Halder-Wagner and Size-strain
plot: Different precipitating agent approach Journal of Alloys and Compounds 895
162694.
[44] Chen W, Feng Y, Wan Y, Zhang L, Yang D, Gao X, Yu Q and Wang D 2023
Investigation on anti-wear and corrosion-resistance behavior of steel-steel friction pair
enhanced by ionic liquid additives under conductive conditions Tribology International
177 108002.
[45] Kourouklis G, Jayaraman A and Espinosa G 1988 High-pressure Raman study of
CeO2 to 35 GPa and pressure-induced phase transformation from the fluorite structure
Physical Review B: Condens Matter 37 4250.
[46] Mi W, Guo Z, Wang Q, Yang Y and Bai H 2013 Charge ordering in reactive
sputtered (100) and (111) oriented epitaxial Fe3O4 films Scripta Materialia 68 972.
[47] Qian Y, Zheng W, Chen W, Feng T, Liu T and Fu Y 2021 Enhanced functional
properties of CeO2 modified graphene/epoxy nanocomposite coating through interface
engineering Surface and Coatings Technology 409 126819.
[48] Xiang T, Ding S, Li C, Zheng S, Hu W, Wang J and Liu P 2017 Effect of current
density on wettability and corrosion resistance of superhydrophobic nickel coating
deposited on low carbon steel Materials & Design 114 65.
[49] Lotfi M, Yari H, Sari M and Azizi A 2022 Fabrication of a highly hard yet tough
epoxy nanocomposite coating by incorporating graphene oxide nanosheets dually
modified with amino silane coupling agent and hyperbranched polyester-amide
Progress in Organic Coatings 162 106570.
[50] Li L, Chen D, Long Y, Wang F and Kang Z 2023 Silane modification of semi-
curing epoxy surface: High interfacial adhesion for conductive coatings Progress in
Organic Coatings 174 107228.
[51] Cai G, Hu P, Cao X, Chen J, Zhang X and Dong Z 2023 pH-triggered self-
inhibition epoxy coating based on cerium-polyphenolic network wrapped carbon
nanotube Progress in Organic Coatings 175 107355.
[52] Lei Y, Zhang X, Liu Q, Tao Y, Sun Y and You B 2022 Skin-mimetic assembly
strategy for fabricating a transparent and highly anti-corrosive FSO-GO/epoxy
nanocomposite coating Progress in Organic Coatings 173 107184.
[53] Alipanaha N, Yaria H, Mahdaviana M, Ramezanzadeha B and Bahlakeh G 2021
MIL-88A (Fe) filler with duplicate corrosion inhibitive/barrier effect for epoxy coatings:
Electrochemical, molecular simulation, and cathodic delamination studies Journal of
Industrial and Engineering Chemistry 97 200.
[54] Xiao Z, Wu X, Tan H and Hao S 2023 Design and synthesis of Fe-Ce-O@C with
efficient photocatalytic activity Journal of Rare Earths 41 91.
[53] Yin W and Cao H 2016 One-step synthesis of SnO2-reduced graphene oxide (SOG)
composites for efficient removal of organic dyes from wastewater RSC Advances 6
100636.
[54] Choi J, Reddy D, Islam M, Ma R and Kim T 2016 Self-assembly of CeO2
nanostructures/reduced graphene oxide composite aerogels for efficient photocatalytic
degradation of organic pollutants in water Journal of Alloys and Compounds 688 527.
[55] Song H, Wu R, Yang J, Dong J and Ji G 2018 Fabrication of CeO2 nanoparticles
decorated three-dimensional flower-like BiOI composites to build p-n heterojunction
with highly enhanced visible-light photocatalytic performance J Colloid Interface Sci
512 325.
[56] Munawar T, Mukhtar F, Nadeem M, Riaz M, Naveed M, Mahmood K, Hasan M,
Arshad M, Hussain F, Hussain A and Iqbal F 2020 Novel photocatalyst and antibacterial
agent; direct dual Z-scheme ZnO–CeO2-Yb2O3 heterostructured nanocomposite Solid
State Sciences 109 106446.
[57] Maria C, Kaviyarasu K, Judith J, Siddhardha B and Jeyaraj B 2017 Facile synthesis
of heterostructured cerium oxide/yttrium oxide nanocomposite in UV light induced
photocatalytic degradation and catalytic reduction: Synergistic effect of antimicrobial
studies Journal of Photochemistry and Photobiology B: Biology 173 23.
[58] Liyanage A, Perera S, Tan K Chabal Y and Balkus K 2014 Synthesis,
Characterization, and Photocatalytic Activity of Y-Doped CeO2 Nanorods ACS
Catalysis 4 577.
[59] Yao J, Gao Z, Meng Q, He G and Chen H 2021 One-step synthesis of reduced
graphene oxide based ceric dioxide modified with cadmium sulfide (CeO2/CdS/RGO)
heterojunction with enhanced sunlight-driven photocatalytic activity Journal of Colloid
and Interface Science 594 621.
[60] Djellabi R, Ghorab F, Nouacer S, Smara A and Khireddine O 2016 Cr(VI)
photocatalytic reduction under sunlight followed by Cr(III) extraction from TiO2
surface Materials Letters 176 106.
[61] Dai W, Kheireddin B, Gao H and Liang H 2016 Roles of nanoparticles in oil
lubrication Tribology International 102 88.
[62] Shen T, Wang D, Yun J, Liu Q, Liu X and Peng Z 2016 Tribological properties and
tribochemical analysis of nano-cerium oxide and sulfurized isobutene in titanium
complex grease Tribology International 93 332.
[63] Englert J, Dotzer C, Yang G, Schmid M, Papp C, Gottfried J, Steinruck H, Spiecker
E, Hauke F and Hirsch A 2011 Covalent bulk functionalization of graphene Nature
Chemistry 3 279.
Revised Manuscript with tracked changes Click here to view linked References

1 CeO2 modified reduced graphene oxide nanocomposites for


2
3
4 tribology, anti-corrosion and photocatalysis applications
5
6
7 Bo Sun1, Wenge Chen1,*, Hui Zhang1, Tao Feng1, Wanqing Xing1, Ahmed Elmarakbi2,
8
9 Yong-Qing Fu2,*
10
11
12
1
School of Materials Science and Engineering, Xi’an University of Technology, Xi’an,
13
14
15
Shaanxi, 710048, P.R. China
16
17 2
Faculty of Engineering and Environment, Northumbria University, Newcastle upon
18
19
20 Tyne, NE1 8ST, UK.
21
22 *Corresponding authors: Prof. Wenge Chen; Prof. Richard Yongqing Fu.
23
24
25 E-mail: wgchen001@263.net (W.G. Chen), richard.fu@northumbria.ac.uk (Y.Q. Fu)
26
27
28 Abstract
29
30
31 CeO2 modified reduced graphene oxide (CeO2@rGO) nanocomposites
32
33
34
synthesized using a hydrothermal method were studied for their performance and
35
36 mechanisms for tribology, anti-corrosion and photocatalysis applications. Results
37
38
39 showed that ceria nanoparticles with an average diameter of ~12.24 nm were uniformly
40
41
42 distributed and covalently bonded onto the surface of rGO. When the nanocomposites
43
44
45
were added into PAO5w-40 lubricating oil, the coefficient of friction was 31.9% lower
46
47 than that of pure lubricating oil. The corresponding wear mechanism was changed from
48
49
50 oil-film lubrication to reaction film lubrication, which is mainly attributed to the dense
51
52
53 filling effect of the nanostructured CeO2 and the lubrication effect of rGO. The
54
55
56
nanocomposites were also added into chromium-free Dacromet coating and the
57
58 corrosion current density in the corrosion test was decreased by an order of magnitude.
59
60
61
62
63
64
65
1 This is mainly due to the formation of various bonds and passivation layers in the
2
3
4
modified coating, which improves its density. It is also attributed to the uniformly
5
6 dispersed rGO, which forms a conductive channel, resulting in the preferred corrosion
7
8
9 of the Zn-Al alloy in the coating thus achieving an effective protection effect to the
10
11
12 substrate. When the composites were applied as the photocatalysts in an RhB solution
13
14
15
under the irradiation of ultraviolet light, the RhB dye was effectively removed with an
16
17 efficiency up to 90.5%, which is mainly because the rGO can extend the life time of
18
19
20 active electron-hole pairs and prevent CeO2 nanoparticles from agglomeration.
21
22
23
24
25 Keywords: Graphene; Rare earth metals; Lubrication; Corrosion resistance;
26
27 Photocatalysis
28
29
30 1. Introduction
31
32 Graphene is considered as a promising nanomaterial with its unique two-dimensional
33
34 layered structure, low shear strength, good chemical stability and carbon-only
35
36 composition [1]. Graphene and its derivatives, such as graphene oxide (GO), reduced
37
38 graphene oxide (rGO), graphene nanoribbon and their nanocomposites, have
39
40 considerable application potentials in the fields of photocatalysis, lubrication and
41
42
43
corrosion resistance [2, 3]. rGO is commonly obtained by reducing GO by chemical
44
45
and thermal reduction methods. The reduction of oxygen atoms makes the rGO more
46
47 stable than GO, so that it can provide unique lubrication and wear properties, as ideal
48
49 nano-additives [4, 5]. It is also widely used in the fields of metal corrosion protection
50
51 due to its high chemical and thermal stabilities and excellent chemical barrier properties
52
53 [6, 7]. If rGO is added into a protective coating, the layered graphene will be dispersed
54
55 in the coating and effectively delay the penetration of corrosive media, form a good
56
57 physical barrier to corrosive media and improve the corrosion resistance of the coating
58
59 [8]. In addition, attributed to its large specific surface areas, high electron mobility and
60
61
62
63
64
65
a large number of reaction sites, graphene has been used as a perfect substrate for
1
2
3
photocatalytic materials [9, 10].
4
5
However, rGO is easily agglomerated due to its large specific surface area, high-
6
7 aspect-ratio and strong van der Waals effect, which hinders its wide-range applications
8
9 [11]. For this reason, many studies aimed to improve its dispersion in the composites,
10
11 using various methods such as molecular level mixing, additions of metal nanoparticles,
12
13 and in-situ growth of rGO [12-14]. It is worthwhile to note that doping of various
14
15 inorganic metal oxides has also been proposed as a simple but effective method to
16
17 improve the physical and chemical properties of graphene-based materials, such as
18
19 stability and compatibility [15, 16].
20
21 Being one of the most abundant rare earth elements in the earth's crust, cerium is
22
23 chemically active and has strong affinities with C, H, O, N, and other non-metallic
24
25 elements. It is often used as the additives in tribology, corrosion resistance, catalysis
26
27 and other fields [17-19]. For examples, Chen et al. [20] prepared stearic acid-capped
28
29 cerium borate composite nanoparticles using a hydrothermal method, which was used
30
31 as an additive in rape-seed oil, showing an excellent dispersing stability and outstanding
32
33
34
tribological properties. Mohamed et al. [21] used bis(2-ethylhexyl) phosphate to
35
36
disperse CeO2 in synthetic engine oil, and reported that CeO2 nano-lubricant enhanced
37
38 the antifriction and anti-wear characteristics by 12-21% and 62-80%, respectively.
39
40 Valentine et al. [22] synthesized Ce based metal organic framework (MOF) using a wet
41
42 chemical synthesis, and effectively inhibited the cathodic and anodic corrosion reaction
43
44 of X65 steel in CO2 solution. Sujay et al. [23] synthesized CeO2 with rich oxygen
45
46 vacancies by ultrasound irradiation and used it as an efficient photocatalyst of CO2.
47
48 However, rare earth oxides also face the same problem as graphene of severe
49
50 agglomeration in their applications.
51
52 Chromium-containing Dacromet coating is widely used as a corrosion inhibitor
53
54 for iron-based structural components in automotive, power, and marine industries.
55
56 However, there are various issues such as the toxic chromium solutions, poor bonding
57
58 strength and corrosion resistance with the substrate, and adding modified graphene
59
60 additives into the coatings has become one of the solutions [24]. Rare earth elements
61
62
63
64
65
can react with oxygen-containing functional groups of graphene oxide to produce new
1
2
3
functional groups, and reduce the interface energy of graphene oxides, thereby
4
5
improving the dispersion of graphene oxides. If the rare earth oxides are strongly
6
7 bonded with the rGO nanosheets, it will not only solve the problem of rGO
8
9 agglomeration, but also promote their binding abilities with other substances. For
10
11 example, Zhang et al. [25] loaded lanthanum oxides onto the surface of rGO using a
12
13 reflux reduction method, and prevented the aggregation of graphene nanosheets.
14
15 Thangavelu et al. [26] prepared Sm2O3/RGO composites using a sonochemical method
16
17 and showed that the addition of samarium can effectively avoid the agglomeration of
18
19 graphene. Yu et al. [27] synthesized CeO2/rGO with rich oxygen vacancies, showing its
20
21 excellent catalytic activity and oxygen reduction durability.
22
23 Although there are a lot of studies on the rare earth elements modified rGO
24
25 composites, the structures and properties of rGO are easily changed during its
26
27 modification process and the rare earth elements easily become agglomerated. So far,
28
29 there are few studies to use the rare earth element modified rGO to improve the anti-
30
31 wear performance of lubricating oil and as an additive of chromium-free Dacromet
32
33
34
coating to improve the corrosion resistance. In addition, when rare earth elements
35
36
modified rGO is used as a photocatalyst, it still has issues such as agglomeration and
37
38 poor stability, resulting in low utilization of light energy.
39
40 To solve these issues, in this study, CeO2 modified reduced graphene oxide
41
42 nanocomposite was prepared using a one-step hydrothermal method. The
43
44 nanocomposites were firstly investigated as the additives in PAO5w-40 lubricating oil
45
46 for improving its tribological properties, and also in the chromium-free Dacromet
47
48 coating for enhancing anti-corrosion properties of coatings. It has further been applied
49
50 as a photocatalyst for degrading the organic dye RhB in water solutions. The
51
52 enhancement mechanisms of CeO2 modified reduced graphene oxides for these
53
54 applications were systematically studied.
55
56 2. Experimental
57
58 2.1. Preparation and Characterization of CeO2@rGO nanocomposite
59
60 CeO2 modified reduced graphene oxide nanocomposite was prepared using a one-
61
62
63
64
65
step hydrothermal method as shown in Fig. 2. Graphite oxide powder (80 mg) was
1
2
3
mixed into 40 mL of deionized water, and the mixture was ultrasonically dispersed for
4
5
1 h to obtain a homogeneous graphene oxide (GO) solution (2 mg/mL). Then
6
7 Ce(NO3)·6H2O was added and continuously stirred for 1 h. The pH value of the mixture
8
9 was adjusted to 10 with the diluted ammonia solution. After continually stirred for 2 h,
10
11 the above mixture was then transferred into a 100 ml Teflon-lined and tightly sealed
12
13 autoclave and maintained at 180℃ for 12 h. After the autoclave was cooled down to
14
15 room temperature, the obtained product was dialyzed into neutrality with the deionized
16
17 water and then dried using a vacuum freeze-drying method to obtain CeO2 modified
18
19 reduced graphene oxide nanocomposite, namely CeO2@rGO (CeO2 wt.%:rGO
20
21 wt.%=3:1).
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43 Fig. 1. Illustration of preparation processes of CeO2@rGO.
44
45 X-ray diffraction (XRD) patterns of the nanocomposite was obtained using an X-
46
47 ray diffractometer (XRD, Bruker D8 ADVANCE) with Cu kα radiation (40 kV,
48
49 λ=0.154 nm) at a scan rate of 2°/min within two theta range from 5 to 90°. Raman
50
51 spectra were obtained using a Raman spectrometer (inVia reflex, Renishaw, UK) in the
52
53
range from 400 cm-1 to 3000 cm-1 with a resolution of 4 cm-1. Microstructures of the
54
55
56
composite were characterized using a transmission electron microscope (TEM, FEI-
57
58 Talos, F200X) equipped with the selected area electron diffraction (SAED).
59
60 2.2. Tribology, corrosion resistance and photocatalytic tests
61
62
63
64
65
The tribological performance of pure lubricating oil (PAO5w-40, Netherlands) and
1
2
3
lubricating oil mixed with 3% concentration of rGO, CeO2 and CeO2@rGO was
4
5
evaluated using a ball and disc friction and wear tester (MRS-10G, China). The applied
6
7 load was 10 N, and the rotating speed of the steel disk was 300 r/min. The rotating
8
9 radius of the ball on the steel disk was 3 mm, and the duration was 30 min. The material
10
11 of the grinding ball used in the experiment was GCr15 bearing steel, with a hardness of
12
13 62 HRC and a diameter of 5 mm. The material of the steel disk was quenched and
14
15 tempered 45 steel, with a surface roughness of 0.8-1.6 microns. The wear scar and
16
17 sliding tracks were observed using the SEM and EDX.
18
19 In terms of anti-corrosion coating, 1.8 wt.% sodium molybdate (Na2MoO4), 7.2
20
21 wt.% Tween-20, 4.8 wt.% acrylic resin lotion, 8.4 wt.% ethylene glycol, and 24 wt.%
22
23 zinc aluminum alloy powder (ZnAl20) were mixed thoroughly, and then 0.18 wt.% of
24
25 the CeO2@rGO nanocomposite was added under an ultrasonic agitation. The thickener
26
27 and defoamer (<1 wt.%) were added in a sequence and stirred for about 5~8 h to obtain
28
29 a Dacromet coating solution added with different contents of CeO2@rGO. The coating
30
31 was cold sprayed onto the surface of Q235 steel substrate, and then heated at 85℃ for
32
33
34
15 min, then further sintered/solidified at 280℃ for 25 min. Electrochemical tests were
35
36
carried out using a P4000 electrochemical workstation with a three-electrode system. A
37
38 platinum electrode and saturated calomel electrode were served as the counter and
39
40 reference electrodes, respectively, while the sample was used as the working electrode
41
42 with an active working area of 1 cm2. The electrochemical tests were performed after
43
44 the open circuit potential (OCP) was stabilized in a 3.5% NaCl solution. The
45
46 polarization curves were recorded at a scanning speed of 0.5 mV/s within the range
47
48 from -250 mV to +250 mV. To study the electrochemical performances of the coatings
49
50 with CeO2@rGO, the EIS tests were conducted using the sample with its working area
51
52 of 1 cm2, the frequency ranges from 10-2 Hz to 105 Hz and a sine wave amplitude of 10
53
54 mV. During the electrochemical impedance spectroscopy (EIS) tests, the samples were
55
56 immersed in a 3.5% NaCl solution for different durations, and the impedance and
57
58 corrosion resistance of the coatings were evaluated. The salt spray tests were conducted
59
60 continuously for 720 h in a NaCl solution of 50 g/L, pH=6.5-7.2.
61
62
63
64
65
CeO2@rGO was directly used as a photocatalyst to degrade rhodamine B (RhB)
1
2
3
in a water solution to explore its photocatalytic performance. The nanocomposite was
4
5
dispersed into 15 mg/L RhB aqueous solution with a mass ratio (e.g., CeO2@rGO:RhB)
6
7 of 20:3 under continuously magnetic stirring. Meanwhile, the mixture was exposed to
8
9 UV irradiation (with a UV source power of 30 W and a wavelength of 254 nm). The
10
11 samples were taken out periodically and then centrifuged to study the variations of
12
13 absorbance of the RhB using a UV-5100 UV-Vis spectrophotometer with a wavelength
14
15 of 554 nm. For comparisons, the degradation experiments using the same conditions
16
17 were also performed in a dark environment without any UV light irradiation. In the
18
19 recycle experiments, after the photocatalysts were separated from solution, they were
20
21 washed with ethanol and DI water before being re-dispersed into the RhB dye solution
22
23 for another testing cycle.
24
25 3. Results and discussion
26
27 3.1. Microstructures of CeO2@rGO nanocomposite
28
29 Fig. 2(a) shows XRD patterns of CeO2, CeO2@rGO and GO. It can be seen that
30
31 the diffraction peaks at the 2θ angles of 28.6°, 33.1°, 47.6°, 56.4°, 59.2°, 69.5°, 76.8°,
32
33
34
79.1° and 88.5° are related to the planes of (111), (200), (220), (311), (222), (400), (331),
35
36
(420) and (422) of CeO2, respectively (JCPDS card no. 43-1002). The (001) diffraction
37
38 peak of GO appears at 10.5°, which is corresponding to a d-spacing of 8.4 Å, due to the
39
40 presence of oxygen-containing functional groups on the surface of graphene oxide sheet
41
42 [28]. Another weak (100) diffraction peak of the GO is at 41.8°. Except for the peaks
43
44 corresponding to CeO2, there is no peak of rGO. The reason is attributed to that CeO2
45
46 was generated between the rGO layers after hydrothermal process, which destroys the
47
48 regular structure of rGO [29]. In addition, it may also be that the crystalline peaks of
49
50 CeO2 is relatively high, which covers the weak signal of rGO [30].
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
Fig. 2. (a) XRD patterns of CeO2, GO and CeO2@rGO; (b) Raman spectroscopy of GO and
16
17
18 CeO2@rGO.
19
20 Fig. 2(b) shows Raman spectra of GO and CeO2@rGO nanocomposite. The
21
22 typical D and G bands are found at ~1350 cm-1 and ~1588 cm-1 in both the GO and
23
24 CeO2@rGO. The D band is ascribed to the sp3-hybridized carbon caused by the
25
26 presence of oxygen-containing functional groups on the carbon basal plane and the G
27
28 band is associated with the in-plane stretching vibration of sp2-hybridized carbon [31,
29
30 32]. The intensity ratio of the D band to G band (ID/IG) reflects the defects density in
31
32 the graphitic materials [33]. It is calculated that the ID/IG of CeO2@rGO (1.13) is larger
33
34 than that of GO (0.98), indicating that the structure of GO has many defects which are
35
36 increased during hydrothermal reduction. In addition, the Raman spectra of
37
38 CeO2@rGO exhibit a peak at 462 cm-1, corresponding to the symmetrical stretching
39
40 mode of the Ce-O-Ce vibrational unit, assigned as F2g [34] . The peak of CeO2@rGO is
41
42 very weak, which is related to the small size of CeO2 nanocrystals [35].
43
44
45
Fig. 3(a) shows the FT-IR spectra of the GO, rGO and CeO2@rGO composite. All
46
47
the typical characteristic peaks of rGO are visible, and the peak at 3400 cm-1 is due to
48
49 the stretching vibration mode of the O-H on the GO surface. The peak at 1386 cm-1 is
50
51 attributed to the in-plane bending vibration mode of the C-OH on the rGO surface. The
52
53 peaks at 1720 cm-1 and 1650 cm-1 are due to the stretching vibration modes of the C=O
54
55 and C=C on the carboxyl group, respectively, the peaks at 1230 cm-1 and 1053 cm-1 are
56
57 due to the stretching vibration modes of the C-O-C [36, 37]. In comparison to that of
58
59 rGO, the CeO2@rGO exhibits a new Ce-O peak at 557 cm-1, which implies that the rare
60
61
62
63
64
65
earth oxide CeO2 has been successfully grafted onto the surface of rGO, providing an
1
2
3
effective foundation for its usage as a lubricant additive [38]. Comparing the FT-IR
4
5
spectra of GO and rGO, it can be observed that the peak intensity of the C=O bond
6
7 stretching vibration mode at around 1720 cm-1 is much stronger in rGO than that in GO.
8
9 Furthermore, it can be observed that the peak width of the epoxy group in the GO is
10
11 much wider than that in the rGO at the same position, which is likely due to the partial
12
13 reduction of the product during the reaction process at various reaction temperature and
14
15 reagents.
16
17 Fig. 3(b) shows the XPS survey spectrum of the CeO2@rGO composite. The major
18
19 peaks in the survey spectrum are attributed to C 1s, O 1s, Ce 3d and Ce 4d, indicating
20
21 that the main components of the composite are carbon, oxygen and cerium. Fig. 3(c)
22
23 shows the high-resolution C 1s spectrum of CeO2@rGO, which exhibits a typical C 1s
24
25 peak for the rGO. The curve fitting results reveal three characteristic peaks at 284.8 eV,
26
27 285.9 eV, and 287.8 eV, corresponding to C-C, C-O, and C=O bonds, respectively [39].
28
29 Fig. 3(d) shows the high-resolution Ce 3d spectrum of the CeO2@rGO composite,
30
31 which exhibits a typical Ce 3d peak consisting of a single peak and two multiplet peaks,
32
33
34
indicating the typical Ce 3d peak of CeO2. The Ce 3d peak of CeO2 can be decomposed
35
36
into ten peaks through Gaussian-Lorentz function fitting, corresponding to v0 (880.8
37
38 eV), v (882.8 eV), v' (885.2 eV), v'' (888.5 eV), v''' (898.6 eV), u0 (899.9 eV), u (901.2
39
40 eV), u' (903.6 eV), u'' (906.5 eV), and u''' (916.9 eV). Based on literature, those peaks
41
42 of v0, v', u0, and u' are corresponding to 3d5/2 and 3d3/2, which are the signature states of
43
44 Ce(III) 3d. Whereas those peaks of v, v'', v''', u, u'', and u''' are corresponding to 3d5/2
45
46 and 3d3/2, which are the signature states of Ce(IV) 3d [40]. These results clearly show
47
48 that the CeO2 of the composite is mainly in the form of Ce(IV), but there is also a certain
49
50 amount of Ce(III). This may be due to the reduction of coordination numbers of Ce
51
52 caused by oxygen vacancies in the CeO2 lattice, or it is because Ce3+ is not fully
53
54 oxidized during thermal treatments due to insufficient O sources on the rGO [41].
55
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32 Fig. 3. (a) FT-IR spectra of GO, rGO and CeO2@rGO; (b) XPS survey spectra of
33
34 CeO2@rGO; (c) high resolution C1s spectrum of CeO2@rGO; (d) high resolution Ce 3d spectrum
35
36 of CeO2@rGO.
37
38 Fig. 4 shows the TEM images of rGO and CeO2@rGO. It can be observed from
39
40 Fig. 4(a1) that the rGO nanosheets is transparent and has a layered structure. Its surface
41
42 is relatively flat, although with wrinkles. From the HRTEM images shown in Fig. 4(a2),
43
44
45
it can be seen that the thickness of rGO layered stacking is 0.85 nm to 1.7 nm. The
46
47
attached figure shows that the thickness of rGO monolayer is 0.335 nm, and about 5~7
48
49 atomic layers can be clearly observed. As shown in Fig. 4(a3), the selected area electron
50
51 diffraction (SAED) pattern of rGO is a typical hexagonal diffraction pattern of graphene
52
53 [42]. From the TEM images (see Fig. 4(b1)), the rGO nanosheets are smooth, the
54
55 wrinkles on the translucent sheet disappear, whereas the CeO2 nanoparticles are
56
57 uniformly distributed. From the HRTEM images shown in Fig. 4(b2), the average
58
59 particle size of CeO2 nanoparticles is about 12.24 nm, and the number of rGO
60
61
62
63
64
65
nanosheets is 1~2, which indicates the effective improved dispersion of rGO after
1
2
3
cerium functionalization. The thickness of rGO nanosheets after functionalization is
4
5
reduced, which makes it easy for the rGO to be filled into the coating as nano-
6
7 reinforcement phase. This will form a dense and intercalated structure, which can
8
9 effectively block the channel for corrosion medium to immerse into the substrate. The
10
11 lattice fringes of CeO2 nanoparticles can be observed from the HRTEM images. Its fast
12
13 Fourier transform (FFT) image is linked with the (200) crystal plane of cubic CeO2,
14
15 indicating that CeO2 with a good crystallinity is in-situ grown on the rGO. Fig. 4(b3)
16
17 shows the SAED image of CeO2. A series of polycrystalline diffraction spots in the
18
19 diffraction pattern correspond to the crystal plane spacings of cubic CeO2 (JCPDF card
20
21 no.34-0394).
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
Fig. 4. (a1-a3) and (b1-b3) are TEM and SAED images of rGO and CeO2@rGO, respectively.
47
48
49 According to the characteristic peaks in the XRD spectrum of CeO2@rGO, the
50
51 Debye-Scherrer equation (Eq. 1) was used to estimate the crystallite size of CeO2 [43].
52
53 The average crystallite size was calculated to be ~11.02 nm, which is very close to the
54
55 result obtained by TEM.
56

57 D= βcosθ (1)
58
59
60 where D (nm) is crystallites size. K is Scherrer constant. λ (nm) is wavelength of the X-
61
62
63
64
65
ray sources. β (radians) is the value of full width at the half maximum (FWHM). θ
1
2
3
(radians) is diffraction angle positions.
4
5
3.2. Enhanced tribological performance of lubricating oil by CeO2@rGO
6
7 Fig. 5(a) shows the variation curves of the friction coefficient of lubricating oil
8
9 added with different additives tested under the same conditions. It can be seen that the
10
11 average values of friction coefficients of pure lubricating oil and lubricating oil added
12
13 with rGO, CeO2@rGO and CeO2 are 0.141, 0.107, 0.096 and 0.168, respectively. The
14
15 friction coefficient of lubricating oil added with CeO2@rGO is ~31.9% lower than that
16
17 of pure lubricating oil. Results clearly shows that CeO2@rGO nanocomposite as an
18
19 additive can effectively improve the lubricating ability. On the one hand, due to the
20
21 two-dimensional layered structure of the rGO, the atoms in the layers are bonded
22
23 through covalent bonds, which can improve the stability of the lubrication. At the same
24
25 time, the rGO layers are bonded by van der Waals force, and they are easily slide under
26
27 the action of shearing force. In addition, 2D structured rGO has a higher surface area
28
29 than that of the traditional materials. When they are existed in the friction interface,
30
31 they can cover a large area of the worn surfaces, thus reducing the probability of direct
32
33
34
contact of two counterfaces and decreasing the friction. On the other hand, CeO2, as a
35
36
rare earth oxide, has the characteristics of large electronegativity and high activity due
37
38 to its special 4f electronic configuration. It can not only clean the surface of graphene,
39
40 but also form C-O-Ce or mixed hybrid to make a stable rGO. At the same time, a friction
41
42 film was mainly composed of chemical reacted films of Fe2O3, which was formed
43
44 during the friction process to avoid direct contact of the friction pair, thus improving
45
46 the friction reduction and lubrication performance [44]. Therefore, CeO2 will not only
47
48 protect the structure of rGO, but also reduce the surface energy of rGO and improve its
49
50 uniform dispersion in the substrate, thus improving the lubricating ability of the pure
51
52 lubricating oil.
53
54 To confirm the effect of CeO2 modified reduced graphene oxide nanocomposites
55
56 on the contact surface, Raman spectra at any two positions on the surface of 45 steel
57
58 were obtained after the friction tests of lubricating oil containing CeO2@rGO, and the
59
60 obtained results are shown in Fig. 5(b). It can be observed that the Ce-O-Ce vibrational
61
62
63
64
65
mode (F2g) shows a typical band at 535.2 cm-1. Compared with the Raman spectrum of
1
2
3
CeO2@rGO nanocomposite, the F2g peak of Ce-O-Ce is shifted to the right-hand side
4
5
by 70.2 cm-1, which is due to the deflection of its diffraction peaks under the external
6
7 load, consistent with that reported in literature [45]. At the same time, the A1g peak of
8
9 Fe3O4 is appeared at 666.8 cm-1, which is because the Fe element in 45 steel reacts with
10
11 oxygen in the air during the friction experiment [46]. The typical D and G bands of
12
13 CeO2@rGO are appeared at ~1350 cm-1 and ~1590 cm-1, and the 2D band are appeared
14
15 at ~2700 cm-1, indicating that rGO effectively fills the pits on the surface of the friction
16
17 pairs during the friction process and forms a good lubricating film. However, results
18
19 also show that the peak intensities of D and G bands after the wear tests are significantly
20
21 lower than those of CeO2@rGO nanocomposite before the friction tests, and the
22
23 intensity ratio of D and G bands are also lower than that of the original CeO2@rGO
24
25 nanocomposite. These results indicate that the structure of CeO2@rGO nanocomposite
26
27 in the worn surface is different from that of the original nanocomposite before the wear
28
29 tests. This is because CeO2@rGO nanocomposite was damaged under the action of
30
31 friction and shear force after entering the friction contact area, resulting in more defects
32
33
34
and structural changes.
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
Fig. 5. (a) The variation curves of the friction coefficient of lubricating oil with different additives;
1
2 (b) Raman spectra of the wear scar on the surface of 45 steel after friction experiments with
3
4
lubricating oil containing CeO2@rGO; (c-e) SEM and EDX images of the wear scar on 45 steel
5
6
7 surface after friction experiments with different additives: (c-c1) pure lubricating oil; (d-d1)
8
9 lubricating oil containing rGO; (e-e1) lubricating oil containing CeO2@rGO.
10
11 Figs. 5(c-e) show the SEM images and EDX results of the worn track on the
12
13 surface of 45 steel after the friction tests using pure lubricating oil, lubricating oil
14
15 containing rGO and CeO2@rGO. Fig. 5(c-c1) shows that when the pure lubricating oil
16
17 was used as the lubricating medium, the worn surface of 45 steel shows deep wear
18
19 marks and gullies. The number of formed furrows is obvious, and the groove marks are
20
21 quite wide. The composition analysis of the worn surface shows that it only contains
22
23 Fe elements, indicating that no other impurity elements are introduced during the wear
24
25 tests. Fig. 5(d-d1) shows that when the lubricating oil containing the rGO nanoparticles
26
27 is used as the lubricating medium, the gullies and pear grooves on the surface of 45
28
29 steel are much shallow, although there is more abrasive wear observed. The
30
31 composition analysis of the wear mark area shows that only C and Fe elements are
32
33
34
detected, which indicates that rGO nanoparticles are covered on the surface of the
35
36
friction pair during the friction process.
37
38 Results showed that adding a certain amount of rGO nanoparticles improved the
39
40 lubricating performance of the lubricating oil. However, due to the poor dispersion
41
42 stability of rGO in the lubricating oil, agglomeration phenomenon was observed which
43
44 greatly affects the lubricating effect of rGO as a lubricating additive. However, after
45
46 the friction test of the lubricating oil containing CeO2@rGO, the furrow phenomenon
47
48 in the wear mark area on the surface of 45 steel became insignificant, and the surface
49
50 was relatively smooth without any protrusions and depressions (see Fig. 5(e-e1)). These
51
52 results show that the friction phenomenon in the wear mark area on the surface of 45
53
54 steel is gradually weakened with the increase of Ce3+ concentration, and the grooves
55
56 and gullies on the surface are gradually become much shallower. The composition
57
58 analysis of the wear mark area shows that the worn surface only contains C, O, Ce, and
59
60 Fe elements, which indicates that CeO2@rGO nanocomposites are successfully
61
62
63
64
65
deposited on the surface of the friction pair during the friction experiment. This is
1
2
3
consistent with the analysis in Figs. 5(a) and 5(b). To sum up, the wear mechanism is
4
5
changed from oil-film lubrication to reaction film lubrication after adding the
6
7 nanocomposite into the lubricating oil.
8
9 3.3. Enhanced corrosion resistance of coating by CeO2@rGO
10
11 Fig. 6 shows SEM images and EDX results of the coatings with 0%, 0.18%, and
12
13 0.27% CeO2@rGO after the 720 h salt spray test. It can be seen that the pore contents
14
15 on the surface of the coatings with 0.18%, 0.09%, 0%, and 0.27% CeO 2@rGO are
16
17 gradually increased. From the enlarged images, the morphology of the coating surface
18
19 is gradually changed from a metallic powder shape to a sponge-like structure, and
20
21 finally to a typical needle-shaped structure of the corrosion product. Among them, the
22
23 surface of the coating with 0.18% CeO2@rGO is still very dense, and the morphology
24
25 of the metallic powder can still be observed. Whereas for the coatings with 0% and
26
27 0.27% CeO2@rGO, the typical corrosion products of the coatings, i.e., zinc aluminum
28
29 chloride hydrates, have an obvious needle-shaped structure. EDX analysis of the
30
31 surface of each coating shows that the surface of the coating with 0.18% CeO2@rGO
32
33
34
is mainly composed of Zn, Al, and O elements. Compared with the uncorroded coating,
35
36
the amount of O has been increased, indicating that the metallic powder has generated
37
38 zinc aluminum oxides after the corrosion test. However, the corroded coating is still
39
40 effective for shielding. For the coatings with 0% and 0.27% CeO2@rGO, Fe elements
41
42 were detected, and there was more oxygen detected, indicating that more corrosion
43
44 products and Fe oxides were generated on the surface of the coating. This can also be
45
46 clearly seen from the image where the red rust was produced on the surface of the
47
48 substrate. The reason why the EDX results of the coating with 0.18% CeO2@rGO did
49
50 not show Fe elements is that this coating has the best corrosion resistance. It can also
51
52 be seen that there was no red rust produced on the surface of the substrate coated with
53
54 this composite layer.
55
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22 Fig. 6. SEM and EDX images of surface of chromium-free Dacroment coating reinforced with
23
24
different contents of CeO2@rGO before and after corrosion: (a-a4) 0% CeO2@rGO; (b-b4) 0.09%
25
26
27
CeO2@rGO; (c-c4) 0.18% CeO2@rGO; (d-d4) 0.27% CeO2@rGO.
28
29 Fig. 7(a) shows the Tafel curves of the coatings with different contents of
30
31 CeO2@rGO. Except for the coating with 0.27% CeO2@rGO, the corrosion potentials
32
33 of all the other coatings are lower than that of the substrate. The potential difference is
34
35 between 0.0504 V to 0.1092 V. All these results indicate that the coatings added with
36
37 CeO2@rGO can provide a certain degree of cathodic protection for the substrate [47].
38
39 Comparing the coatings added with different contents of CeO2@rGO, the corrosion
40
41 potential of the coating with 0.18% CeO2@rGO shows the most negative value, and the
42
43 corrosion potential of the coating with 0.27% CeO2@rGO shows the most positive
44
45 value. In addition, the corrosion current densities of the coatings with 0.09% and 0.18%
46
47 CeO2@rGO are decreased by an order of magnitude if compared to that of the coating
48
49 without adding the CeO2@rGO. The corrosion resistance was significantly increased
50
51 after adding the CeO2@rGO, indicating that a certain amount of CeO2@rGO can reduce
52
53
the Icorr value of the coating and improve its corrosion resistance. However, the Icorr
54
55
56
value of the coating with 0.27% CeO2@rGO is increased. This is because after adding
57
58 an excessive amount of CeO2@rGO, the rGO could become contacted with each other
59
60 and can form an effective corrosion path, thus increasing the conductivity of the coating
61
62
63
64
65
and the corrosion current density. It is also possible that adding too much amount of
1
2
3
composite causes severe agglomeration, producing more pores for the corrosion
4
5
medium easier to pass through, thus reducing the corrosion resistance of the coating.
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34 Fig. 7. (a) Tafel polarization curves of chromium-free Dacromet coatings reinforced with different
35
36 contents of CeO2@rGO; (b) Nyquist diagram of chromium-free Dacromet coatings reinforced
37
38 with different contents of CeO2@rGO; Bode diagram of chromium-free Dacromet coatings
39
40 reinforced with different contents of CeO2@rGO: (c) phase frequency diagram; (d) amplitude-
41
42 frequency diagram.
43
44
45
Fig. 7(b) shows the Nyquist diagram of the coatings with different contents of
46
47
CeO2@rGO in a 3.5% NaCl solution. It can be seen that the capacitive arc radii of the
48
49 coatings with 0.09% and 0.18% CeO2@rGO are much larger than that of the coating
50
51 with 0% CeO2@rGO, and the density of the coating is increased after adding a certain
52
53 amount of composite, both of which can effectively improve the corrosion resistance
54
55 of the coating [48]. This is because the graphene is uniformly dispersed in the coating.
56
57 It can effectively fill up the defects in the coating, which plays a positive role as a
58
59 physical barrier to hinder the intrusion of the corrosive medium. When the CeO2@rGO
60
61
62
63
64
65
content is 0.27%, the capacitive arc radius of the coating decreases, indicating that
1
2
3
adding an excessive amount of composite will reduce its dispersion in the coating,
4
5
leading to agglomeration, increased defects in the coating, and also decreased corrosion
6
7 resistance of the coating.
8
9 Figs. 7(c-d) show the Bode diagrams of coatings added with different amounts of
10
11 the composite. From the frequency response diagram, the |Z| values of the coatings with
12
13 0.09% and 0.18% CeO2@rGO are higher than those of the coatings without adding the
14
15 composite. Among them, the coating with 0.18% CeO2@rGO exhibits the maximum
16
17 low-frequency impedance value (|Z|f→0) of about 3.68×103 Ω·cm2, indicating its
18
19 minimum corrosion rate and relatively good corrosion resistance. When the added
20
21 amount of CeO2@rGO reaches 0.27%, the |Z| value of the coating becomes decreased,
22
23 and its corrosion resistance is also decreased. This again proves that adding a certain
24
25 amount of composite enhances the corrosion resistance of the coating, but adding
26
27 excessive amounts of composite will reduce the corrosion resistance of the coating.
28
29 This may be due to the van der Waals force which causes the rGO to be agglomerated
30
31 locally, thus increasing the pores in the coating and decreasing its shielding effect.
32
33
34
Fig. 8 shows the equivalent circuit model of the coating’s electrochemical
35
36
impedance spectrum, which is Rs(Cc(Rc(QfRf)))(CpRp). Here, Rs, Rc, Rf, and Rp are the
37
38 resistances of the solution, coating, surface film layer of zinc-aluminum metal powder,
39
40 and surface activation dissolution reaction of zinc-aluminum metal powder,
41
42 respectively. Cc, Qf, and Cp are the capacitances of the coating, surface film layer of
43
44 zinc-aluminum alloy powder, and surface activation dissolution reaction of zinc-
45
46 aluminum metal powder, respectively. Table 1 summarizes the fitting results of the
47
48 electrochemical impedance spectra of all the coatings. It can be seen that the Rc values
49
50 increase with the increase of the amount of composites. The Rc values of coatings with
51
52 0%, 0.09%, and 0.18% CeO2@rGO are relatively small, around 25 Ω·cm2, indicating
53
54 that the metallic powder is not completely covered by the passivation film, and cannot
55
56 provide effective cathodic protection to the substrate. In Table 1, Rf represents the
57
58 resistance of the corrosion product film formed on the surface of the metallic powder.
59
60 The larger the value of Rf, the denser the corrosion product layer formed on the surface
61
62
63
64
65
of the metallic powder, and the better its protection effect. This will make it more
1
2
3
difficult for the metallic powder to undergo activation dissolution, thus enhancing the
4
5
anti-corrosion resistance of the coating. In addition, the value of Rf increases firstly but
6
7 then decreases with the further increase of CeO2@rGO. Among them, the coating with
8
9 0.18% CeO2@rGO exhibits the largest Rf value of 4064 Ω·cm2, indicating that the
10
11 corrosion product film formed by this coating is the densest, thus effectively forming a
12
13 physical barrier and delaying the corrosion of the metallic powder. Rp represents the
14
15 resistance of the activation dissolution reaction of the metallic powder. The larger the
16
17 value of Rp, the more difficult the reaction of the zinc-aluminum powder can occur, and
18
19 the better the corrosion resistance of the coating. According to the Table 1, Rp value is
20
21 increased with increasing the amounts of CeO2@rGO, and the coating added with 0.27%
22
23 CeO2@rGO exhibits the largest Rp. Since both the Rf and Rp values are critical to the
24
25 corrosion resistance of the coating, the sum of Rf and Rp is used in this study to represent
26
27 the corrosion resistance of the coating, with a larger Rf+Rp indicating a better corrosion
28
29 resistance. As shown in the Table 1, the coating with 0.18% CeO2@rGO exhibits the
30
31 largest Rf+Rp value, indicating its best corrosion resistance.
32
33
34
35
36
37
38
39
40
41
42 Fig. 8. Equivalent circuit diagram.
43
44 Table 1. Fitting parameters of electrochemical impedance spectrum for chromium-free Dacormet
45
46 coatings reinforced with different contents of CeO2@rGO
47
48 wt.% Rs Cc Rc CPEf n Rf Cp Rp
49 Ω·cm2 F/cm2 Ω·cm2 S-n/cm2 Ω·cm2 F/cm2 Ω·cm2
50
0% 14.92 7.353×10-7 24.47 3.203×10-4 0.7175 2712 4.212×10-8 64.16
51
52 0.09% 13.44 6.706×10-7 25.47 2.784×10-4 0.6558 3985 3.994×10-8 66.31
53 0.18% 11.45 1.321×10-7 26.51 3.145×10-4 0.6434 4064 4.459×10-8 74
54
55 0.27% 12.74 6.024×10-8 54 5.357×10-4 0.5357 3081 8.404×10-8 608.1
56
57
Fig. 9 shows the FT-IR spectrum of the Q235 steel substrate with the chromium-
58
59 free Dacromet coating with the CeO2@rGO. It can be seen that the coating exhibits the
60
61
62
63
64
65
characteristic peaks corresponding to those of O-Ce, C-O-Si and Si-O-Si bonds [49,
1
2
3
50]. The hydrolyzed silane alcohol produced by the silane coupling agent in the coating
4
5
can form C-O-Si bond, and at the same time, the silane alcohol itself becomes
6
7 dehydrated and condensed to form linear or cyclic Si-O-Si network structures. The
8
9 Ce(OH)3 passivation layer is formed at the interface between the substrate and coating,
10
11 which prevents direct contact between the substrate and the external environment [51].
12
13 The formation of Si-O-Fe and Ce-Fe bonds at the interface between the coating and
14
15 substrate is also beneficial for improving the adhesion between the coating and
16
17 substrate, thereby improving its corrosion resistance [32, 52, 53]. Formation of these
18
19 bonds can also improve the density of the coating and enhance the bonding strength
20
21 between the coating and substrate.
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41 Fig. 9. FT-IR spectrum of the coatings with CeO2@rGO.
42
43 3.4. Enhanced photocatalytic performance by CeO2@rGO
44
45 Fig. 10(a) shows the photodegradation curves of the organic dye RhB by using
46
47 CeO2, GO and CeO2@rGO in the dark and under UV irradiation. It can be seen that the
48
49 removal of RhB by GO mainly occurs in the adsorption process under dark conditions,
50
51 and the removal rate under the UV irradiation is not decreased significantly, with the
52
53
maximum removal rate of 53.2%. CeO2 also mainly removes the RhB through the
54
55
56
adsorption process under dark conditions. However, under the UV irradiation, it
57
58 exhibits photocatalytic activity, with the maximum removal rate of 20.3%. CeO2@rGO
59
60 achieves a high removal rate of 90.5% for RhB, of which 77.8% is removed through
61
62
63
64
65
the adsorption process under dark conditions. The subsequent photocatalytic
1
2
3
degradation process further increases the removal rate up to 90.5%. Fig. 10(b) shows
4
5
the UV-vis spectra before and after the RhB adsorption. This shows that the RhB has a
6
7 maximum absorption peak at 554 nm before adsorption, but this peak is gradually
8
9 weakened after the RhB removal. The reason for this phenomenon may be that the rGO
10
11 nanosheets are excellent electron acceptors, while the UV-excited CeO2 is an excellent
12
13 electron donor. Therefore, the interactions between CeO2 and rGO nanosheets can
14
15 cause excited electrons to be effectively transferred from CeO2 to rGO nanosheets, thus
16
17 reducing the possible recombination of electron-hole pairs. Modifications of covalent
18
19 bonding between rGO and CeO2 can also improve the dispersion of rGO in the solution.
20
21 Furthermore, the layered structure of rGO can prevent the severe aggregation of CeO2
22
23 nanoparticles, thereby exposing more active sites to enhance the photocatalytic
24
25 performance of the CeO2@rGO composite.
26
27 Adsorption and photocatalytic efficiency the CeO2@rGO nanocomposite was
28
29 evaluated by analyzing the removal of organic dye RhB in both the dark environment
30
31 and under UV irradiation. Fig. 10(c) shows the concentration changes of the RhB dye
32
33
34
solution with time for the CeO2@rGO at different experimental conditions. The red
35
36
curve is the degradation curve of CeO2@rGO to the RhB under UV light conditions,
37
38 and the black one is the curve of adsorption-desorption equilibrium of CeO2@rGO to
39
40 the RhB dye in the dark conditions. After 15 min testing, the concentration of RhB
41
42 shows a plateau (see the black line in Fig. 10(c)), indicating that the adsorption-
43
44 desorption equilibrium between the nanocomposite and dye molecules is reached after
45
46 15 min. It can be seen that ~88% of the RhB dye is absorbed at the equilibrium
47
48 adsorption state under dark reaction conditions, indicating that the CeO2@rGO has
49
50 shown as good physical adsorption ability. The rGO has a large specific surface area,
51
52 and CeO2 nanoparticles are uniformly dispersed on the surface of rGO, exposing
53
54 numerous active sites, so that the CeO2@rGO composite has significantly improved the
55
56 adsorption activity for an organic dye (RhB). The adsorption kinetics can be described
57
58 using Eq. 2 [53]:
59
60 ln(C/C0) = -kt (2)
61
62
63
64
65
in which the ratio of C (e.g., the concentration of RhB at an adsorption time of t) to C0
1
2
3
(e.g., the initial concentration of RhB) can be given from the relative intensity ratio of
4
5
the respective absorbance (A/A0). k is the apparent first-order rate constant (min-1) and
6
7 t is adsorption time. Fig. 10(d) shows the ln(C/C0) versus time and a good linear
8
9 relationship is observed with the correlation coefficients (R2) value of 0.991, indicating
10
11 that the adsorption process follows the pseudo-first-order kinetics. Calculated from the
12
13 slope of the straight line, the kinetic adsorption rate constants k is 0.807 min-1.
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
Fig. 10. (a) Photodegradation curves of different concentrations for RhB by the CeO2, GO and
58
59
60 CeO2@rGO under UV irradiation; (b) UV-vis spectra of RhB before and after photocatalytic
61
62
63
64
65
degradation; (c) Concentration changes of RhB dye solutions with time for CeO2@rGO in the
1
2 dark and under UV irradiation; (d) The pseudo-first-order reaction kinetics of the CeO2@rGO
3
4
composite for RhB degradation; (e) Comparisons of the removal efficiency for RhB dye of
5
6
7 different CeO2 based catalyst; (f) Recycling performance of the CeO2@rGO during adsorption-
8
9 photodegradation of the RhB solution.
10
11 Under UV irradiation conditions, it is observed that the concentration of RhB is
12
13 decreased rapidly in the first 15 min due to the superior adsorption ability of
14
15 CeO2@rGO. Compared with those in the dark conditions, the RhB concentration is
16
17 gradually decreased from 15 min to 120 min, which is attributed to the photo-
18
19 degradation of RhB enhanced by CeO2. In the initial stage, CeO2@rGO rapidly adsorbs
20
21 the RhB dye molecules due to the action of electrostatic attraction. As the adsorption
22
23 process is progressed further, the electrostatic attraction between CeO2@rGO and RhB
24
25 molecules becomes weaker, eventually reaching an adsorption-desorption equilibrium.
26
27 When exposed to the UV light, the CeO2 nanostructures on the rGO generate electrons
28
29 and holes. The rGO in the nanocomposite acts as both acceptor and transporter for the
30
31 electrons due to its good conductivity. Therefore, the photogenerated electrons from the
32
33
34
CeO2 is effectively transferred to the rGO nanosheets [32]. The photogenerated holes
35
36
are transferred to the surface of CeO2, and after the reaction, they will degrade the
37
38 adsorbed RhB dye molecules, so that the concentration of RhB solution will continue
39
40 to decrease. Results show that about 90.5% of RhB is degraded by CeO2@rGO after
41
42 120 min. As noted in Fig. 10(e), the removal efficiency for the RhB dye of our work is
43
44 significantly higher than those of the other CeO2 based catalyst reported in literature
45
46 [54-59]. Fig. 10(f) shows the cycling performance of the nanocomposite for the RhB
47
48 photocatalytic degradation. The nanocomposite shows only slight decrease in the
49
50 photocatalytic activity after three cycles, which is mainly due to an incomplete removal
51
52 of the RhB decomposition products during the regeneration process and the
53
54 photoetching induced by accumulation of photogenerated holes on the CeO2 surface
55
56 [60].
57
58 4. Mechanisms of CeO2@rGO in lubrication, anti-corrosion and photocatalysis
59
60 Based on the characteristics for the integrated composite of rGO and CeO2, we
61
62
63
64
65
proposed the mechanisms of using CeO2@rGO nanocomposites for improvement of
1
2
3
tribological properties, enhancement of corrosion resistance of coating, and
4
5
photocatalytic performance for degradation of dyes, as shown in Fig. 11.
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49 Fig. 11. Schematic illustrations of mechanisms using CeO2@rGO nanocomposite for lubrication,
50
51 anti-corrosion and photocatalysis applications. (a) Lubrication mechanisms of different types of
52
53 lubricating oil (in which (a1) is for pure lubricating oil; lubricating oil with (a2) CeO2; (a3) rGO;
54
55
(a4) CeO2@rGO); (b) Anti-corrosion mechanism of chromium-free Dacromet coating added with
56
57
CeO2@rGO ((b1) before adding CeO2, (b2) after adding, (b3) interfacial structure); (c)
58
59
60 Photocatalytic mechanism.
61
62
63
64
65
When the lubricating oil is used as a lubricating medium, it will become "extruded"
1
2
3
under the action of large contact stress. This will cause the direct contact between the
4
5
steel ball and substrate, resulting in increased friction and wear [61], as shown in Fig.
6
7 11(a1). When CeO2 is added inside the lubricating oil, the three-dimensional structure
8
9 of CeO2 nanoparticles can effectively fill the pits on the friction pair surface and
10
11 smoothen the surface. However, the surface energy of CeO2 nanoparticles is still quite
12
13 large, and thus they are easily agglomerated in the lubricating oil, which will reduce the
14
15 lubricating performance of the lubricating oil [62], as shown in Fig. 11(a2). When
16
17 graphene nanosheets are used as the additives, the layered structure can fill the pits of
18
19 friction pairs, and graphene nanosheets are easily broken under the action of shearing
20
21 force, thus forming a good lubricating film to prevent the direct contact of friction pairs
22
23 [63], as shown in Fig. 11(a3). Similarly, graphene itself is easily agglomerated due to
24
25 its large surface energy. Using the CeO2 modified reduced graphene oxide
26
27 nanocomposites (CeO2@rGO) can solve this issue as a good lubricating oil additive, as
28
29 shown in Fig. 11(a4). Rare earth elements such as Ce easily interact with the oxygen-
30
31 containing functional groups on the surface of rGO to form O-Ce. Grafting CeO2 onto
32
33
34
the surface of rGO can effectively reduce the surface energy of rGO, thus dispersing
35
36
the CeO2 uniformly in the lubricating oil and avoiding the severe agglomeration. In
37
38 addition, CeO2@rGO nanocomposites have a dual lubrication mechanism from both
39
40 the CeO2 nanoparticles and rGO nanosheets. In this case, the surface of friction pairs
41
42 can be worn smoothly by filling the vacancies with a polishing effect, and a good
43
44 lubricating film can be formed between friction pairs to prevent any direct contact of
45
46 friction pairs. The PAO5w-40 lubricating oil containing CeO2@rGO can not only fill
47
48 the rough physical interface, but also form a lubricating film, greatly reducing the
49
50 friction coefficient. Accordingly, the friction and wear resistance of PAO5w-40
51
52 lubricating oil with the nanocomposites as additive has been significantly improved.
53
54 Figs. 11(b1) to 11(b3) show the mechanisms of corrosion resistance of the
55
56 CeO2@rGO composites which are dispersed uniformly in the coating. In this study of
57
58 the chromium-free Dacromet coating, we used the molybdic acid instead of the
59
60 traditionally used chromic acid. The corrosion resistance mechanism of the coating has
61
62
63
64
65
also been changed to the formation of insoluble molybdate on the substrate as shown
1
2
3
in Eq. 3. The formed oxide film on the substrate’s surface passivates it from corrosion,
4
5
which competes with Cl- during corrosion, thus reducing the corrosion of Cl-. In
6
7 addition, the corrosion resistance performance of the coating is further improved after
8
9 the addition of CeO2@rGO. The modified rGO is uniformly dispersed in the coating,
10
11 and its layered structure easily adheres to the zinc-aluminum alloy powder and plays a
12
13 key role in the solidification of the CeO2@rGO. On the other hand, the nanoflakes and
14
15 nanoparticles fill the porous structure of the coating, greatly increasing its density and
16
17 hardness. The homogeneous distribution of the CeO2@rGO in the coating promotes the
18
19 formation of a conductive channel, facilitates the electron transfer, and is thus beneficial
20
21 for sacrificing zinc-aluminum alloy with a lower potential for protecting the substrate.
22
23 The hydrolysis of the silane coupling agent in the coating produces a silanol that can
24
25 form C-O-Si, and Si-O-Fe is also formed at the interface between the coating and the
26
27 substrate [49]. At the same time, the self-dehydration and condensation of silanol form
28
29 a linear or cyclic Si-O-Si network structure. The Ce(OH)3 passivation layer is formed
30
31 at the interface between the substrate and the coating, which actively prevents direct
32
33
34
contact between the substrate and the external environment. This indicates that the
35
36
compound in the anti-corrosion film has barrier/active dual-functions. The presence of
37
38 Ce-Fe bonds can also improve the adhesion between the coating and substrate, thereby
39
40 improving the wear resistance and corrosion resistance of the coating [52]. Therefore,
41
42 the corrosion resistance of the chromium-free Dacromet coating with the added
43
44 CeO2@rGO has been improved.
45
46 MoO42- + Fe2+ → FeMoO4 → Fe2(MoO4)3 (3)
47
48 Fig. 11(c) shows the photocatalytic degradation mechanism of the composite for
49
50 RhB solution. The rGO acts as both acceptor and transporter for the photo-generated
51
52 electrons due to its good conductivity, thus prolonging the life of electron/hole pairs
53
54 [32]. The interactions between CeO2 and rGO can lead to the effective transfer of
55
56 excited electrons from CeO2 to the surface of rGO nanosheets, thus reducing the
57
58 electron-hole pairs recombination [54]. The 2D π-conjugation system of rGO facilitates
59
60 the easy movement of electrons towards the conduction band to participate in
61
62
63
64
65
degradation reaction (generating superoxide radical which in turn produces hydroxyl
1
2
3
radicals) rather than recombination with holes. In addition, the uniform loading of rare
4
5
earth oxides on the rGO surface without significant agglomeration helps to expose more
6
7 active sites. Due to these reasons, the CeO2@rGO nanocomposites will have good
8
9 properties for photocatalysis.
10
11 5. Conclusions
12
13 In this paper, CeO2 modified reduced graphene oxide nanocomposite (CeO2@rGO)
14
15 was successfully prepared with cerium nitrate hexahydrate and graphene nanosheets as
16
17 the raw materials using a hydrothermal method. When the nanocomposite was added
18
19 into PAO5w-40 lubricating oil, the friction coefficient was 31.9% lower than that of
20
21 pure lubricating oil. The wear mechanism was changed from oil-film lubrication to
22
23 reaction film lubrication. The corrosion current density of the chromium-free Dacromet
24
25 coating added with the nanocomposite was decreased by an order of magnitude
26
27 compared with that of the coating without adding the nanocomposite. This is mainly
28
29 due to the addition of the composites will increase the density of the coating and form
30
31 a conductive channel which easily corrode the alloy powder at a much lower potential.
32
33
34
When it was used as the photocatalyst, the removal rate of RhB dye can reach 90.5%,
35
36
because rGO can reduce the recombination of electron-hole pairs and prevent CeO2
37
38 from agglomeration, thereby exposing more active sites and extending the life of
39
40 electron-hole pairs.
41
42 Acknowledgement
43
44 The authors would like to acknowledge the financial supports from the Science and
45
46 Technology Project of Shaanxi Province of China (No.2020ZDLGY12-06), Xi’an
47
48 Science research project of China (No.2021XJZZ0042) and International Exchange
49
50 Grant (IEC/NSFC/201078) through Royal Society and National Science Foundation of
51
52 China (NSFC).
53
54 Conflicts of interest
55
56 We declare that we do not have any commercial or associative interest that
57
58 represents a conflict of interest in connection with the work submitted.
59
60
61
62
63
64
65
1 References
2
3 [1] Lee C, Li Q, Kalb W, Liu X, Berger H, Carpick R and Hone J 2010 Frictional
4 Characteristics of Atomically Thin Sheets Science 328 76
5
6 [2] Kim K, Lee H, Lee C, Lee S, Jang J, Ahn J, Kim J and Lee H 2011 Chemical Vapor
7 Deposition-Grown Graphene: The Thinnest Solid Lubricant ACS Nano 5 5107
8 [3] Berman D, Erdemir A and Sumant A 2014 Graphene: a new emerging lubricant
9
10 Materials Today 17 31.
11 [4] Xu H, Zang J, Yuan Y, Tian P and Wang Y 2019 Preparation of multilayer graphene
12
13
coatings with interfacial bond to mild steel via covalent bonding for high performance
14 anticorrosion and wear resistance Carbon 154 156.
15 [5] Cui G, Bi Z, Zhang R, Liu J, Yu X and Li Z 2019 A comprehensive review on
16
17 graphene-based anti-corrosive coatings Chemical Engineering Journal 373 104.
18 [6] Ji X, Lin Y, Zeng J, Ren Z, Lin Z, Mu Y, Qiu Y and Yu J 2021
19 Graphene/MoS2/FeCoNi(OH)x and Graphene/MoS2/FeCoNiPx multilayer-stacked
20
21 vertical nanosheets on carbon fibers for highly efficient overall water splitting Nature
22 Communication 12 1380.
23
24
[7] Dong C, Lu J, Qiu B, Shen B, Xing M and Zhang J 2018 Developing stretchable
25 and graphene-oxide-based hydrogel for the removal of organic pollutants and metal
26 ions Applied Catalysis B: Environmental 222 146.
27
28 [8] Wang J, Jin X, Li C, Wang W, Wu H and Guo S 2019 Graphene and graphene
29 derivatives toughening polymers: Toward high toughness and strength Chemical
30 Engineering Journal 370 831.
31
32 [9] Hu X, Deng F, Huang W, Zeng G, Luo X and Dionysiou D 2018 The band structure
33 control of visible-light-driven rGO/ZnS-MoS2 for excellent photocatalytic degradation
34
35
performance and long-term stability Chemical Engineering Journal 350 248.
36 [10] Ho P, Lee W, Liou Y, Chiu Y, Shih Y, Chen C, Su P, Li M, Chen H, Liang C and
37 Chen C 2015 Sunlight-activated graphene-heterostructure transparent cathodes:
38
39 enabling high-performance n-graphene/p-Si Schottky junction photovoltaics Energy &
40 Environmental Science 8 2085.
41 [11] Lee B, Koo M, Jin S, Kim K and Hong S 2014 Simultaneous strengthening and
42
43 toughening of reduced graphene oxide/alumina composites fabricated by molecular-
44 level mixing process Carbon 78 212.
45
46
[12] Peng Y, Lu B and Chen S 2018 Carbon-Supported Single Atom Catalysts for
47 Electrochemical Energy Conversion and Storage Advanced Materials 30 1801995.
48 [13] Feng Y, Zhang S, Zhu L, Li G, Zhao N, Zhang H andChen B 2022 Reduced
49
50 graphene oxide-supported ruthenium nanocatalysts for highly efficient electrocatalytic
51 hydrogen evolution reaction International Journal of Hydrogen Energy 47 39853.
52 [14] Yang Z, Liu C, Liu X, Du Y, Jin H, Ding F, Li B, Ouyang Y, Bai L and Yuan F
53
54 2022 In-situ synthesis of graphene nanosheets encapsulated silicon nanospheres by
55 thermal plasma for ultra-stable lithium storage Carbon 199 424.
56
57
[15] Gong C, Hinojos D, Wang W, Nijem N, Shan B, Wallace R, Cho K and Chabal Y
58 2012 Metal-Graphene-Metal Sandwich Contacts for Enhanced Interface Bonding and
59 Work Function Control ACS Nano 6 5381.
60
61
62
63
64
65
[16] Sun Y, Liu Q, Gao S, Cheng H, Lei F, Sun Z, Jiang Y, Su H, Wei S and Xie Y 2013
1
Pits confined in ultrathin cerium(IV) oxide for studying catalytic centers in carbon
2
3 monoxide oxidation Nature Communication 4 2899.
4 [17] Zhang Z, Wang Z, Liang B and La P 2006 Effects of CeO2 on friction and wear
5
6 characteristics of Fe–Ni–Cr alloy coatings Tribology International 39 971.
7 [18] Pereira G, Prada O, Avila P, Avila J, Pinto H, Miyazaki M, de Melo H and Bose W
8 2022 Cerium conversion coating and sol-gel coating for corrosion protection of the
9
10 WE43 Mg alloy Corrosion Science 206 110527.
11 [19] Liang X, Wang P, Gao Y, Huang H, Tong F, Zhang Q, Wang Z, Liu Y, Zheng Z,
12
13
Dai Y and Huang B 2020 Design and synthesis of porous M-ZnO/CeO2 microspheres
14 as efficient plasmonic photocatalysts for nonpolar gaseous molecules oxidation: Insight
15 into the role of oxygen vacancy defects and M=Ag, Au nanoparticles Applied Catalysis
16
17 B: Environmental 260 118151.
18 [20] Boshui C, Kecheng G, Jianhua F, Jiang W, Jiu W and Nan Z 2015 Tribological
19 characteristics of monodispersed cerium borate nanospheres in biodegradable rapeseed
20
21 oil lubricant Applied Surface Science 353 326.
22 [21] Ali M and Xianjun H 2022 Exploring the lubrication mechanism of CeO2
23
24
nanoparticles dispersed in engine oil by bis(2-ethylhexyl) phosphate as a novel antiwear
25 additive Tribology International 165 107321.
26 [22] Anadebe V, Chukwuike V, Ramanathan S and Barik R 2022 Cerium-based metal
27
28 organic framework (Ce-MOF) as corrosion inhibitor for API 5L X65 steel in CO2-
29 saturated brine solution: XPS, DFT/MD-simulation, and machine learning model
30 prediction Process Safety and Environmental Protection 168 499.
31
32 [23] Sujay S, Alkanad K, Thejaswini B, Alnaggar G, Zaqri N, Drmosh Q, Boshaala A
33 and Lokanath N 2022 Alkaline mediated sono-synthesis of surface oxygen-vacancies-
34
35
rich cerium oxide for efficient photocatalytic CO2 reduction to methanol Surfaces and
36 Interfaces 34 102389.
37 [24] Gao Z, Zhang D, Li X, Jiang S and Zhang Q 2018 Current status, opportunities
38
39 and challenges in chemical conversion coatings for zinc Colloids and Surfaces A 546
40 221.
41 [25] Zhang J, Zhang Z, Jiao Y, Yang H, Li Y, Zhang J and Gao P 2019 The
42
43 graphene/lanthanum oxide nanocomposites as electrode materials of supercapacitors
44 Journal of Power Sources 419 99.
45
46
[26] Sakthi T, Nataraj N, Chen T, Chen S and Kokulnathan T 2022 Synergistic
47 formation of samarium oxide/graphene nanocomposite: A functional electrocatalyst for
48 carbendazim detection Chemosphere 307 135711.
49
50 [27] Yu Y, Wang X, Gao W, Li P, Yan W, Wu S, Cui Q, Song W and Ding K 2017
51 Trivalent cerium-preponderant CeO2/graphene sandwich-structured nanocomposite
52 with greatly enhanced catalytic activity for the oxygen reduction reaction Journal of
53
54 Materials Chemistry A 5 6656.
55 [28] Qian X, Jin L, Zhu L, Yao S, Rao D, Shen X, Xi X, Xiao K and Qin S 2016 CeO2
56
57
nanodots decorated ketjen black for high performance lithium–sulfur batteries RSC
58 Advances 6 111190.
59 [29] Bakry A, Alamier W, Salama R, El-Shall M and Awad F 2022 Remediation of
60
61
62
63
64
65
water containing phosphate using ceria nanoparticles decorated partially reduced
1
graphene oxide (CeO2-PRGO) composite Surfaces and Interfaces 31 102006.
2
3 [30] Min C, He Z, Liu D, Jia W, Qian J, Jin Y and Li S 2019 Ceria/reduced Graphene
4 Oxide Nanocomposite: Synthesis, Characterization, and Its Lubrication Application
5
6 ChemistrySelect 4 4615.
7 [31] Ferrari A and Basko D 2013 Raman spectroscopy as a versatile tool for studying
8 the properties of graphene Nature Nanotechnology 8 235.
9
10 [32] Ferrari A, Meyer J, Scardaci V, Casiraghi C, Lazzeri M, Mauri F, Piscanec S, Jiang
11 D, Novoselov K, Roth S and Geim A 2006 Raman spectrum of graphene and graphene
12
13
layers Physical Review Letters 97 187401.
14 [33] Dong L, Chen W, Deng N and Zheng C 2016 A novel fabrication of graphene by
15 chemical reaction with a green reductant Chemical Engineering Journal 306 754.
16
17 [34] Verma R and Samdarshi S 2016 In Situ Decorated Optimized CeO2 on Reduced
18 Graphene Oxide with Enhanced Adsorptivity and Visible Light Photocatalytic Stability
19 and Reusability The Journal of Physical Chemistry C 120 22281.
20
21 [35] Murugan R, Vijayaprasath G and Ravi G 2015 The influence of substrate
22 temperature on the optical and micro structural properties of cerium oxide thin films
23
24
deposited by RF sputtering Superlattices and Microstructures 85 321.
25 [36] Deosarkar M, Pawar S and Bhanvase B 2014 In-situ sonochemical synthesis of
26 Fe3O4-graphene nanocomposite for lithium rechargeable batteries Chemical
27
28 Engineering and Processing 83 49.
29 [37] Gu M, Farooq U, Lu S, Zhang X, Qiu Z and Sui Q 2018 Degradation of
30 trichloroethylene in aqueous solution by rGO supported nZVI catalyst under several
31
32 oxic environments Journal of Hazardous Materials 349 35.
33 [38] Abbasi A, Amin K, Ali M, Ali Z, Atif M, Ensinger W and Khalid W 2022
34
35
Synergetic effect of adsorption-photocatalysis by GO-CeO2 nanocomposites for
36 photodegradation of doxorubicin Journal of Environmental Chemical Engineering 10
37 107078.
38
39 [39] Stankovich S, Dikin D, Piner R, Kohlhaas K, Kleinhammes A, Jia Y, Wu Y,
40 Nguyen S and Ruoff R 2007 Synthesis of graphene-based nanosheets via chemical
41 reduction of exfoliated graphite oxide Carbon 45 1558.
42
43 [40] Beche E, Charvin P, Perarnau D, Abanades S and Flamant G 2008 Ce 3d XPS
44 investigation of cerium oxides and mixed cerium oxide (CexTiyOz) Surface and
45
46
Interface Analysis 40 264.
47 [41] Yuan S, Zhang Q, Xu B, Jin Z, Zhang Y, Yang Y, Zhang M and Ohno T 2014
48 Porous cerium dioxide hollow spheres and their photocatalytic performance RSC
49
50 Advances 4 62255.
51 [42] Dreyer D, Park S, Bielawski C and Ruoff R 2010 The chemistry of graphene oxide
52 Chemical Society Reviews 39 228.
53
54 [43] Basak M, Rahman M, Ahmed M, Biswas B and Sharmin N 2022 The use of X-ray
55 diffraction peak profile analysis to determine the structural parameters of cobalt ferrite
56
57
nanoparticles using DebyeScherrer, Williamson-Hall, Halder-Wagner and Size-strain
58 plot: Different precipitating agent approach Journal of Alloys and Compounds 895
59 162694.
60
61
62
63
64
65
[44] Chen W, Feng Y, Wan Y, Zhang L, Yang D, Gao X, Yu Q and Wang D 2023
1
Investigation on anti-wear and corrosion-resistance behavior of steel-steel friction pair
2
3 enhanced by ionic liquid additives under conductive conditions Tribology International
4 177 108002.
5
6 [45] Kourouklis G, Jayaraman A and Espinosa G 1988 High-pressure Raman study of
7 CeO2 to 35 GPa and pressure-induced phase transformation from the fluorite structure
8 Physical Review B: Condens Matter 37 4250.
9
10 [46] Mi W, Guo Z, Wang Q, Yang Y and Bai H 2013 Charge ordering in reactive
11 sputtered (100) and (111) oriented epitaxial Fe3O4 films Scripta Materialia 68 972.
12
13
[47] Qian Y, Zheng W, Chen W, Feng T, Liu T and Fu Y 2021 Enhanced functional
14 properties of CeO2 modified graphene/epoxy nanocomposite coating through interface
15 engineering Surface and Coatings Technology 409 126819.
16
17 [48] Xiang T, Ding S, Li C, Zheng S, Hu W, Wang J and Liu P 2017 Effect of current
18 density on wettability and corrosion resistance of superhydrophobic nickel coating
19 deposited on low carbon steel Materials & Design 114 65.
20
21 [49] Lotfi M, Yari H, Sari M and Azizi A 2022 Fabrication of a highly hard yet tough
22 epoxy nanocomposite coating by incorporating graphene oxide nanosheets dually
23
24
modified with amino silane coupling agent and hyperbranched polyester-amide
25 Progress in Organic Coatings 162 106570.
26 [50] Li L, Chen D, Long Y, Wang F and Kang Z 2023 Silane modification of semi-
27
28 curing epoxy surface: High interfacial adhesion for conductive coatings Progress in
29 Organic Coatings 174 107228.
30 [51] Cai G, Hu P, Cao X, Chen J, Zhang X and Dong Z 2023 pH-triggered self-
31
32 inhibition epoxy coating based on cerium-polyphenolic network wrapped carbon
33 nanotube Progress in Organic Coatings 175 107355.
34
35
[52] Lei Y, Zhang X, Liu Q, Tao Y, Sun Y and You B 2022 Skin-mimetic assembly
36 strategy for fabricating a transparent and highly anti-corrosive FSO-GO/epoxy
37 nanocomposite coating Progress in Organic Coatings 173 107184.
38
39 [53] Alipanaha N, Yaria H, Mahdaviana M, Ramezanzadeha B and Bahlakeh G 2021
40 MIL-88A (Fe) filler with duplicate corrosion inhibitive/barrier effect for epoxy coatings:
41 Electrochemical, molecular simulation, and cathodic delamination studies Journal of
42
43 Industrial and Engineering Chemistry 97 200.
44 [54] Xiao Z, Wu X, Tan H and Hao S 2023 Design and synthesis of Fe-Ce-O@C with
45
46
efficient photocatalytic activity Journal of Rare Earths 41 91.
47 [53] Yin W and Cao H 2016 One-step synthesis of SnO2-reduced graphene oxide (SOG)
48 composites for efficient removal of organic dyes from wastewater RSC Advances 6
49
50 100636.
51 [54] Choi J, Reddy D, Islam M, Ma R and Kim T 2016 Self-assembly of CeO2
52 nanostructures/reduced graphene oxide composite aerogels for efficient photocatalytic
53
54 degradation of organic pollutants in water Journal of Alloys and Compounds 688 527.
55 [55] Song H, Wu R, Yang J, Dong J and Ji G 2018 Fabrication of CeO2 nanoparticles
56
57
decorated three-dimensional flower-like BiOI composites to build p-n heterojunction
58 with highly enhanced visible-light photocatalytic performance J Colloid Interface Sci
59 512 325.
60
61
62
63
64
65
[56] Munawar T, Mukhtar F, Nadeem M, Riaz M, Naveed M, Mahmood K, Hasan M,
1
Arshad M, Hussain F, Hussain A and Iqbal F 2020 Novel photocatalyst and antibacterial
2
3 agent; direct dual Z-scheme ZnO–CeO2-Yb2O3 heterostructured nanocomposite Solid
4 State Sciences 109 106446.
5
6 [57] Maria C, Kaviyarasu K, Judith J, Siddhardha B and Jeyaraj B 2017 Facile synthesis
7 of heterostructured cerium oxide/yttrium oxide nanocomposite in UV light induced
8 photocatalytic degradation and catalytic reduction: Synergistic effect of antimicrobial
9
10 studies Journal of Photochemistry and Photobiology B: Biology 173 23.
11 [58] Liyanage A, Perera S, Tan K Chabal Y and Balkus K 2014 Synthesis,
12
13
Characterization, and Photocatalytic Activity of Y-Doped CeO2 Nanorods ACS
14 Catalysis 4 577.
15 [59] Yao J, Gao Z, Meng Q, He G and Chen H 2021 One-step synthesis of reduced
16
17 graphene oxide based ceric dioxide modified with cadmium sulfide (CeO2/CdS/RGO)
18 heterojunction with enhanced sunlight-driven photocatalytic activity Journal of Colloid
19 and Interface Science 594 621.
20
21 [60] Djellabi R, Ghorab F, Nouacer S, Smara A and Khireddine O 2016 Cr(VI)
22 photocatalytic reduction under sunlight followed by Cr(III) extraction from TiO2
23
24
surface Materials Letters 176 106.
25 [61] Dai W, Kheireddin B, Gao H and Liang H 2016 Roles of nanoparticles in oil
26 lubrication Tribology International 102 88.
27
28 [62] Shen T, Wang D, Yun J, Liu Q, Liu X and Peng Z 2016 Tribological properties and
29 tribochemical analysis of nano-cerium oxide and sulfurized isobutene in titanium
30 complex grease Tribology International 93 332.
31
32 [63] Englert J, Dotzer C, Yang G, Schmid M, Papp C, Gottfried J, Steinruck H, Spiecker
33 E, Hauke F and Hirsch A 2011 Covalent bulk functionalization of graphene Nature
34
35
Chemistry 3 279.
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
Response to Reviewers

To: Editor of Materials Chemistry and Physics

Sub: Revised manuscript of MATCHEMPHYS-D-22-05852.

Dear Editor:

We sincerely thank you for your help in processing our paper and all the reviewers for
their professional and helpful comments and suggestions about the manuscript entitled
“CeO2 modified reduced graphene oxides nanocomposites for tribology, anti-corrosion
and photocatalysis” (MATCHEMPHYS-D-22-05852). We have studied these
comments carefully and have made major revisions. We hope that the revised paper
now meets the high standard of the journal. The following are the lists of major
corrections of the paper and our point-by-point responses to all the review comments.

Many thanks and best regards.

Wenge Chen
On behalf of all co-authors

Detailed replies to comments

Replies to the comments from Reviewer #1:

The authors prepared CeO2@rGO nanocomposite by hydrothermal method and


exploited its three different properties (i) as additive into PAO5W-40 lubricating oil for
improving tribological properties, (ii) as chromium-free Dacromat coating for
enhancing anti-corrosion properties of coating, and (iii) as a photocatalyst for degrading
the organic dye RhB in aqueous solution. In my view the authors could have thoroughly
studies any one of the properties of the prepared CeO2@rGO nanocomposite. The
quality of the work is not up to the mark with three different properties of CeO2@rGO
nanocomposite; the authors need to improve its quality with the following observations:

1. Other characterization techniques should be included for CeO2@rGO


nanocomposite, such as FT-IR and XPS.

Reply:
Thank you for your valuable comments.
We have performed these measurements and now provided the FT-IR and XPS
spectra for the composite. The corresponding context has been changed as following:
“Fig. 3(a) shows the FT-IR spectra of the GO, rGO and CeO2@rGO composite. All
the typical characteristic peaks of rGO are visible, and the peak at 3400 cm -1 is due to
the stretching vibration mode of the O-H on the GO surface. The peak at 1386 cm-1 is
attributed to the in-plane bending vibration mode of the C-OH on the rGO surface. The
peaks at 1720 cm-1 and 1650 cm-1 are due to the stretching vibration modes of the C=O
and C=C on the carboxyl group, respectively, the peaks at 1230 cm-1 and 1053 cm-1 are
due to the stretching vibration modes of the C-O-C [36, 37]. In comparison to that of
rGO, the CeO2@rGO exhibits a new Ce-O peak at 557 cm-1, which implies that the rare
earth oxide CeO2 has been successfully grafted onto the surface of rGO, providing an
effective foundation for its usage as a lubricant additive [38]. Comparing the FT-IR
spectra of GO and rGO, it can be observed that the peak intensity of the C=O bond
stretching vibration mode at around 1720 cm-1 is much stronger in rGO than that in GO.
Furthermore, it can be observed that the peak width of the epoxy group in the GO is
much wider than that in the rGO at the same position, which is likely due to the partial
reduction of the product during the reaction process at various reaction temperature and
reagents.”
“Fig. 3(b) shows the XPS survey spectrum of the CeO2@rGO composite. The
major peaks in the survey spectrum are attributed to C 1s, O 1s, Ce 3d and Ce 4d,
indicating that the main components of the composite are carbon, oxygen and cerium.
Fig. 3(c) shows the high-resolution C 1s spectrum of CeO2@rGO, which exhibits a
typical C 1s peak for the rGO. The curve fitting results reveal three characteristic peaks
at 284.8 eV, 285.9 eV, and 287.8 eV, corresponding to C-C, C-O, and C=O bonds,
respectively [39]. Fig. 3(d) shows the high-resolution Ce 3d spectrum of the
CeO2@rGO composite, which exhibits a typical Ce 3d peak consisting of a single peak
and two multiplet peaks, indicating the typical Ce 3d peak of CeO2. The Ce 3d peak of
CeO2 can be decomposed into ten peaks through Gaussian-Lorentz function fitting,
corresponding to v0 (880.8 eV), v (882.8 eV), v' (885.2 eV), v'' (888.5 eV), v''' (898.6
eV), u0 (899.9 eV), u (901.2 eV), u' (903.6 eV), u'' (906.5 eV), and u''' (916.9 eV).
Based on literature, those peaks of v0, v', u0, and u' are corresponding to 3d5/2 and 3d3/2,
which are the signature states of Ce(III) 3d. Whereas those peaks of v, v'', v''', u, u'', and
u''' are corresponding to 3d5/2 and 3d3/2, which are the signature states of Ce(IV) 3d [40].
These results clearly show that the CeO2 of the composite is mainly in the form of
Ce(IV), but there is also a certain amount of Ce(III). This may be due to the reduction
of coordination numbers of Ce caused by oxygen vacancies in the CeO2 lattice, or it is
because Ce3+ is not fully oxidized during thermal treatments due to insufficient O
sources on the rGO [41].”
Fig. 3. (a) FT-IR spectra of GO, rGO and CeO2@rGO; (b) XPS survey spectra of CeO2@rGO; (c)
high resolution C1s spectrum of CeO2@rGO; (d) high resolution Ce3d spectrum of CeO2@rGO.

2. The authors are not thoroughly investigated the anti-corrosion property of the
synthesized nanocomposite. The authors need to include open circuit potential values,
corrosion current, corrosion potential and charge transfer resistance of the blank and
different wt% of the CeO2@rGO nanocomposite dispersed coating material. Further,
more discussion related to the anti-corrosion mechanism need to be included.

Reply:
We thank you for your professional comments.
Accordingly, we have added the results for the coatings with the 0.09% and 0.27%
CeO2@rGO composites as following:
“Fig. 7(a) shows the Tafel curves of the coatings with different contents of
CeO2@rGO. Except for the coating with 0.27% CeO2@rGO, the corrosion potentials
of all the other coatings are lower than that of the substrate. The potential difference is
between 0.0504 V to 0.1092 V. All these results indicate that the coatings added with
CeO2@rGO can provide a certain degree of cathodic protection for the substrate [47].
Comparing the coatings added with different contents of CeO2@rGO, the corrosion
potential of the coating with 0.18% CeO2@rGO shows the most negative value, and the
corrosion potential of the coating with 0.27% CeO2@rGO shows the most positive
value. In addition, the corrosion current densities of the coatings with 0.09% and 0.18%
CeO2@rGO are decreased by an order of magnitude if compared to that of the coating
without adding the CeO2@rGO. The corrosion resistance was significantly increased
after adding the CeO2@rGO, indicating that a certain amount of CeO2@rGO can reduce
the Icorr value of the coating and improve its corrosion resistance. However, the Icorr
value of the coating with 0.27% CeO2@rGO is increased. This is because after adding
an excessive amount of CeO2@rGO, the rGO could become contacted with each other
and can form an effective corrosion path, thus increasing the conductivity of the coating
and the corrosion current density. It is also possible that adding too much amount of
composite causes severe agglomeration, producing more pores for the corrosion
medium easier to pass through, thus reducing the corrosion resistance of the coating.”

Fig. 7. (a) Tafel polarization curves of chromium-free Dacromet coatings reinforced with different
contents of CeO2@rGO; (b) Nyquist diagram of chromium-free Dacromet coatings reinforced
with different contents of CeO2@rGO; Bode diagram of chromium-free Dacromet coatings
reinforced with different contents of CeO2@rGO: (c) phase frequency diagram; (d) amplitude-
frequency diagram.
“Fig. 7(b) shows the Nyquist diagram of the coatings with different contents of
CeO2@rGO in a 3.5% NaCl solution. It can be seen that the capacitive arc radii of the
coatings with 0.09% and 0.18% CeO2@rGO are much larger than that of the coating
with 0% CeO2@rGO, and the density of the coating is increased after adding a certain
amount of composite, both of which can effectively improve the corrosion resistance
of the coating [48]. This is because the graphene is uniformly dispersed in the coating.
It can effectively fill up the defects in the coating, which plays a positive role as a
physical barrier to hinder the intrusion of the corrosive medium. When the CeO2@rGO
content is 0.27%, the capacitive arc radius of the coating decreases, indicating that
adding an excessive amount of composite will reduce its dispersion in the coating,
leading to agglomeration, increased defects in the coating, and also decreased corrosion
resistance of the coating.”
“Figs. 7(c-d) show the Bode diagrams of coatings added with different amounts of
the composite. From the frequency response diagram, the |Z| values of the coatings with
0.09% and 0.18% CeO2@rGO are higher than those of the coatings without adding the
composite. Among them, the coating with 0.18% CeO2@rGO exhibits the maximum
low-frequency impedance value (|Z|f→0) of about 3.68×103 Ω·cm2, indicating its
minimum corrosion rate and relatively good corrosion resistance. When the added
amount of CeO2@rGO reaches 0.27%, the |Z| value of the coating becomes decreased,
and its corrosion resistance is also decreased. This again proves that adding a certain
amount of composite enhances the corrosion resistance of the coating, but adding
excessive amounts of composite will reduce the corrosion resistance of the coating.
This may be due to the van der Waals force which causes the rGO to be agglomerated
locally, thus increasing the pores in the coating and decreasing its shielding effect.”
“Fig. 8 shows the equivalent circuit model of the coating’s electrochemical
impedance spectrum, which is Rs(Cc(Rc(QfRf)))(CpRp). Here, Rs, Rc, Rf, and Rp are the
resistances of the solution, coating, surface film layer of zinc-aluminum metal powder,
and surface activation dissolution reaction of zinc-aluminum metal powder,
respectively. Cc, Qf, and Cp are the capacitances of the coating, surface film layer of
zinc-aluminum alloy powder, and surface activation dissolution reaction of zinc-
aluminum metal powder, respectively. Table 1 summarizes the fitting results of the
electrochemical impedance spectra of all the coatings. It can be seen that the Rc values
increase with the increase of the amount of composites. The Rc values of coatings with
0%, 0.09%, and 0.18% CeO2@rGO are relatively small, around 25 Ω·cm2, indicating
that the metallic powder is not completely covered by the passivation film, and cannot
provide effective cathodic protection to the substrate. In Table 1, Rf represents the
resistance of the corrosion product film formed on the surface of the metallic powder.
The larger the value of Rf, the denser the corrosion product layer formed on the surface
of the metallic powder, and the better its protection effect. This will make it more
difficult for the metallic powder to undergo activation dissolution, thus enhancing the
anti-corrosion resistance of the coating. In addition, the value of Rf increases firstly but
then decreases with the further increase of CeO2@rGO. Among them, the coating with
0.18% CeO2@rGO exhibits the largest Rf value of 4064 Ω·cm2, indicating that the
corrosion product film formed by this coating is the densest, thus effectively forming a
physical barrier and delaying the corrosion of the metallic powder. Rp represents the
resistance of the activation dissolution reaction of the metallic powder. The larger the
value of Rp, the more difficult the reaction of the zinc-aluminum powder can occur, and
the better the corrosion resistance of the coating. According to the Table 1, Rp value is
increased with increasing the amounts of CeO2@rGO, and the coating added with 0.27%
CeO2@rGO exhibits the largest Rp. Since both the Rf and Rp values are critical to the
corrosion resistance of the coating, the sum of Rf and Rp is used in this study to represent
the corrosion resistance of the coating, with a larger Rf+Rp indicating a better corrosion
resistance. As shown in the Table 1, the coating with 0.18% CeO2@rGO exhibits the
largest Rf+Rp value, indicating its best corrosion resistance.”

Fig. 8. Equivalent circuit diagram.


Table 1. Fitting parameters of electrochemical impedance spectrum for chromium-free Dacormet
coatings reinforced with different contents of CeO2@rGO.
wt.% Rs Cc Rc CPEf n Rf Cp Rp
Ω·cm2 F/cm2 Ω·cm2 S-n/cm2 Ω·cm2 F/cm2 Ω·cm2
0% 14.92 7.353×10-7 24.47 3.203×10-4 0.7175 2712 4.212×10-8 64.16
-7 -4
0.09% 13.44 6.706×10 25.47 2.784×10 0.6558 3985 3.994×10-8 66.31
-7 -4
0.18% 11.45 1.321×10 26.51 3.145×10 0.6434 4064 4.459×10-8 74
0.27% 12.74 6.024×10-8 54 5.357×10-4 0.5357 3081 8.404×10-8 608.1
We have also added the FT-IR spectrum of the coatings with CeO2@rGO:
“Fig. 9 shows the FT-IR spectrum of the Q235 steel substrate with the chromium-
free Dacromet coating with the CeO2@rGO. It can be seen that the coating exhibits the
characteristic peaks corresponding to those of O-Ce, C-O-Si and Si-O-Si bonds [49,
50]. The hydrolyzed silane alcohol produced by the silane coupling agent in the coating
can form C-O-Si bond, and at the same time, the silane alcohol itself becomes
dehydrated and condensed to form linear or cyclic Si-O-Si network structures. The
Ce(OH)3 passivation layer is formed at the interface between the substrate and coating,
which prevents direct contact between the substrate and the external environment [51].
The formation of Si-O-Fe and Ce-Fe bonds at the interface between the coating and
substrate is also beneficial for improving the adhesion between the coating and
substrate, thereby improving its corrosion resistance [32, 52, 53]. Formation of these
bonds can also improve the density of the coating and enhance the bonding strength
between the coating and substrate.”

Fig. 9. FT-IR spectrum of the coatings with CeO2@rGO.


We have conducted more corrosion tests and added the results. The corrosion
resistance mechanism has also been added:
“In this study of the chromium-free Dacromet coating, we used the molybdic acid
instead of the traditionally used chromic acid. The corrosion resistance mechanism of
the coating has also been changed to the formation of insoluble molybdate on the
substrate as shown in Eq. 3. The formed oxide film on the substrate’s surface passivates
it from corrosion, which competes with Cl- during corrosion, thus reducing the
corrosion of Cl-. In addition, the corrosion resistance performance of the coating is
further improved after the addition of CeO2@rGO. The modified rGO is uniformly
dispersed in the coating, and its layered structure easily adheres to the zinc-aluminum
alloy powder and plays a key role in the solidification of the CeO2@rGO. On the other
hand, the nanoflakes and nanoparticles fill the porous structure of the coating, greatly
increasing its density and hardness. The homogeneous distribution of the CeO2@rGO
in the coating promotes the formation of a conductive channel, facilitates the electron
transfer, and is thus beneficial for sacrificing zinc-aluminum alloy with a lower
potential for protecting the substrate. The hydrolysis of the silane coupling agent in the
coating produces a silanol that can form C-O-Si, and Si-O-Fe is also formed at the
interface between the coating and the substrate [49]. At the same time, the self-
dehydration and condensation of silanol form a linear or cyclic Si-O-Si network
structure. The Ce(OH)3 passivation layer is formed at the interface between the
substrate and the coating, which actively prevents direct contact between the substrate
and the external environment. This indicates that the compound in the anti-corrosion
film has barrier/active dual-functions. The presence of Ce-Fe bonds can also improve
the adhesion between the coating and substrate, thereby improving the wear resistance
and corrosion resistance of the coating [52]. Therefore, the corrosion resistance of the
chromium-free Dacromet coating with the added CeO2@rGO has been improved.
MoO42- + Fe2+ → FeMoO4 → Fe2(MoO4)3 (3)”

3. Why the authors have not reported the photodegradation efficiency of CeO2 alone
for referencing? This need to be included for comparison to the photodegradation
efficiency of CeO2@rGO nanocomposite.

Reply:
Thanks for this comment.
We added the photocatalytic performance of CeO2, GO and CeO2@rGO to provide
relevant comparisons:
“Fig. 10(a) shows the photodegradation curves of the organic dye RhB by using
CeO2, GO and CeO2@rGO in the dark and under UV irradiation. It can be seen that the
removal of RhB by GO mainly occurs in the adsorption process under dark conditions,
and the removal rate under the UV irradiation is not decreased significantly, with the
maximum removal rate of 53.2%. CeO2 also mainly removes the RhB through the
adsorption process under dark conditions. However, under the UV irradiation, it
exhibits photocatalytic activity, with the maximum removal rate of 20.3%. CeO2@rGO
achieves a high removal rate of 90.5% for RhB, of which 77.8% is removed through
the adsorption process under dark conditions. The subsequent photocatalytic
degradation process further increases the removal rate up to 90.5%. Fig. 10(b) shows
the UV-vis spectra before and after the RhB adsorption. This shows that the RhB has a
maximum absorption peak at 554 nm before adsorption, but this peak is gradually
weakened after the RhB removal. The reason for this phenomenon may be that the rGO
nanosheets are excellent electron acceptors, while the UV-excited CeO2 is an excellent
electron donor. Therefore, the interactions between CeO2 and rGO nanosheets can cause
excited electrons to be effectively transferred from CeO2 to rGO nanosheets, thus
reducing the possible recombination of electron-hole pairs. Modifications of covalent
bonding between rGO and CeO2 can also improve the dispersion of rGO in the solution.
Furthermore, the layered structure of rGO can prevent the severe aggregation of CeO2
nanoparticles, thereby exposing more active sites to enhance the photocatalytic
performance of the CeO2@rGO composite.”
Fig. 10. (a) Photodegradation curves of different concentrations for RhB by the CeO2, GO and
CeO2@rGO under UV irradiation; (b) UV-vis spectra of RhB before and after photocatalytic
degradation; (c) Concentration changes of RhB dye solutions with time for CeO2@rGO in the
dark and under UV irradiation; (d) The pseudo-first-order reaction kinetics of the CeO2@rGO
composite for RhB degradation; (e) Comparisons of the removal efficiency for RhB dye of
different CeO2 based catalyst; (f) Recycling performance of the CeO2@rGO during adsorption-
photodegradation of the RhB solution.

Replies to the comments from Reviewer #2:

Reviewer #2:
This manuscript report lubricating properties, corrosion resistance properties and
photocatalytic properties of a CeO2@rGO nanocomposite. A chemical bonding
mechanism is proposed to contribute to the anticorrosion properties. However, this
statement is presented without any evidence of chemical bonding. Polarization and
impedance analyses should be revised as well. I consider this manuscript is incomplete
and unready for publication.
1. Authors claimed in the abstract that C-O-Si, Si-O-Me, and Si-O-Si bonds are formed
in the modified coating. However, no evidence is provided in the result section for this
statement. Chemical characterization of the CeO2@rGO-modified dacromet coating
should be provided, required for the proposed anticorrosion hypothesis in Figure 1b3,
in order to observe which chemical bonds are formed.

Reply:
Thank you for your comments.
We have added the FT-IR spectrum of the coatings with CeO2@rGO as following:
“Fig. 9 shows the FT-IR spectrum of the Q235 steel substrate with the chromium-
free Dacromet coating with the CeO2@rGO. It can be seen that the coating exhibits the
characteristic peaks corresponding to those of O-Ce, C-O-Si and Si-O-Si bonds [49,
50]. The hydrolyzed silane alcohol produced by the silane coupling agent in the coating
can form C-O-Si bond, and at the same time, the silane alcohol itself becomes
dehydrated and condensed to form linear or cyclic Si-O-Si network structures. The
Ce(OH)3 passivation layer is formed at the interface between the substrate and coating,
which prevents direct contact between the substrate and the external environment [51].
The formation of Si-O-Fe and Ce-Fe bonds at the interface between the coating and
substrate is also beneficial for improving the adhesion between the coating and
substrate, thereby improving its corrosion resistance [32, 52, 53]. Formation of these
bonds can also improve the density of the coating and enhance the bonding strength
between the coating and substrate.”
Fig. 9. FT-IR spectrum of the coatings with CeO2@rGO.
We have conducted in-depth exploration and modification of the corrosion
resistance mechanism:
“In this study of the chromium-free Dacromet coating, we used the molybdic acid
instead of the traditionally used chromic acid. The corrosion resistance mechanism of
the coating has also been changed to the formation of insoluble molybdate on the
substrate as shown in Eq. 3. The formed oxide film on the substrate’s surface passivates
it from corrosion, which competes with Cl- during corrosion, thus reducing the
corrosion of Cl-. In addition, the corrosion resistance performance of the coating is
further improved after the addition of CeO2@rGO. The modified rGO is uniformly
dispersed in the coating, and its layered structure easily adheres to the zinc-aluminum
alloy powder and plays a key role in the solidification of the CeO2@rGO. On the other
hand, the nanoflakes and nanoparticles fill the porous structure of the coating, greatly
increasing its density and hardness. The homogeneous distribution of the CeO2@rGO
in the coating promotes the formation of a conductive channel, facilitates the electron
transfer, and is thus beneficial for sacrificing zinc-aluminum alloy with a lower
potential for protecting the substrate. The hydrolysis of the silane coupling agent in the
coating produces a silanol that can form C-O-Si, and Si-O-Fe is also formed at the
interface between the coating and the substrate [49]. At the same time, the self-
dehydration and condensation of silanol form a linear or cyclic Si-O-Si network
structure. The Ce(OH)3 passivation layer is formed at the interface between the
substrate and the coating, which actively prevents direct contact between the substrate
and the external environment. This indicates that the compound in the anti-corrosion
film has barrier/active dual-functions. The presence of Ce-Fe bonds can also improve
the adhesion between the coating and substrate, thereby improving the wear resistance
and corrosion resistance of the coating [52]. Therefore, the corrosion resistance of the
chromium-free Dacromet coating with the added CeO2@rGO has been improved.
MoO42- + Fe2+ → FeMoO4 → Fe2(MoO4)3 (3)”

2. Authors attributed the corrosion protection of CeO2@rGO-modified Dacromet


coating to the formation of C-O-Si, Si-O-Me and Si-O-Si bonds in the modified coating.
However, Si-O-Me and Si-O-Si bonds do not originate from the addition of
CeO2@rGO. Please explain.

Reply:
Thank you for your comments.
The addition of CeO2@rGO improves the corrosion resistance of the coating due
to the following three reasons. Firstly, uniform dispersion of CeO2@rGO in the
Dacromet coating helps to improve its adhesion to zinc-aluminum alloy powders, which
plays a crucial role in the solidification of CeO2@rGO. Secondly, the filling of pores in
the coatings using the nanometer-sized flakes and particles significantly increases the
density and hardness of the coating. The strong bonding between the composite and
substrate forms a good passivation layer to improve the corrosion resistance of the
substrate and also improve its adhesion to the coating. Finally, the uniform distribution
of CeO2@rGO in the coating easily form the conductive channels, facilitates electron
transmission, and is beneficial to etch zinc-aluminum alloys at a much lower potential,
thus achieving the goal of protecting the substrate.
The corrosion resistance of the modified Dacromet coating is also attributed to the
formation of C-O-Si, Si-O-Fe and Si-O-Si bonds in the modified coating [1, 2]. The C-
O-Si bond is formed due to the hydrolysis of silane coupling agent (one of the main
components in the coating) with the hydroxyl and carboxyl groups of rGO. The Si-O-
Fe bond is formed due to the reaction between the hydroxyl and carboxyl groups of
rGO and the hydroxyl groups on the substrate surface. In addition, Si-O-Si network
structures are formed due to the self-dehydration condensation of silanol. Formation of
these strong bonds improves the density of the coating and enhances the bonding
strength between the coating and the substrate. The schematic diagram of the interface
bonding between the coating and substrate is shown in Fig. R1. However, these bonding
processes do not involve with CeO2, and the main function of CeO2 is to prevent the
severe agglomeration of rGO, thus uniformly dispersing it in substrate and maximizing
its corrosion resistance.

Fig. R1. Schematic diagram of bonding interface of chromium-free Dacromet coating containing
CeO2@rGO.

[1] Lotfi M, Yari H, Sari M and Azizi A 2022 Fabrication of a highly hard yet tough
epoxy nanocomposite coating by incorporating graphene oxide nanosheets dually
modified with amino silane coupling agent and hyperbranched polyester-amide
Progress in Organic Coatings 162 106570.
[2] Li L, Chen D, Long Y, Wang F and Kang Z 2023 Silane modification of semi-curing
epoxy surface: High interfacial adhesion for conductive coatings Progress in Organic
Coatings 174 107228.

3. Polarization studies related to Tafel curves in Figure 7a are not explained in the
manuscript, only mentioned in the abstract. Please add its discussion part.

Reply:
We sincerely thank for this suggestion.
We added the Tafel curves of the coatings with different contents of the composites
in the modified paper as following:
“Fig. 7(a) shows the Tafel curves of the coatings with different contents of
CeO2@rGO. Except for the coating with 0.27% CeO2@rGO, the corrosion potentials
of all the other coatings are lower than that of the substrate. The potential difference is
between 0.0504 V to 0.1092 V. All these results indicate that the coatings added with
CeO2@rGO can provide a certain degree of cathodic protection for the substrate [47].
Comparing the coatings added with different contents of CeO2@rGO, the corrosion
potential of the coating with 0.18% CeO2@rGO shows the most negative value, and the
corrosion potential of the coating with 0.27% CeO2@rGO shows the most positive
value. In addition, the corrosion current densities of the coatings with 0.09% and 0.18%
CeO2@rGO are decreased by an order of magnitude if compared to that of the coating
without adding the CeO2@rGO. The corrosion resistance was significantly increased
after adding the CeO2@rGO, indicating that a certain amount of CeO2@rGO can reduce
the Icorr value of the coating and improve its corrosion resistance. However, the Icorr
value of the coating with 0.27% CeO2@rGO is increased. This is because after adding
an excessive amount of CeO2@rGO, the rGO could become contacted with each other
and can form an effective corrosion path, thus increasing the conductivity of the coating
and the corrosion current density. It is also possible that adding too much amount of
composite causes severe agglomeration, producing more pores for the corrosion
medium easier to pass through, thus reducing the corrosion resistance of the coating.”
Fig. 7. (a) Tafel polarization curves of chromium-free Dacromet coatings reinforced with different
contents of CeO2@rGO.

4. Discussion of Nyquist diagrams is repeated twice in the text. The first paragraph
should be replaced by explanation of Tafel polarization curves in Figure 7a, which is
currently missing.

Reply:
Thanks for this good comment, and we have done the modifications.
We modified the Tafel curves of the coatings and the Nyquist diagram of the
coatings with 0.09% and 0.27% CeO2@rGO as following:
“Fig. 7(b) shows the Nyquist diagram of the coatings with different contents of
CeO2@rGO in a 3.5% NaCl solution. It can be seen that the capacitive arc radii of the
coatings with 0.09% and 0.18% CeO2@rGO are much larger than that of the coating
with 0% CeO2@rGO, and the density of the coating is increased after adding a certain
amount of composite, both of which can effectively improve the corrosion resistance
of the coating [48]. This is because the graphene is uniformly dispersed in the coating.
It can effectively fill up the defects in the coating, which plays a positive role as a
physical barrier to hinder the intrusion of the corrosive medium. When the CeO2@rGO
content is 0.27%, the capacitive arc radius of the coating decreases, indicating that
adding an excessive amount of composite will reduce its dispersion in the coating,
leading to agglomeration, increased defects in the coating, and also decreased corrosion
resistance of the coating.”
“Figs. 7(c-d) show the Bode diagrams of coatings added with different amounts of
the composite. From the frequency response diagram, the |Z| values of the coatings with
0.09% and 0.18% CeO2@rGO are higher than those of the coatings without adding the
composite. Among them, the coating with 0.18% CeO2@rGO exhibits the maximum
low-frequency impedance value (|Z|f→0) of about 3.68×103 Ω·cm2, indicating its
minimum corrosion rate and relatively good corrosion resistance. When the added
amount of CeO2@rGO reaches 0.27%, the |Z| value of the coating becomes decreased,
and its corrosion resistance is also decreased. This again proves that adding a certain
amount of composite enhances the corrosion resistance of the coating, but adding
excessive amounts of composite will reduce the corrosion resistance of the coating.
This may be due to the van der Waals force which causes the rGO to be agglomerated
locally, thus increasing the pores in the coating and decreasing its shielding effect.”

Fig. 7. (a) Tafel polarization curves of chromium-free Dacromet coatings reinforced with different
contents of CeO2@rGO; (b) Nyquist diagram of chromium-free Dacromet coatings reinforced
with different contents of CeO2@rGO; Bode diagram of chromium-free Dacromet coatings
reinforced with different contents of CeO2@rGO: (c) phase frequency diagram; (d) amplitude-
frequency diagram.

5. Fitting of the equivalent electric circuit pictured in Figure 7b is absent, in which


values for each component of the electric circuit are not calculated. Circuit elements
(Rs, Cc, Rc, CPEf, Rf, Cp, Rp) are neither explained nor assigned to values.

Reply:
Thank you for your valuable comments.
We added the fitting of the equivalent circuit and assigned the circuit values:
“Fig. 8 shows the equivalent circuit model of the coating’s electrochemical
impedance spectrum, which is Rs(Cc(Rc(QfRf)))(CpRp). Here, Rs, Rc, Rf, and Rp are the
resistances of the solution, coating, surface film layer of zinc-aluminum metal powder,
and surface activation dissolution reaction of zinc-aluminum metal powder,
respectively. Cc, Qf, and Cp are the capacitances of the coating, surface film layer of
zinc-aluminum alloy powder, and surface activation dissolution reaction of zinc-
aluminum metal powder, respectively. Table 1 summarizes the fitting results of the
electrochemical impedance spectra of all the coatings. It can be seen that the Rc values
increase with the increase of the amount of composites. The Rc values of coatings with
0%, 0.09%, and 0.18% CeO2@rGO are relatively small, around 25 Ω·cm2, indicating
that the metallic powder is not completely covered by the passivation film, and cannot
provide effective cathodic protection to the substrate. In Table 1, Rf represents the
resistance of the corrosion product film formed on the surface of the metallic powder.
The larger the value of Rf, the denser the corrosion product layer formed on the surface
of the metallic powder, and the better its protection effect. This will make it more
difficult for the metallic powder to undergo activation dissolution, thus enhancing the
anti-corrosion resistance of the coating. In addition, the value of Rf increases firstly but
then decreases with the further increase of CeO2@rGO. Among them, the coating with
0.18% CeO2@rGO exhibits the largest Rf value of 4064 Ω·cm2, indicating that the
corrosion product film formed by this coating is the densest, thus effectively forming a
physical barrier and delaying the corrosion of the metallic powder. Rp represents the
resistance of the activation dissolution reaction of the metallic powder. The larger the
value of Rp, the more difficult the reaction of the zinc-aluminum powder can occur, and
the better the corrosion resistance of the coating. According to the Table 1, Rp value is
increased with increasing the amounts of CeO2@rGO, and the coating added with 0.27%
CeO2@rGO exhibits the largest Rp. Since both the Rf and Rp values are critical to the
corrosion resistance of the coating, the sum of Rf and Rp is used in this study to represent
the corrosion resistance of the coating, with a larger Rf+Rp indicating a better corrosion
resistance. As shown in the Table 1, the coating with 0.18% CeO2@rGO exhibits the
largest Rf+Rp value, indicating its best corrosion resistance.”

Fig. 8. Equivalent circuit diagram.


Table 1. Fitting parameters of electrochemical impedance spectrum for chromium-free Dacormet
coatings reinforced with different contents of CeO2@rGO.
wt.% Rs Cc Rc CPEf n Rf Cp Rp
Ω·cm2 F/cm2 Ω·cm2 S-n/cm2 Ω·cm2 F/cm2 Ω·cm2
0% 14.92 7.353×10-7 24.47 3.203×10-4 0.7175 2712 4.212×10-8 64.16
-7 -4
0.09% 13.44 6.706×10 25.47 2.784×10 0.6558 3985 3.994×10-8 66.31
-7 -4
0.18% 11.45 1.321×10 26.51 3.145×10 0.6434 4064 4.459×10-8 74
-8 -4
0.27% 12.74 6.024×10 54 5.357×10 0.5357 3081 8.404×10-8 608.1
Replies to the comments from Reviewer #3:

Reviewer #3:
This manuscript deals with oil-lubricating, anti-corrosion, and photocatalytic behaviors
of Ce-functionalized RGO nanoparticles. The results and analyses presented in this
manuscript clarify that the paper needs many major revisions and cannot be worthy of
possible publication. The main points are mentioned below:
1. Language writing style (including some spelling mistakes) needs to be further
polished.

Reply:
Thank you for your comments.
We have carefully and thoroughly checked the whole paper, and improved all the
grammar and English in the revised manuscript, including the ones mentioned above.
They are marked with red color.

2. The evidence for the novelty of the research is insufficient, and the performance
advantages of the prepared CeO2@RGO nanoparticle compared to the GO and even
CeO2 should be further pointed out in the text.

Reply:
Thank you for your valuable and thoughtful comments.
We added the photocatalytic performance of CeO2, GO and CeO2@rGO and
provide relevant comparisons as following:
“Fig. 10(a) shows the photodegradation curves of the organic dye RhB by using
CeO2, GO and CeO2@rGO in the dark and under UV irradiation. It can be seen that the
removal of RhB by GO mainly occurs in the adsorption process under dark conditions,
and the removal rate under the UV irradiation is not decreased significantly, with the
maximum removal rate of 53.2%. CeO2 also mainly removes the RhB through the
adsorption process under dark conditions. However, under the UV irradiation, it
exhibits photocatalytic activity, with the maximum removal rate of 20.3%. CeO2@rGO
achieves a high removal rate of 90.5% for RhB, of which 77.8% is removed through
the adsorption process under dark conditions. The subsequent photocatalytic
degradation process further increases the removal rate up to 90.5%. Fig. 10(b) shows
the UV-vis spectra before and after the RhB adsorption. This shows that the RhB has a
maximum absorption peak at 554 nm before adsorption, but this peak is gradually
weakened after the RhB removal. The reason for this phenomenon may be that the rGO
nanosheets are excellent electron acceptors, while the UV-excited CeO2 is an excellent
electron donor. Therefore, the interactions between CeO2 and rGO nanosheets can cause
excited electrons to be effectively transferred from CeO2 to rGO nanosheets, thus
reducing the possible recombination of electron-hole pairs. Modifications of covalent
bonding between rGO and CeO2 can also improve the dispersion of rGO in the solution.
Furthermore, the layered structure of rGO can prevent the severe aggregation of CeO2
nanoparticles, thereby exposing more active sites to enhance the photocatalytic
performance of the CeO2@rGO composite.”

Fig. 10. (a) Photodegradation curves of different concentrations for RhB by the CeO2, GO and
CeO2@rGO under UV irradiation; (b) UV-vis spectra of RhB before and after photocatalytic
degradation; (c) Concentration changes of RhB dye solutions with time for CeO2@rGO in the
dark and under UV irradiation; (d) The pseudo-first-order reaction kinetics of the CeO2@rGO
composite for RhB degradation; (e) Comparisons of the removal efficiency for RhB dye of
different CeO2 based catalyst; (f) Recycling performance of the CeO2@rGO during adsorption-
photodegradation of the RhB solution.

3. Properties and applications of the Dacromet solution should be explained in the


"Introduction". Also, the authors have neglected the explanation of the role of the silane
agent in this solution and the interactions of Dacromet constituents with CeO2@RGO
nanoparticles.

Reply:
Thank you very much for your good suggestions, and we have added the
performance and applications of Dacromet coating in the introduction as following:
“Chromium-containing Dacromet coating is widely used as a corrosion inhibitor
for iron-based structural components in automotive, power, and marine industries.
However, there are various issues such as the toxic chromium solutions, poor bonding
strength and corrosion resistance with the substrate, and adding modified graphene
additives into the coatings has become one of the solutions [24].”
Silane coupling agent is very important for corrosion resistance of coating. We have
added the FT-IR spectrum of the coating with CeO2@rGO in the modified paper as the
following:
“Fig. 9 shows the FT-IR spectrum of the Q235 steel substrate with the chromium-
free Dacromet coating with the CeO2@rGO. It can be seen that the coating exhibits the
characteristic peaks corresponding to those of O-Ce, C-O-Si and Si-O-Si bonds [49,
50]. The hydrolyzed silane alcohol produced by the silane coupling agent in the coating
can form C-O-Si bond, and at the same time, the silane alcohol itself becomes
dehydrated and condensed to form linear or cyclic Si-O-Si network structures. The
Ce(OH)3 passivation layer is formed at the interface between the substrate and coating,
which prevents direct contact between the substrate and the external environment [51].
The formation of Si-O-Fe and Ce-Fe bonds at the interface between the coating and
substrate is also beneficial for improving the adhesion between the coating and
substrate, thereby improving its corrosion resistance [32, 52, 53]. Formation of these
bonds can also improve the density of the coating and enhance the bonding strength
between the coating and substrate.”

Fig. 9. FT-IR spectrum of the coatings with CeO2@rGO.


The interactions between CeO2@rGO and Dacromet coating can be explained in
three aspects. Firstly, the uniform dispersion of the CeO2@rGO (with its layered
structure) in Dacromet coating improve the adhesion onto the zinc-aluminum alloy
powders, which plays a crucial role in the solidification of the composite. Secondly, the
filling of pores within the coating with these nanometer-sized flakes and particles
greatly increases the density and hardness of the coating. The improved bonding
between the composite and the substrate and the formation of the passivation layer
enhance the corrosion resistances of the substrate and its adhesion to the coating. Finally,
the uniform distribution of CeO2@rGO in the coating helps to form the effectively
conductive channels, facilitate electron transmission. Therefore, it is beneficial to etch
the zinc-aluminum alloys with a lower potential, thus achieving the goal of protecting
the substrate.

4. Section 2 is better to be moved to the end of the paper (before the conclusion part).

Reply:
We thank the reviewer for this comment and we have moved the section 2 before
the conclusion part.

5. What is the difference between the conductivity of GO and RGO?

Reply:
Thanks for your good suggestion.
The reduced graphene oxide (RGO) has a better conductivity than graphene oxide.
Graphene oxide (GO) introduces a large number of oxygen and hydrogen functional
groups on the surface of graphene, which causes it to lose the π electron conjugation
structure of graphene, resulting in the decreased conductivity. Whereas RGO is
obtained by reducing the oxygen and hydrogen functional groups in the GO to restore
the π electron conjugation structure of graphene, thereby improving its conductivity.

6. The experimental section must be completed. For instance, the conditions of EIS
(frequency domain, perturbation amplitude, potential of measurements) and
polarization (scan rate, measurement range around Ecorr) tests are missed. Also, some
materials are not introduced, e.g., used CeO2 for comparison.
- Where did the reported ratio of CeO2/RGO (3:1) come from? What is the authors'
proof for that?
- Incorrect abbreviations of "SADE" or "SEAD" have been repeated many times, but
the correct abbreviation should be "SAED" (selected-area electron diffraction). Please
assimilate throughout the text.

Reply:
Thanks for these comments.
We have added the following contents in the experimental section:
“Electrochemical tests were carried out using a P4000 electrochemical workstation
with a three-electrode system. A platinum electrode and saturated calomel electrode
were served as the counter and reference electrodes, respectively, while the sample was
used as the working electrode with an active working area of 1 cm2. The
electrochemical tests were performed after the open circuit potential (OCP) was
stabilized in a 3.5% NaCl solution. The polarization curves were recorded at a scanning
speed of 0.5 mV/s within the range from -250 mV to +250 mV. To study the
electrochemical performances of the coatings with CeO2@rGO, the EIS tests were
conducted using the sample with its working area of 1 cm2, the frequency ranges from
10-2 Hz to 105 Hz and a sine wave amplitude of 10 mV. During the electrochemical
impedance spectroscopy (EIS) tests, the samples were immersed in a 3.5% NaCl
solution for different durations, and the impedance and corrosion resistance of the
coatings were evaluated. The salt spray tests were conducted continuously for 720 h in
a NaCl solution of 50 g/L, pH=6.5-7.2.”
CeO2 was also applied and added for comparisons in the photocatalytic section, as
shown in Fig. 10, which shows the removal of the organic dye RhB by CeO2, GO and
CeO2@rGO in both the dark and under UV irradiation. The removal of RhB by GO
mainly occurred in the adsorption process under dark conditions, and the removal rate
under UV irradiation was not decreased significantly, with the maximum removal rate
of 53.2%. CeO2 also mainly removes the RhB through the adsorption process under
dark conditions. However, under the UV irradiation, it exhibits photocatalytic activity,
with the maximum removal rate of 20.3%. CeO2@rGO achieved a high removal rate
of 90.5% for RhB, of which 77.8% was removed through the adsorption process under
dark conditions. The subsequent photocatalytic degradation process further increased
the removal rate up to 90.5%. Of which, 77.8% was removed through the adsorption
process under dark conditions, and the subsequent photocatalytic degradation process
further increased the removal rate to 90.5%. This is because the CeO2 loaded on the
rGO surface can form a good interface structure with the rGO surface, making it easier
to absorb photons and have a highly efficient electron transfer, thereby improving the
photocatalytic performance. In addition, the band structure of CeO2 will also be
changed on the rGO surface, thereby enhancing its photocatalytic performance. At the
same time, it can also prevent CeO2 agglomeration. Pure CeO2, on the other hand, has
surface electron defects and poor conductivity, and is prone to agglomeration during
application, thereby limiting its photocatalytic performance.
We made a mistake for the denoation of CeO2/RGO (3:1) and now we have made
the corresponding modifications for all the writings and abbreviation errors in the entire
paper.

Fig. 10(a). Photodegradation curves of different concentrations for RhB by the CeO2, GO and
CeO2@rGO under UV irradiation.

7. XRD and TEM sections. The interlayer spacing should have increased if the CeO2
species could be placed between the RGO layers. However, the related GO peak in
XRD disappeared after Ce-grafting! Instead, in TEM assessment, the authors assert the
formation of the dense intercalated structure thereby Ce-based reduction of GO, while
it is previously mentioned that by reduction of GO via Ce decoration, the lamellar
structure of graphene is destroyed. Destruction of RGO regular structure (despite its
properties' diminution, e.g. barrier effect) has not resulted from other tests. Please
clarify this conflict and revise the text.
- The size of CeO2 crystals should be calculated from XRD curves (using the Debye-
Scherrer equation).

Reply:
We thank the reviewer for this comment and suggestion.
The disappearance of the diffraction peak of rGO may be due to the crystal growth
of CeO2 nanoparticles between the layers of rGO, causing the disruption of the regular
layered stacking structure of rGO and separating layers of the rGO. In addition, it may
also be that the crystalline peaks of CeO2 is relatively high, which covers the weak
signal of rGO [3, 4]. And we added the following contents in the revised paper:
“The reason is attributed to that CeO2 was generated between the rGO layers after
hydrothermal process, which destroys the regular structure of rGO [29]. In addition, it
may also be that the crystalline peaks of CeO2 is relatively high, which covers the weak
signal of rGO [30].”
We added the calculation results of the crystallite sizes using the Debye-Scherrer
equation as following:
“According to the characteristic peaks in the XRD spectrum of CeO2@rGO, the
Debye-Scherrer equation (Eq. 1) was used to estimate the crystallite size of CeO2 [43].
The average crystallite size was calculated to be ~11.02 nm, which is very close to the
result obtained by TEM.

D= βcosθ (1)

where D (nm) is crystallites size. K is Scherrer constant. λ (nm) is wavelength of the X-


ray sources. β (radians) is the value of full width at the half maximum (FWHM). θ
(radians) is diffraction angle positions.”

[3] Bakry A, Alamier W, Salama R, El-Shall M and Awad F 2022 Remediation of water
containing phosphate using ceria nanoparticles decorated partially reduced graphene
oxide (CeO2-PRGO) composite Surfaces and Interfaces 31 102006.
[4] Min C, He Z, Liu D, Jia W, Qian J, Jin Y and Li S 2019 Ceria/reduced Graphene
Oxide Nanocomposite: Synthesis, Characterization, and Its Lubrication Application
ChemistrySelect 4 4615.

8. EDX section. What is the RGO layer thickness on the substrate? Does the EDX test
(with a micron-scale penetration depth of X-ray) suitable for the detection of this layer?
Why the Fe element has not been detected in these specimens?!

Reply:
Thank you very much for your question.
The reason why the iron element was not detected is that the coating has a dense
structure and its thickness is around 10 μm, as shown in Fig. R2. It can be seen that the
coating without CeO2@rGO has obvious holes, whereas the coating containing
CeO2@rGO shows a dense interfacial structure. In addition, it can be seen from the
energy spectrum that the C element content of the coating with CeO2@rGO is
significantly increased, indicating the presence of rGO in the coating. EDX is not
suitable for detecting coated surfaces, as X-rays has limited penetrating depth.

Fig. R2. SEM and EDX of cross section of the coatings with different contents of the CeO2@rGO:
(a-a1) 0%; (b-b1) 0.18%.
We added the SEM images and EDX results of the coatings with 0%, 0.18%, and
0.27% CeO2@rGO after corrosion (Fig. 6). Iron was also detected in coatings with other
formulations, as shown in Figs. 6(a, d):
“Fig. 6 shows SEM images and EDX results of the coatings with 0%, 0.18%, and
0.27% CeO2@rGO after the 720 h salt spray test. It can be seen that the pore contents
on the surface of the coatings with 0.18%, 0.09%, 0%, and 0.27% CeO 2@rGO are
gradually increased. From the enlarged images, the morphology of the coating surface
is gradually changed from a metallic powder shape to a sponge-like structure, and
finally to a typical needle-shaped structure of the corrosion product. Among them, the
surface of the coating with 0.18% CeO2@rGO is still very dense, and the morphology
of the metallic powder can still be observed. Whereas for the coatings with 0% and
0.27% CeO2@rGO, the typical corrosion products of the coatings, i.e., zinc aluminum
chloride hydrates, have an obvious needle-shaped structure. EDX analysis of the
surface of each coating shows that the surface of the coating with 0.18% CeO2@rGO
is mainly composed of Zn, Al, and O elements. Compared with the uncorroded coating,
the amount of O has been increased, indicating that the metallic powder has generated
zinc aluminum oxides after the corrosion test. However, the corroded coating is still
effective for shielding. For the coatings with 0% and 0.27% CeO2@rGO, Fe elements
were detected, and there was more oxygen detected, indicating that more corrosion
products and Fe oxides were generated on the surface of the coating. This can also be
clearly seen from the image where the red rust was produced on the surface of the
substrate. The reason why the EDX results of the coating with 0.18% CeO 2@rGO did
not show Fe elements is that this coating has the best corrosion resistance. It can also
be seen that there was no red rust produced on the surface of the substrate coated with
this composite layer.”

Fig. 6. SEM and EDX images of surface of chromium-free Dacroment coating reinforced with
different contents of CeO2@rGO before and after corrosion: (a-a4) 0% CeO2@rGO; (b-b4) 0.09%
CeO2@rGO; (c-c4) 0.18% CeO2@rGO; (d-d4) 0.27% CeO2@rGO.

9. Section 4.2, "To sum up, the wear mechanism is changed from oil-film lubrication to
reaction film lubrication after adding the nanocomposite into the lubricating oil". The
lubricating mechanism changing via CeO2@RGO addition is not explained precisely
and is ambiguous in the text. What is the difference between oil-film lubrication and
reaction-film one?

Reply:
We thank the reviewer for this good comment and questions.
Oil-film lubrication refers to the surface of two friction pairs undergoing relative
motions, which are fully separated by the oil film layer between the lubricated surfaces,
without direct contact of the friction pairs [5]. Reaction film lubrication refers to that
various types of chemical elements which are contained in the lubricating oil undergo
chemical reactions with metals at high temperatures, generating new chemical
substances. These chemical substances cover the surfaces of the metal friction pairs and
effectively isolate the contacting friction pairs. Meanwhile, due to their excellent
lubricating properties, they can easily slide under the action of shear force, reducing the
frictional resistance of metal moving parts [6].
As the friction process progresses, the small-sized CeO2@rGO quickly enters the
surface of the friction pairs. Under a heavy load, the steel disk surface has been worn
and the worn depth is gradually increased. The CeO2@rGO that enters the worn surface
at this time forms a discontinuous oil-film, which is relatively thin and unstable. After
a short period, the composite will fill the pits caused by the friction on the steel disk
surface. This is beneficial for the CeO2@rGO nano-additives to form a continuous,
thicker, and more stable boundary lubricating oil-film in the worn area of the friction
pairs. This lubricating oil-film can completely prevent the direct contact between the
friction pairs, and its formation improves the anti-friction and anti-wear properties of
the lubricating oil. In addition, in the CeO2@rGO composite, the existence of CeO2
nanoparticles can transform the sliding friction between worn surfaces into rolling
friction, improving the lubricating effect of the lubricating oil. In summary, the main
lubrication mechanism for improving the friction performance of the lubricating oil
with CeO2@rGO composite is that small-sized nanoparticles can quickly enter the
surface of the friction pairs, form boundary lubrication oil-film in the worn area, and
effectively fill the worn pits.

[5] Yu X, Feng Y, Gao W, Shi G, Li S, Chen M, Zhang R, Wang J, Jia W, Jiao J and
Jiang H 2022 Study on lubrication performance of hydrostatic clearance oil film
considering multi-factor coupling International Journal of Hydrogen Energy 47 40083.
[6] Kinoshita H, Nishina Y, Alias A and Fujii M 2014 Tribological properties of
monolayer graphene oxide sheets as water-based lubricant additives Carbon 66 720.

10. The evaluations/interpretations of corrosion test results are very infirm and
presented with many ambiguities/missing data! This part requires comprehensive
revisions:
- The text of EIS analyses is replicated twice in the manuscript. Instead, the polarization
section is omitted!
- There is no data reporting for corrosion tests such as coating and charge transfer
resistances, coating and double layer capacitances, and heterogeneity constant values
as well as Ecorr, the Tafel branches, and corrosion current density for accurately
interpreting the anti-corrosion performance of RGO and CeO2@RGO in the Dacromet
coating.
(1)- Do the CeO2@RGO nanoparticles have only the barrier effect in the coating?
Formation of Ce(OH)3 passive/inhibitive layer on the steel/coating interface (feasible
reaction in cathodic regions: Ce+3 + 3OH- → Ce(OH)3↓) could be concluded. These
compounds have at least a barrier/active dual-functional effect in the anti-corrosion
films. Cerium grafting grants this property to the GO nanoparticles in addition to their
dispersion improvement.
- In the Nyquist plots, the horizontal and vertical axes should be the same in length and
major/minor axis ticks. In this form, the depression of curves from ideal capacitance
can be perceived accurately.
- It is stated: "It can be seen that the capacitive arc radius of CeO2@rGO coating is
larger than that of coating without adding the CeO2@rGO." Please magnify the Nyquist
plo
- The electrical equivalent circuit (EEC) utilized for fitting the curves is wrong! In non-
ideal conditions of the electrode/electrolyte systems (heterogeneous corroded/inhibited
surfaces), CPE has to be used for the coating-related time constant instead of the C
parameter. The corrosion behavior should be explained regarding the curves fitting via
appropriate EEC. Furthermore, it is better to give the corresponding Bode plots for a
better understanding of the scientific readers of this paper.
- The anti-corrosion mechanism(s) of these compounds in the Dacromet-coated steel
must be discussed more in detail.
(2)- What is the effect of physical/chemical bonding between Ce and steel substrate
or between Ce and GO on the anti-corrosion performance of the coating? The role of
Ce-Fe bonding on adhesion reinforcement and thus corrosion resistance increment of
the Cr-free Dacromet coating should be clearly explained. The presented explanation,
in this matter, is very sketchy.

Reply:
Thanks for your detailed comments and questions. We have modified the lengths
of the horizontal and vertical axes and the major/minor axis ticks on the Nyquist plot to
be the same. The Nyquist plot has been magnified, and the corresponding Bode plots
are also provided.
We also added the results for the coatings with the 0.09% and 0.27% CeO2@rGO
as following:
“Fig. 7(a) shows the Tafel curves of the coatings with different contents of
CeO2@rGO. Except for the coating with 0.27% CeO2@rGO, the corrosion potentials
of all the other coatings are lower than that of the substrate. The potential difference is
between 0.0504 V to 0.1092 V. All these results indicate that the coatings added with
CeO2@rGO can provide a certain degree of cathodic protection for the substrate [47].
Comparing the coatings added with different contents of CeO2@rGO, the corrosion
potential of the coating with 0.18% CeO2@rGO shows the most negative value, and the
corrosion potential of the coating with 0.27% CeO2@rGO shows the most positive
value. In addition, the corrosion current densities of the coatings with 0.09% and 0.18%
CeO2@rGO are decreased by an order of magnitude if compared to that of the coating
without adding the CeO2@rGO. The corrosion resistance was significantly increased
after adding the CeO2@rGO, indicating that a certain amount of CeO2@rGO can reduce
the Icorr value of the coating and improve its corrosion resistance. However, the Icorr
value of the coating with 0.27% CeO2@rGO is increased. This is because after adding
an excessive amount of CeO2@rGO, the rGO could become contacted with each other
and can form an effective corrosion path, thus increasing the conductivity of the coating
and the corrosion current density. It is also possible that adding too much amount of
composite causes severe agglomeration, producing more pores for the corrosion
medium easier to pass through, thus reducing the corrosion resistance of the coating.”
Fig. 7. (a) Tafel polarization curves of chromium-free Dacromet coatings reinforced with different
contents of CeO2@rGO; (b) Nyquist diagram of chromium-free Dacromet coatings reinforced
with different contents of CeO2@rGO; Bode diagram of chromium-free Dacromet coatings
reinforced with different contents of CeO2@rGO: (c) phase frequency diagram; (d) amplitude-
frequency diagram.
“Fig. 7(b) shows the Nyquist diagram of the coatings with different contents of
CeO2@rGO in a 3.5% NaCl solution. It can be seen that the capacitive arc radii of the
coatings with 0.09% and 0.18% CeO2@rGO are much larger than that of the coating
with 0% CeO2@rGO, and the density of the coating is increased after adding a certain
amount of composite, both of which can effectively improve the corrosion resistance
of the coating [48]. This is because the graphene is uniformly dispersed in the coating.
It can effectively fill up the defects in the coating, which plays a positive role as a
physical barrier to hinder the intrusion of the corrosive medium. When the CeO2@rGO
content is 0.27%, the capacitive arc radius of the coating decreases, indicating that
adding an excessive amount of composite will reduce its dispersion in the coating,
leading to agglomeration, increased defects in the coating, and also decreased corrosion
resistance of the coating.”
“Figs. 7(c-d) show the Bode diagrams of coatings added with different amounts of
the composite. From the frequency response diagram, the |Z| values of the coatings with
0.09% and 0.18% CeO2@rGO are higher than those of the coatings without adding the
composite. Among them, the coating with 0.18% CeO2@rGO exhibits the maximum
low-frequency impedance value (|Z|f→0) of about 3.68×103 Ω·cm2, indicating its
minimum corrosion rate and relatively good corrosion resistance. When the added
amount of CeO2@rGO reaches 0.27%, the |Z| value of the coating becomes decreased,
and its corrosion resistance is also decreased. This again proves that adding a certain
amount of composite enhances the corrosion resistance of the coating, but adding
excessive amounts of composite will reduce the corrosion resistance of the coating.
This may be due to the van der Waals force which causes the rGO to be agglomerated
locally, thus increasing the pores in the coating and decreasing its shielding effect.”
“Fig. 8 shows the equivalent circuit model of the coating’s electrochemical
impedance spectrum, which is Rs(Cc(Rc(QfRf)))(CpRp). Here, Rs, Rc, Rf, and Rp are the
resistances of the solution, coating, surface film layer of zinc-aluminum metal powder,
and surface activation dissolution reaction of zinc-aluminum metal powder,
respectively. Cc, Qf, and Cp are the capacitances of the coating, surface film layer of
zinc-aluminum alloy powder, and surface activation dissolution reaction of zinc-
aluminum metal powder, respectively. Table 1 summarizes the fitting results of the
electrochemical impedance spectra of all the coatings. It can be seen that the Rc values
increase with the increase of the amount of composites. The Rc values of coatings with
0%, 0.09%, and 0.18% CeO2@rGO are relatively small, around 25 Ω·cm2, indicating
that the metallic powder is not completely covered by the passivation film, and cannot
provide effective cathodic protection to the substrate. In Table 1, Rf represents the
resistance of the corrosion product film formed on the surface of the metallic powder.
The larger the value of Rf, the denser the corrosion product layer formed on the surface
of the metallic powder, and the better its protection effect. This will make it more
difficult for the metallic powder to undergo activation dissolution, thus enhancing the
anti-corrosion resistance of the coating. In addition, the value of Rf increases firstly but
then decreases with the further increase of CeO2@rGO. Among them, the coating with
0.18% CeO2@rGO exhibits the largest Rf value of 4064 Ω·cm2, indicating that the
corrosion product film formed by this coating is the densest, thus effectively forming a
physical barrier and delaying the corrosion of the metallic powder. Rp represents the
resistance of the activation dissolution reaction of the metallic powder. The larger the
value of Rp, the more difficult the reaction of the zinc-aluminum powder can occur, and
the better the corrosion resistance of the coating. According to the Table 1, Rp value is
increased with increasing the amounts of CeO2@rGO, and the coating added with 0.27%
CeO2@rGO exhibits the largest Rp. Since both the Rf and Rp values are critical to the
corrosion resistance of the coating, the sum of Rf and Rp is used in this study to represent
the corrosion resistance of the coating, with a larger Rf+Rp indicating a better corrosion
resistance. As shown in the Table 1, the coating with 0.18% CeO2@rGO exhibits the
largest Rf+Rp value, indicating its best corrosion resistance.”

Fig. 8. Equivalent circuit diagram.


Table 1. Fitting parameters of electrochemical impedance spectrum for chromium-free Dacormet
coatings reinforced with different contents of CeO2@rGO.
wt.% Rs Cc Rc CPEf n Rf Cp Rp
Ω·cm2 F/cm2 Ω·cm2 S-n/cm2 Ω·cm2 F/cm2 Ω·cm2
0% 14.92 7.353×10-7 24.47 3.203×10-4 0.7175 2712 4.212×10-8 64.16
0.09% 13.44 6.706×10-7 25.47 2.784×10-4 0.6558 3985 3.994×10-8 66.31
0.18% 11.45 1.321×10-7 26.51 3.145×10-4 0.6434 4064 4.459×10-8 74
0.27% 12.74 6.024×10-8 54 5.357×10-4 0.5357 3081 8.404×10-8 608.1
We have supplemented the FT-IR spectrum of the coatings with CeO2@rGO as
following:
“Fig. 9 shows the FT-IR spectrum of the Q235 steel substrate with the chromium-
free Dacromet coating with the CeO2@rGO. It can be seen that the coating exhibits the
characteristic peaks corresponding to those of O-Ce, C-O-Si and Si-O-Si bonds [49,
50]. The hydrolyzed silane alcohol produced by the silane coupling agent in the coating
can form C-O-Si bond, and at the same time, the silane alcohol itself becomes
dehydrated and condensed to form linear or cyclic Si-O-Si network structures. The
Ce(OH)3 passivation layer is formed at the interface between the substrate and coating,
which prevents direct contact between the substrate and the external environment [51].
The formation of Si-O-Fe and Ce-Fe bonds at the interface between the coating and
substrate is also beneficial for improving the adhesion between the coating and
substrate, thereby improving its corrosion resistance [32, 52, 53]. Formation of these
bonds can also improve the density of the coating and enhance the bonding strength
between the coating and substrate.”

Fig. 9. FT-IR spectrum of the coatings with CeO2@rGO.


From the FT-IR spectrum of the coatings with CeO2@rGO, it can be observed that
there exist O-Ce bonds and O-H groups from the epoxy resin in the coating, indicating
the formation of Ce(OH)3 passivation layer at the interface between the substrate and
the coating, as shown in Eq. R1. The main function of the passivation layer is to prevent
direct contact between the steel and the external environment, thus slowing down the
corrosion rate of the steel. Additionally, it also exhibits a certain degree of activity. This
suggests that the compound has a barrier/active dual-functional effect in the anti-
corrosion films.
Ce3+ + 3OH- → Ce(OH)3↓ (R1)
The physical/chemical bonding between Ce and the steel substrate or Ce and rGO
can have a positive impact on the anti-corrosion performance of the coating. First, it
can enhance the anti-corrosion performance. The physical/chemical bonding between
CeO2 and the steel substrate or between CeO2 and rGO can form a stable and inert
protective film to protect the steel surface against external corrosion. This can greatly
reduce the corrosion rate of the steel and the production of corrosion products on the
surface, thus improving the anti-corrosion performance of the coating. Secondly, it can
also enhance the adhesion. The physical/chemical bonding of the compounds can
effectively enhance the adhesion and wear resistance between the coating and the
substrate, thereby enhancing the stability and durability of the coating.
In the chromium-free Dacromet coating, the Ce-Fe bond can have a positive impact
on the enhancement of adhesion and corrosion resistance. Firstly, it can enhance the
adhesion. The Ce-Fe bond can make the bond strength between the Dacromet coating
and the substrate stronger, thereby improving the adhesion, wear resistance, and
corrosion resistance of the coating. Secondly, it can improve the corrosion resistance.
The Ce-Fe bond can form a stable oxide film on the surface of the chromium-free
Dacromet coating, effectively preventing the generation of corrosion products on the
surface of the coating and improving the anti-corrosion performance of the coating. In
addition, CeO2 itself has anti-corrosion properties, which can further improve the anti-
corrosion performance of the chromium-free Dacromet coating.
We have conducted in-depth exploration and modification of the corrosion
resistance mechanism:
“In this study of the chromium-free Dacromet coating, we used the molybdic acid
instead of the traditionally used chromic acid. The corrosion resistance mechanism of
the coating has also been changed to the formation of insoluble molybdate on the
substrate as shown in Eq. 3. The formed oxide film on the substrate’s surface passivates
it from corrosion, which competes with Cl- during corrosion, thus reducing the
corrosion of Cl-. In addition, the corrosion resistance performance of the coating is
further improved after the addition of CeO2@rGO. The modified rGO is uniformly
dispersed in the coating, and its layered structure easily adheres to the zinc-aluminum
alloy powder and plays a key role in the solidification of the CeO2@rGO. On the other
hand, the nanoflakes and nanoparticles fill the porous structure of the coating, greatly
increasing its density and hardness. The homogeneous distribution of the CeO2@rGO
in the coating promotes the formation of a conductive channel, facilitates the electron
transfer, and is thus beneficial for sacrificing zinc-aluminum alloy with a lower
potential for protecting the substrate. The hydrolysis of the silane coupling agent in the
coating produces a silanol that can form C-O-Si, and Si-O-Fe is also formed at the
interface between the coating and the substrate [49]. At the same time, the self-
dehydration and condensation of silanol form a linear or cyclic Si-O-Si network
structure. The Ce(OH)3 passivation layer is formed at the interface between the
substrate and the coating, which actively prevents direct contact between the substrate
and the external environment. This indicates that the compound in the anti-corrosion
film has barrier/active dual-functions. The presence of Ce-Fe bonds can also improve
the adhesion between the coating and substrate, thereby improving the wear resistance
and corrosion resistance of the coating [52]. Therefore, the corrosion resistance of the
chromium-free Dacromet coating with the added CeO2@rGO has been improved.
MoO42- + Fe2+ → FeMoO4 → Fe2(MoO4)3 (3)”
Highlights (for review)

Highlights

 CeO2 modified reduced graphene oxide nanocomposite (CeO2@rGO) was

prepared by a hydrothermal method.

 Lubrication mechanism of lubricating oil with CeO2@rGO was changed from oil

film lubrication to reaction film lubrication.

 Adding CeO2@rGO to the Dacromet coating can improve its density and corrosion

resistance.

 CeO2@rGO was used as a photocatalyst to degrade dyes, and the good adsorption-

photocatalytic activity was demonstrated.

You might also like