You are on page 1of 28

 

 
Preparation of graphene oxide modified by titanium dioxide to enhance the
anti-corrosion performance of epoxy coatings

Zongxue Yu, Yu Ma, Yi He, Ling Liang, Liang Lv, Xiang Ran, Yang
Pan, Zhi Luo

PII: S0257-8972(15)30067-0
DOI: doi: 10.1016/j.surfcoat.2015.06.027
Reference: SCT 20326

To appear in: Surface & Coatings Technology

Received date: 14 January 2015


Revised date: 12 June 2015
Accepted date: 15 June 2015

Please cite this article as: Zongxue Yu, Yu Ma, Yi He, Ling Liang, Liang Lv, Xiang
Ran, Yang Pan, Zhi Luo, Preparation of graphene oxide modified by titanium dioxide to
enhance the anti-corrosion performance of epoxy coatings, Surface & Coatings Technology
(2015), doi: 10.1016/j.surfcoat.2015.06.027

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
ACCEPTED MANUSCRIPT

Preparation of graphene oxide modified by titanium dioxide


to enhance the anti-corrosion performance of epoxy coatings
Zongxue Yu a,*, Yu Ma a, Yi He a,b,†, Ling Liang a, Liang Lv a, Xiang Ran a,
Yang Pan a, Zhi Luo c

T
IP
a
Southwest Petroleum University, School of Chemistry and Chemical Engineering, Chengdu, 610500, China.
b
State Key Lab of Oil and Gas Reservoir Geology and Exploitation, Chengdu, 610500, China.

R
c
Sichuan TongKai Energy Science and Technology Development Co., LTD, Bazhong, 636600, China

SC
Abstract

Solvent-based epoxy resins are often used for the anti-corrosion purpose but their cured process fabricating

NU
plentiful micro-pore via solvent evaporation is an intrinsic shortcoming and it is thus necessary to obstacle their
MA
micro-pore for enhancement antiseptic property. With the purpose of the enhancement, we synthesized TiO2–GO

sheet hybrids using titanium dioxide loading on graphene oxide sheets with the help of
D

(3-aminopropyl)trimethoxysilane, and dispersing the sheets into epoxy resin at a low weight fraction of 2%. The
TE

electrochemical impedance spectroscopy (EIS) test and monitoring coatings’ morphology in corrosion process
P

reveal that the corrosion resistant performance is significantly enhanced by the addition of TiO2–GO hybrids to
CE

epoxy. Comparisons with other nanofillers including TiO2 and graphene oxide (GO) indicate that TiO2–GO
AC

hybrids exhibit an obvious superiority in enhancing the corrosion resistant of epoxy coatings at the same contents.

The superiority of the TiO2–GO hybrids is related to their exfoliation, dispersion and excellent plugging

micro-pore property arising from their laminated structure. Furthermore, the corrosion resistant mechanisms were

tentatively proposed for the TiO2–GO/epoxy coatings.

Keywords: Graphene oxide, Titanium dioxide, Nanocomposites, Composite coatings, Anti-corrosion

1. Introduction

The oxidation of graphene has recently attracted considerable interest, mainly due to prospects of

* Corresponding author. Tel.: +86 28 83037315


E-mail address: haiqingy@163.com (Zongxue Yu).
† He also supported the research.

1
ACCEPTED MANUSCRIPT

using graphene oxide (GO) for low-cost production of large quantities of graphene [1]. The primary

route for the GO production is the Hummers method based on oxidation treatment of the natural

graphite [2, 3], which manufactures an inhomogeneous binding of oxygen functional groups on the

T
IP
graphene sheets. Those interesting groups provide a handle for further surface chemical

R
functionalizations via well-developed carbon chemistry [4-7]. Existing experimental evidence shows

SC
that GO sheets contain epoxy, hydroxyl and carboxyl groups [5]. GO can be well dispersed and

exfoliated in epoxy at a very low filler content [8]. However, thanks to the high surface area, Van der

NU
Waals’ interaction, and vacuum filtration in preparation process [9, 10], GO has a pronounced tendency
MA
to form severe aggregations at a relatively high loading and its predicted properties fail to be fully

reached in polymer [10-13]. Undeniably, with the purpose of achieving the successful application of
D

GO in polymer composites, the key issue should be resolved, improving dispersion and exfoliation of
TE

sheet-filler. In recent years, a variety of processing methods e.g., in situ polymerization, surface
P

grafting and decorating with nanoparticles, and organic modification have been proposed to achieve the
CE

goal [7, 14]. Especially, nanoparticles decorated on surface have more efficient separating sheets,
AC

resulting from that the layer spacing is increased. Additionally, there are many nanoparticles were

utilized for decoration, such as silica, titanium oxide, aluminium oxide and zinc oxide. The nanometer

powder of titanium dioxide (nano-TiO2) has been widely investigated as nano-TiO2-based materials

because of the good property of the particle. Unfortunately, nano-TiO2 powders also have a strong

tendency to aggregate because of their high surface area and high polarity [15-17]. To reduce the

shortcoming of the nanoparticles, many researchers have focused on surface modification. More

importantly, a novel method of incorporating inorganic nanoparticles into graphene sheets has been

reported in the recent related literatures [4, 5, 18, 19], but only a few associated with nano-TiO2.

2
ACCEPTED MANUSCRIPT

Consequently, we embedded nano-TiO2 into GO sheets (TiO2–GO hybrids), which changed the sheets

closely stacking into loose state as well as precluded the aggregation of nanoparticles under their

utilization alone in polymer.

T
IP
Epoxy resins are widely used as adhesives, coatings, structural materials and composite matrix [20,

R
21]. However, epoxy resins are inherently shortcoming, fabricating the high temperature curing epoxy

SC
solvent-borne coatings, which makes them inevitably produced plentiful micro-pore via solvent

evaporation [20, 22]. The corrosive liquid medium can quickly permeate the coating into the protected

NU
substrate using micro-pores generating poor corrosion resistance. Extensive studies have thus been
MA
conducted to toughen epoxy resins using nanofillers [6, 20, 21]. Epoxy resins containing film

nanofillers have shown the greatest barrier property arising from that sheet-like nanomaterial have a
D

more efficient obstruction for micro-pores than others, and GO own the property. With the purpose of
TE

stick nano-TiO2 to sheets, nanoparticle could be a useful method for forming reactive activity on its
P

surface.
CE

In this communication, we focused on developing a route to prepare a covalently bonded


AC

nanoparticles–GO hybrids material, which could have important properties and applications. The

TiO2–GO nanocomposites were synthesized with the help of 3-aminopropyltriethoxysilane (APTS) in

this work. In this process, the amine group of APTS conjugated to the handle (oxygen functional

groups on the graphene sheets). The GO, nano-TiO2 and TiO2–GO hybrids with respect to low weight

loading were dispersed into epoxy resin as nanofillers, and composite coatings were obtained from

mixtures curing process. Discrepancies in the anti-corrosion, exfoliation and dispersion performance of

above composite coatings were examined.

3
ACCEPTED MANUSCRIPT

2. Experimental

2.1 Materials

The natural flake graphite used to prepare GO was provided by Chengdu Kelong Chemical Reagent

T
Factory. Concentrated sulfuric acid (98%), potassium permanganate (KMnO 4), sodium nitrate

IP
(NaNO3), hydrogen peroxide aqueous solution (30%), N, N-dimethylformamide (DMF), anhydrous

R
SC
ethanol (analytical reagent grade), hydrochloric acid (HCl), and sodium hydroxide (NaOH) were

obtained from Chengdu Kelong Chemical Reagent Factory. Silane coupling agents (3-aminopropyl)

NU
trimethoxysilane (APTS) was supplied by Chengdu Kelong Chemical Reagent Factory. Nano-TiO2
MA
(about 50 nm) was purchased from the Chengdu Branch of the Chinese Academy of Sciences. The

epoxy emulsion (WSP-6101) and its hardener used in this research were supplied by Bluestar
D

technology wuxi resin factory. Deionized water (DI water) was produced by a water purification
TE

machine (UPC-III-40L, Ulupure).


P

2.2 Synthesis of GO
CE

GO was manufactured by Hummer’s method [2, 23]. In a typical procedure, 98% H2SO4 (24 mL)
AC

was added into a 250 mL flask dumped with natural graphite (0.8 g) at 0 °C (ice bath), followed by the

addition of solid NaNO3 (1 g) and KMnO4 (5 g), slowly after 0.5 h. Magnetic stirring was used in the

whole process. At 39 °C, the mixture was stirred with a magnetic stirring bar for 2.5 h. Water (41 mL)

was added into the mixture at 95 °C, and H2O2 (10%, 120 mL) was added after 0.5 h. The solution was

filtered and washed with HCl (0.5 M) aqueous solution (100 mL) to remove metal ions, followed by

repeated washing with water and vacuum filtration. The resulting solid was dispersed in water by

ultrasonication for 3 h to make a GO aqueous dispersion. The obtained brown dispersion was then

subjected to 35 min of centrifugation at 4000 rpm to remove any aggregates. To remove most of the

4
ACCEPTED MANUSCRIPT

water, the suspension was approached though evaporation process. Finally, the obtained GO was dried

under vacuum at 60 °C for 48 h.

2.3 Synthesis of f-TiO2

T
IP
Nano-TiO2 (0.2 g) was added into anhydrous ethanol (63 g) to form a homogeneous solution by

R
stirring and dropwise addition of DI water (5.6 g) at 78 °C for 6 h. Subsequently, the product was

SC
filtrated and washed with anhydrous ethanol three times, and dried at 60 °C for 24 h.

2.4 Synthesis of TiO2–GO hybrids

NU
The preparation of TiO2–GO is schematically illustrated in Fig. 1. First, the f-TiO2 (0.05 g) was
MA
dispersed in DMF (100 mL) to form a homogeneous suspension via ultrasonication technique for 1 h,

and the as-prepared GO (0.15 g) was added into the suspension. Subsequently, the solution was stirred
D

for 5 h at 105 °C. Finally, the product was obtained by filtrating and washing with anhydrous ethanol
TE

five times, and drying at 60 °C for 24 h.


P

2.5 Preparation of composite coatings


CE

To prepare composite coating precursor, the epoxy emulsion was accompanied by a known amount
AC

of nanometer materials (2wt.% mass fractions of GO, nano-TiO2, and TiO2–GO hybrids, respectively).

And then, original and processed mixtures were sprayed on the steel surface through the ultrasonic

oscillation and mechanical stirring process. Besides, redundant emulsion was carefully poured into a

homemade container, respectively. Finally, The epoxy-coated steels and cured epoxy coatings were

attained using two steps: steels and containers were baking in an oven at 120 °C for 1 h followed by

220 °C for other 1.5 h [22]. The dry film thickness of coatings was required to be measured utilizing

screw micrometer and calibrator. And cured epoxy coatings were named pure epoxy coating,

nano-TiO2/epoxy coating and TiO2–GO/epoxy coating, respectively.

5
ACCEPTED MANUSCRIPT

2.6 Characterization

The infrared absorption peaks of nanomaterials were observed through infrared spectrum (FT-IR;

WQF520). The crystalline structure and composition of the as-products in powder form were analyzed

T
IP
by an X-ray diffractometer (XRD; Bruker, D8 ADVNCE) with monochromatized copper Kα radiation

R
at λ = 1.5418 Å, and composite coatings also were analyzed for evaluation filler’s exfoliation and

SC
dispersion. Morphologies of the as-prepared nanocomposites and coatings’ fracture surface were

observed using a scanning electron microscopy (FESEM; JSM-7500F) and a JEOL JEM-2100 high

NU
resolution transmission electron microscope (HR-TEM). Furthermore, the fracture surface was
MA
obtained through breaking the cured epoxy resin in liquid nitrogen. The thermostability of the powders

was determined in nitrogen atmosphere through thermogravimetry analysis (TGA; Mettler-Toledo


D

SDTA85) at a rate of 10 °C/min.


TE

Electrochemical impedance spectroscopy (EIS) measurements were performed by an


P

electrochemical workstation CHI660D (Shanghai Chenhua Device Company, China). In this


CE

configuration, a classic three-electrode cell arrangement was employed [24, 25]: The epoxy-coated
AC

steel coupon served as the working electrode with a square –3.6 cm2; a saturated calomel electrode

(SCE) was the reference; and a platinum plate electrode with a dimension of served as

auxiliary. Electrolyte for all electrochemical evaluations was 3.5 wt.% NaCl solution (immersion

solution) and the temperature was approximately 25 °C. Crucially, the coatings evaluated in the

electrochemical measurements had similar thickness (about 50 μm) as those used in the morphological

monitor since they were prepared following the same procedures. The impedance measurements were

performed in the frequency range of 10 kHz–100 mHz utilizing an AC overpotential with amplitude of

6
ACCEPTED MANUSCRIPT

10 mV in different immersion times up to 10 days. The obtained impedance data were evaluated

employing ZSimpWin software.

After the epoxy-coated steel soaking in the electrolyte solution times up to 20 days, the surface of

T
IP
coatings were monitored utilizing another FESEM (ZEISS, ƩIGMA). And the static contact angle

R
meter (XED-SPJ, Peking Hako experimental instrument plant, China) was employed to estimate the

SC
water contact angle (WCA) of soaked-coatings.

NU
3. Results and Discussion

3.1. Characterization and morphology of TiO2–GO hybrids


MA
FT-IR was utilized to investigate the completion of functionalization and hybridization. The FT-IR

spectrum of the as-synthesized TiO2–GO hybrids and GO are shown in Fig. 2. Fig. 2(a) shows the
D

spectrum of GO. According to Yue Lin [26], the characteristic absorption peaks of GO were observed
TE

at 3447 cm-1 (–OH), 1741 cm-1 (C=O), 1220 cm-1 (C–O), and 1120 cm-1 (C–O–C), indicating that the
P

GO sheets existed in the hydroxyl, carboxyl, and epoxide groups. The peak position of =C–H band was
CE

located at 1400 cm-1 [26, 27]. Fig. 2(b) presents the TiO2–GO hybrids. According to Yue Lin [26] and
AC

Fuan He [5], the absorption peaks at 2928 and 2859 cm-1 representing the C–H bonding, the peak at

1115 cm-1 representing the Si–O–Si bonding [26], the peak at 1035 cm-1 representing the Si–O–C

bonding [26], and the peak at 925 cm-1 representing the stretch vibration band of Si–O–Ti [15], indicate

the successful coating of APTS onto the nano-TiO2 through chemical bonding. The peak at 755–620

cm-1 represents the Ti–O–Ti [15]. The peak at 1545 cm-1 represents the secondary amide N–H–

bending and C–N stretching, and the peak at 801 cm-1 represents the N–H rocking, furthermore, the

peaks at 1220 cm-1 (C–O) and 1120 cm-1 (C–O–C) disappeared, implying the reaction between the

7
ACCEPTED MANUSCRIPT

epoxide group and amino. All these results above will have import referential value for the study on the

bonding between GO sheets and nano-TiO2.

The XRD patterns of APTS, GO, TiO2–GO hybrids and nano-TiO2 are presented in Fig. 3(A). A

T
IP
sharp peak appears at 2 theta nearly equal to 11.97°, which corresponds to the characteristic diffraction

R
peak of GO [9, 18]. Particularly, the sharp peak reveals that GO sheets exhibit a close-packed layered

SC
and highly ordered structure. For TiO2–GO hybrids, there was an emergence of a broad around at 2

theta nearly equal to 21.7° which imply the formation of structure was transformed by APTS. While a

NU
new hump at 2 theta nearly equal to 10.45° is emerging in the curve of hybrids, which attribute to the
MA
diffraction peak of GO. And the hump suggests that TiO2–GO sheets have been sufficiently disordered

and loosened [9]. More importantly, the d-spacing of nanometer materials can be evaluated in
D

accordance with the Bragg’s law:


TE

(1)
P

where n is the diffraction series, λ is the X-ray wavelength, and d is the d-spacing of nanometer
CE

materials. The calculated value of GO is 0.739 nm, but the TiO2–GO hybrids is 0.846 nm. That is, the
AC

decorating process has changed the closely stacking-structure into more loosened and expanded, and

the produced formation could exert beneficial influence on exfoliation and dispersion behavior.

The thermogravimetric curves of APTS, GO and TiO2–GO hybrids are displayed in Fig. 3(B).

Generally, for GO, APTS and TiO2–GO hybrids, the loss of mass at around 100 °C was ascribed to the

removal of absorbed water. For GO, there was a main loss of 43wt.% at around 200 °C arise from the

decomposition of the labile oxygen functional groups, yielding CO 2, CO, and vapor [27]. For APTS,

there was a main loss of 89wt.% at around 200 °C. The absolute mass losses were 67wt.% and 93.6wt.%

for GO and APTS, respectively. Meanwhile, the TiO2–GO hybrids had small mass losses (12wt.%) at

8
ACCEPTED MANUSCRIPT

around 200 °C, and small final weight losses (38wt.%). This suggested that the graphene oxide was

stabilized by the decorated TiO2 because of the interaction between oxygen-containing groups on the

graphene oxide and f-TiO2 particles, leading to decreased weight loss of graphene oxide. Namely,

T
IP
APTS was favorable for combining with the oxygenation of GO and partially restoring graphitic

R
structures.

SC
Figs. 4(a and b) show the characteristic morphology of GO [2, 27]. When nano-TiO2 was introduced

NU
to GO, the nanocomposite displayed a dispersion of TiO2 particles on the surface of GO, and the

dispersion resembled stars in the sky (Figs. 4(c and d)). Furthermore, the TiO2–GO hybrids still had to
MA
fold morphology, which indicated that the lamella of GO was not destroyed in the decorating process.

Figs. 4(e and f) depicted the HR-TEM images of GO and (b) TiO2–GO, respectively. The GO (Fig. 4e)
D
TE

revealed a multilayered structure which demonstrated its preparation process stacking sheets together,

which was disadvantageous to disperse into polymer as nanofiller. While nano-TiO2 decorating on
P
CE

sheets, as shown in Fig. 4f, TiO2–GO hybrids displayed nanoparticles anchoring on monolayer and

maintaining their two dimensional sheet morphology. According to previous research [28, 29], the
AC

more loosened stacking-structure of hybrids was more likely to obtain an excellent dispersion and

exfoliation than closely multilayered stacking-structure.

3.2. Exfoliation and dispersion of TiO2–GO hybrids in epoxy coatings

The dispersion state of GO, nano-TiO2 and TiO2–GO hybrids in cured epoxy resins were studied by

XRD, which was an important tool for determining the exfoliation of GO sheets in the polymer

composites [10]. Fig. 5 showed the XRD patterns of (a) pure epoxy, (b) GO/epoxy, (c)

nano-TiO2/epoxy and TiO2–GO/epoxy. In case of pure epoxy, a wide diffraction from 5 to 30° (Fig.

5(a)) was caused by the scattering of cured epoxy molecules [10, 13], implying its amorphous nature.

9
ACCEPTED MANUSCRIPT

However, after GO, nano-TiO2 and TiO2–GO were incorporated into the epoxy matrix, then all the

epoxy composites, particularly TiO2–GO/epoxy, exhibited similar diffraction patterns as the pure

epoxy, indicating that TiO2–GO hybrids existed an excellent exfoliation and dispersion in the epoxy

T
IP
matrix.

R
In order to evaluate the exfoliation and dispersion of GO, nano-TiO2 and TiO2–GO sheet hybrids in

SC
the matrix, SEM images of cured epoxy composites were also carried out. Figs. 6(a)–6(d) display the

fracture morphology of pure epoxy, 2wt.% GO/epoxy, 2wt.% nano-TiO2/epoxy, and 2wt.%

NU
TiO2–GO/epoxy, respectively. Fig. 6(d) reveals that TiO2–GO had a better exfoliation and dispersion
MA
performance in epoxy resin at 2wt.%, resulting from sheets have been divided to form thin layers.

However, for the 2wt.% GO/epoxy and 2wt.% nano-TiO2/epoxy samples, those nanofillers dispersed
D

inhomogeneously (Fig. 6(b) and (c)). GO dispersed in epoxy have more thicker than as-prepared sheets,
TE

as shown in Fig. 6(b), which implied the laminar nanofiller existing a poor exfoliation performance in
P

epoxy resin. Nano-TiO2 expresses a large-scale aggregation in Fig. 6(c), which corresponded to the
CE

inherent worst dispersion performance as the filler for epoxy resin. Moreover, the superior exfoliation
AC

and dispersion performance of TiO2–GO hybrids demonstrated that the composite process improved

the exfoliation and dispersion of the nanofiller in epoxy resin. The excellent exfoliation and dispersion

could enhance the anti-corrosion property of coatings, and this evolvement can be verified in further

measurements.

3.3. Corrosion resistance of coatings

The electrochemical impedance technique was employed to evaluate and to compare the corrosion

resistance of composite coatings. In addition, the dry film thickness of coatings was nearly equal to 40μm.

As illustrated in Figs. 7(c) and 7(e), the Nyquist impedance of the sample obviously decreased with the

10
ACCEPTED MANUSCRIPT

extension of the immersion time, and this trend was consistent with the corrosion process of coating.

Actually, the process consisted of three stages (the initial soaking, mid-soaking, and terminal soaking)

[19, 25, 30]. In the initial soaking stage, the electrolyte solution penetrated into the coating through

T
IP
inherent micro-pore, but did not reach the surface of the steel sheet. In the mid-soaking stage, the

R
electrolyte solution reached the surface of the steel, and the corrosion reaction occurred. In the terminal

SC
soaking stage, the corrosion products accumulated at metal/coating interface (diffusion process).

Obviously, the anti-corrosion performance of the coating was better with longer time which was long

NU
from the initial stage to the mid-soaking, and the adjudication of soaking-stage played a key role in
MA
research corrosion process.

Equivalent electric circuit is generally used to interpret the EIS data for adjudication the stage. The
D

EIS results were then fitted to the different equivalent circuit models by analyzing the Nyquist plots,
TE

using the ZSimpwin software as shown in Fig. 7(a) and (b). The fitted results represent initial soaking
P

stage and terminal, respectively. The components represent electrolyte resistance, Rs, coating
CE

capacitance, CPEf, coating resistance, Rf, charge transfer resistance, Rct, double layer capacitance, and
AC

CPEdl.

As for the change of impedance in the organic coating failure process, the impedance data have been

simulated by ZSimpWin, and the simulation revealed the process. Fig. 7(c) discloses the impedance

spectra of the composite coatings (2wt.% GO/epoxy, 2wt.% nano-TiO2/epoxy, and 2wt.%

TiO2–GO/epoxy) and the pure epoxy after soaking for 2 h. Fig. 7(e) displays the impedance spectra of

the composite coatings and pure epoxy, which were upheld after soaking for 18, 44, 68, and 90 h,

respectively. The result in Fig. 7(c) indicates that the impedance of coating conformed to the organic

coating at the initial soaking, and the data in Fig. 7(e) is consistent with the model at the mid-soaking

11
ACCEPTED MANUSCRIPT

soaking [22, 30-32]. Thus, the transition time was 16 h for pure epoxy, and 42, 66, and 90 h for 2wt.%

GO/epoxy, 2wt.% nano-TiO2/epoxy, and 2wt.% TiO2–GO/epoxy, respectively. It is verified that the

TiO2–GO nanocomposites effectively enhance the anti-corrosion performance of epoxy. Beyond that,

T
IP
TiO2–GO sheet hybrids have better enhancement than GO and nano-TiO2, owing to well exfoliation

R
and dispersion, and excellent blocking property for micro-pores.

SC
Results of WCA testing are shown in Fig. 8. Figs. 8(a′)–8(d′) indicate that the surface properties of

the coatings were transformed by nanofillers. The typical epoxy coating surface in Fig. 8(a′) presents

NU
hydrophilicity (87.5° ± 1°). When GO, nano-TiO2, and TiO2–GO hybrids were dispersed in epoxy
MA
resin, WCA increased (88.9° ± 1° (Fig. 8(b′)), 95.5° ± 1° (Fig. 8(c′)), and 96.2° ± 1° (Fig. 8(d′)),

respectively). Comparing Fig. 8(a′) with Fig. 8(c′), the nano-TiO2 turned the hydrophilic surface of
D

epoxy coating into a hydrophobic surface (from 87.5° ± 1° to 95.5° ± 1°). Comparing Fig. 8(c′) with
TE

Fig. 8(d′), TiO2–GO hybrids had a stronger influence than nano-TiO2. The larger contact angle
P

responded with better corrosion resistance, which means that the hydrophobic surface reduced the
CE

corrosive medium (electrolyte solution) contacting with the organic coating to protect the steel [31, 33].
AC

And hydrophobic surface can signally extend the initial soaking stage due to hydrophobic property

increase the wettability time of coatings’ surface. After soaking for 20 days, the WCA of the four

anticorrosive coatings decreased obviously (67.9° ± 1° (Fig. 8(A′)), 84.8° ± 1° (Fig. 8(B′)), 73.6° ± 1°

(Fig. 8(C′)), and 89.3° ± 1° (Fig. 8(D′)), respectively) because of the dissociation of the coatings. If

coatings were destroyed, surface’ property turn into more hydrophilia, because most of the electrolyte

generated corrosion products are hydrophile. Nonetheless, the WCA of 2wt.% TiO2–GO/epoxy was

still at a maximum (Fig. 8(D′)), which reflected that the coating had a more better anti-corrosion

12
ACCEPTED MANUSCRIPT

performance among the four coatings. The result was also sustained using the monitor of coatings’

morphologies in soaking process.

Figs. 8(a)–8(d) show the SEM images of the four anticorrosive coatings before soaking, and Figs.

T
IP
8(A)–8(D) show those after soaking for 20 days. Figs. 8(a″)–8(d″) show the surface of electrode (the

R
four coated-steel) before soaking, and Figs. 8(A″)–8(D″) show those after soaking for 20 days. The

SC
morphology of pure epoxy before soaking is illustrated in Fig. 8(a). Fig. 8(A) shows the corresponding

morphology after soaking for 20 days. Fig. 8(A) displays an abundant etch-pit phenomenon, which

NU
generated the serious resin wear on the electrode surface, as shown in Fig. 8(A″). The morphology of
MA
the 2wt.% GO/epoxy before soaking is exhibited in Fig. 8(b), and Fig. 8(B) shows the corresponding

morphology after soaking for 20 days. Fig. 8(B) displays a considerable peeling morphology, which
D

lead to the biggish decrustation on the electrode surface, as shown in Fig. 8(B″). Fig. 8(c) displays the
TE

morphology of the 2wt.% nano-TiO2 /epoxy before soaking, and Fig. 8(C) shows the corresponding
P

morphology after soaking for 20 days. The disintegrating phenomenon is displayed in Fig. 8(C), which
CE

produced blistering on the electrode surface, as shown in Fig. 8(C″). Fig. 8(d) reveals the morphology
AC

of the 2wt.% TiO2–GO /epoxy before soaking, and Fig. 8(D) shows the corresponding morphology

after soaking for 20 days. Comparison Fig. 8(d) and (d″) with Fig. 8(D) and (D″), the result reveals that

there was an inconspicuous transformation on the electrode surface. Obviously, the morphological

transformation of coatings was associated with the anti-corrosion performance, and the lesser change

reflected the better. The comparison between the pure epoxy, 2wt.% GO/epoxy, and 2wt.% nano-TiO2

/epoxy and the 2wt.% TiO2–GO/epoxy explicitly reflects that the 2wt.% TiO2–GO/epoxy had a better

anti-corrosion performance than the other three coatings, and this conclusion also was suggested by the

electrochemical monitoring and WCA testing.

13
ACCEPTED MANUSCRIPT

According to above mentioned data, it is not difficult find that the TiO2–GO hybrids have

effectively enhanced the anti-corrosion performance for epoxy coatings. The enhancement arises from

that the sheet structure provide an extra barrier layer to preeminently obstructed micro-pores for

T
IP
electrolyte permeation lead to prevent the underlying metals from corrosion attack, the contribution of

R
TiO2–GO sheet hybrids as shown in Fig. 9, and its precondition is excellent exfoliation and dispersion.

SC
4. Conclusions

NU
We have synthesized fine titanium dioxide nanoparticles-decorated graphene oxide nanocomposites
MA
via a simple two-step solution approach without using any extra templates or surfactants. The sheet

structure of the TiO2–GO hybrids is influenced obviously by decorating nano-TiO2 on GO surface, and
D

greater interlayer spacing is obtained. Moreover, the hybrids not only have excellent exfoliation and
TE

dispersion in epoxy resin, but also obviously enhance epoxy coatings’ corrosion resistance at a low
P

content (2wt.%). Comparison with other nanofillers reported in this literature indicated that TiO2–GO
CE

hybrids displayed an obvious advantage in enhancement corrosion resistance. There are several reasons
AC

for superiority of the hybrids, which include their high-efficiency plugging inherent micro-pores, sheets

structure, and the excellent exfoliation and dispersion in epoxy resin. Herein, the TiO2–GO hybrids

presented a proactive application in the field of the nanofillers for anticorrosive epoxy coatings.

Acknowledgement

This work was supported by the applied basic research plan of Sichuan Province (No: V200801).

References

[1] Ž. Šljivančanin, A.S. Milošević, Z.S. Popović, F.R. Vukajlović, Carbon, 54 (2013) 482-488.
[2] C. Botas, P. Álvarez, P. Blanco, M. Granda, C. Blanco, R. Santamaría, L.J. Romasanta, R. Verdejo,
M.A. López-Manchado, R. Menéndez, Carbon, 65 (2013) 156-164.
[3] G. Srinivas, Y. Zhu, R. Piner, N. Skipper, M. Ellerby, R. Ruoff, Carbon, 48 (2010) 630-635.

14
ACCEPTED MANUSCRIPT

[4] Y. Li, W. Gao, L. Ci, C. Wang, P.M. Ajayan, Carbon, 48 (2010) 1124-1130.
[5] F. He, J. Fan, D. Ma, L. Zhang, C. Leung, H.L. Chan, Carbon, 48 (2010) 3139-3144.
[6] L. Chen, S. Chai, K. Liu, N. Ning, J. Gao, Q. Liu, F. Chen, Q. Fu, ACS Appl Mater Interfaces, 4
(2012) 4398-4404.
[7] S. Navalon, A. Dhakshinamoorthy, M. Alvaro, H. Garcia, Chemical reviews, 114 (2014)

T
6179-6212.
[8] H. Yang, C. Shan, F. Li, Q. Zhang, D. Han, L. Niu, Journal of Materials Chemistry, 19 (2009)

IP
8856.
[9] R.K. Layek, A.K. Das, M.U. Park, N.H. Kim, J.H. Lee, Journal of Materials Chemistry A, 2 (2014)

R
12158.

SC
[10] L.-Z. Guan, Y.-J. Wan, L.-X. Gong, D. Yan, L.-C. Tang, L.-B. Wu, J.-X. Jiang, G.-Q. Lai, Journal
of Materials Chemistry A, 2 (2014) 15058.
[11] L. Cao, X. Liu, H. Na, Y. Wu, W. Zheng, J. Zhu, Journal of Materials Chemistry A, 1 (2013)

NU
5081.
[12] X.-J. Shen, X.-Q. Pei, S.-Y. Fu, K. Friedrich, Polymer, 54 (2013) 1234-1242.
[13] Y.-J. Wan, L.-X. Gong, L.-C. Tang, L.-B. Wu, J.-X. Jiang, Composites Part A: Applied Science
MA
and Manufacturing, 64 (2014) 79-89.
[14] R. Qian, J. Yu, C. Wu, X. Zhai, P. Jiang, RSC Advances, 3 (2013) 17373.
[15] C.X. Wang, H.Y. Mao, C.X. Wang, S.H. Fu, Ind. Eng. Chem. Res., 50 (2011) 11930-11934.
[16] G. Jiang, Z. Lin, C. Chen, L. Zhu, Q. Chang, N. Wang, W. Wei, H. Tang, Carbon, 49 (2011)
D

2693-2701.
TE

[17] L. Shao, S. Quan, Y. Liu, Z. Guo, Z. Wang, Materials Letters, 107 (2013) 307-310.
[18] N. Yusoff, N.M. Huang, M.R. Muhamad, S.V. Kumar, H.N. Lim, I. Harrison, Materials Letters,
93 (2013) 393-396.
P

[19] E. Bakhshandeh, A. Jannesari, Z. Ranjbar, S. Sobhani, M.R. Saeb, Progress in Organic Coatings,
CE

77 (2014) 1169-1183.
[20] I. Zaman, T.T. Phan, H.C. Kuan, Q.S. Meng, L.T.B. La, L. Luong, O. Youssf, J. Ma, Polymer, 52
(2011) 1603-1611.
AC

[21] Y.-J. Wan, L.-C. Tang, L.-X. Gong, D. Yan, Y.-B. Li, L.-B. Wu, J.-X. Jiang, G.-Q. Lai, Carbon,
69 (2014) 467-480.
[22] H. Yi, C.L. Chen, F. Zhong, Z.H. Xu, High Performance Polymers, 26 (2014) 255-264.
[23] N.M. Huang, H.N. Lim, C.H. Chia, M.A. Yarmo, M.R. Muhamad, Int. J. Nanomed., 6 (2011)
3443-3448.
[24] J. Kim, H.W. Park, Corrosion Science, 90 (2015) 153-160.
[25] M. Behzadnasab, S.M. Mirabedini, K. Kabiri, S. Jamali, Corrosion Science, 53 (2011) 89-98.
[26] Y. Lin, J. Jin, M. Song, Journal of Materials Chemistry, 21 (2011) 3455.
[27] M.S. Ahmed, H.S. Han, S. Jeon, Carbon, 61 (2013) 164-172.
[28] K.-S. Kim, I.-Y. Jeon, S.-N. Ahn, Y.-D. Kwon, J.-B. Baek, Journal of Materials Chemistry, 21
(2011) 7337.
[29] M.M. Gudarzi, F. Sharif, Journal of colloid and interface science, 366 (2012) 44-50.
[30] C. Zhu, R. Xie, J. Xue, L. Song, Electrochimica Acta, 56 (2011) 5828-5835.
[31] S. Shreepathi, S.M. Naik, M.R. Vattipalli, Journal of Coatings Technology and Research, 9 (2011)
411-422.
[32] L. Jianguo, G. Gaoping, Y. Chuanwei, Electrochimica Acta, 50 (2005) 3320-3332.

15
ACCEPTED MANUSCRIPT

[33] K.-C. Chang, M.-H. Hsu, H.-I. Lu, M.-C. Lai, P.-J. Liu, C.-H. Hsu, W.-F. Ji, T.-L. Chuang, Y.
Wei, J.-M. Yeh, W.-R. Liu, Carbon, 66 (2014) 144-153.

T
R IP
SC
NU
MA
D
P TE
CE
AC

16
ACCEPTED MANUSCRIPT

Fig. 1. Schematic of the synthesis procedure of TiO2–GO hybrids.

Fig. 2. FT-IR spectra of GO (a) and TiO2–GO hybrids (b).

Fig. 3. (A) XRD patterns of APTS, GO, TiO2–GO hybrids and nano-TiO2. (B) TGA curves of APTS,
GO and TiO2–GO.

Fig. 4. SEM images of GO (a, b) and TiO2–GO hybrids (c, d). TEM images of (e) GO, (f) TiO2–GO

T
hybrids.

IP
Fig. 5. XRD patterns of (a) pure epoxy, (b) GO/epoxy, (c) nano-TiO2/epoxy, and (d) TiO2–GO/epoxy.

Fig. 6. Fracture surface of pure epoxy (a), 2 wt.% GO/epoxy (b), 2 wt.% nano-TiO2 /epoxy (c) and 2

R
wt.% TiO2–GO/epoxy (d).

SC
Fig. 7. Equivalent circuit models used for numerical simulation of the EIS measurements of the
coatings at different soaking stages (a, b). EIS diagrams obtained from epoxy coating containing same
mass fractions of GO, nano-TiO2, and TiO2–GO. (c) Nyquist diagram of coatings after soaking for 2 h.
(d) Bode diagram of coatings after soaking for 2 h. (e) Nyquist diagram of coatings after 18, 44, 68,

NU
and 90 h (corresponding pure epoxy, 2wt.% GO/epoxy, 2wt.% nano-TiO2/epoxy, and 2wt.%
TiO2–GO/epoxy, respectively). (f) Bode diagram of coatings after soaking for 18, 44, 68, and 90 h,
respectively.

Fig. 8. SEM images of electrodes about pure epoxy (a), 2wt.% GO/epoxy (b), 2wt.% nano-TiO2/epoxy
MA
(c) and 2wt.% TiO2–GO/epoxy (d), before soaking. (A), (B), (C), and (D) correspond to coatings after
soaking for 20 days. WCA of electrodes about pure epoxy (a′), 2wt.% GO/epoxy (b′), 2wt.%
nano-TiO2/epoxy (c′) and 2wt.% TiO2–GO/epoxy (d′), before soaking. (A′), (B′), (C′), and (D′)
correspond to coatings after soaking for 20 days. The surface of electrodes about pure epoxy (a″), 2wt.%
GO/epoxy (b″), 2wt.% nano-TiO2/epoxy (c″) and 2wt.% TiO2–GO/epoxy (d″), before soaking. (A″),
(B″), (C″), and (D″) correspond to coatings after soaking for 20 days.
D

Fig. 9. Schematic representation of the corrosion inhibition properties of TiO 2–GO/ epoxy.
P TE
CE
AC

17
ACCEPTED MANUSCRIPT

T
IP
R
SC
NU
MA
D
TE

Fig. 1
P
CE
AC

18
ACCEPTED MANUSCRIPT

Ti-O-Ti(722~634)
-OH,N-H(3414) Ti-O-Si(925)
(b)

-NH2(1653) -NH(801)
-NH,-CN(1545)

T
-CH(2928,2859) Si-O-C(1035)
C=C(1620)
Si-O-Si(1115)

IP
B

C-O-C(1120)
(a) -OH(3447)

R
SC
=C-H(710)
C=O(1741) C-O(1220)
C=C(1625)
=C-H(1400)

NU
4000 3750 3500 3250 3000 2750 2500 2250 2000 1750 1500 1250 1000 750 500
-1
Wave number (cm )
MA
Fig. 2
D
P TE
CE
AC

19
ACCEPTED MANUSCRIPT

A B APTS

100 GO
TiO2GO
TiO2 90 12%

80

T

10.45  0.846 nm

70
21.7
38%

IP
TiO2GO
Intensity(a.u.)

60

Mass (%)
43%


50

R
11.97 , 0.739 nm

40
67%

SC
GO 30

20

21.6
10

NU
89%
APTS 93.6%
0
10 20 30 40 50 60 70 80 0 100 200 300 400 500 600 700 800

2/(deg.) o
Temperature ( C)
MA
Fig. 3
D
P TE
CE
AC

20
ACCEPTED MANUSCRIPT

T
IP
R
SC
NU
MA
D
PTE
CE
AC

Fig. 4

21
ACCEPTED MANUSCRIPT

(a)

T
(b)

R IP
(c)

SC
(d)

NU
5 10 15 20 25 30 35

2/(deg.)
MA
Fig. 5
D
P TE
CE
AC

22
ACCEPTED MANUSCRIPT

T
IP
R
SC
NU
MA
Fig. 6
D
PTE
CE
AC

23
ACCEPTED MANUSCRIPT

pure epoxy
2wt.% TiO2/epoxy
2wt.% GO/epoxy
2wt.% TiO2-GO/epoxy

fitted
1000000 70
400000
65

350000 60
100000

Phase angle (Degree)


55
300000

T
(c) 50

logZ cm 
Zim cm 

2
2

250000 45
10000

IP
200000 40

S
CPEf 35
150000
1000 30
Rs
100000 25

R
Rct (d) 20
50000 100
15
0

SC
(a) Equivalent circuit models for the initial soaking 10
-50000 10 5
0 200000 400000 600000 800000 0.1 1 10 100 1000 10000
2 log f (Hz)
Zre (cm )

1000000 55

NU
14000
50

12000 100000 45

Phase angle (Degree)


(e) logZ cm 
2
40
10000

CPEf 10000 35
Zim cm 
2

8000
MA
30
6000
Rs
1000 25
CPEdl
4000 20
Rf
(f)
2000 Rct 100 15

10
0 (b)
D

Equivalent circuit models for the mid-soaking soaking


10 5
0 5000 10000 15000 20000 25000 30000 35000 0.1 1 10 100 1000 10000

Zre (cm )
2 log f (Hz)
TE

Fig. 7
P
CE
AC

24
ACCEPTED MANUSCRIPT

T
IP
R
SC
NU
MA
D
PTE
CE
AC

Fig. 8

25
ACCEPTED MANUSCRIPT

T
IP
R
SC
NU
MA
Fig. 9
D
PTE
CE
AC

26
ACCEPTED MANUSCRIPT

Highlights

1. TiO2-GO sheets material was synthesized through a simple method.

2. When TiO2 loaded on GO sheets, the nanocomposites remained lamelliform.


3. TiO2-GO had a well dispersing performance for epoxy resin.

T
4. TiO2-GO dramatically improved the anti-corrosion performance of epoxy coatings.

R IP
SC
NU
MA
D
P TE
CE
AC

27

You might also like