Você está na página 1de 15

Progress in Organic Coatings 77 (2014) 1169–1183

Contents lists available at ScienceDirect

Progress in Organic Coatings


journal homepage: www.elsevier.com/locate/porgcoat

Anti-corrosion hybrid coatings based on epoxy–silica


nano-composites: Toward relationship between the morphology
and EIS data
Ehsan Bakhshandeh a,∗ , Ali Jannesari a , Zahra Ranjbar b , Sarah Sobhani a ,
Mohammad Reza Saeb a
a
Department of Resin and Additives, Institute for Color Science and Technology, Tehran, Iran
b
Department of Surface Coating and Corrosion, Institute for Color Science and Technology, Tehran, Iran

a r t i c l e i n f o a b s t r a c t

Article history: This work reports on design and manufacture of organic–inorganic hybrid coatings based on diglycidyl
Received 19 May 2013 ether of bisphenol A (DGEBA) epoxy resin pursuing hydrolyzation of tetraethoxysilane (TEOS) through a
Received in revised form 17 March 2014 sol–gel process. The resulting hybrid materials were cured to be used as potential anticorrosive coatings.
Accepted 3 April 2014
The assigned materials were modified molecules made of DGEBA and 3-aminopropyl triethoxylsilane
(APTES), in which the molar ratio of epoxide group of DGEBA to NH of APTES varied in the order of
Keywords:
2:1, 4:1, 8:1 and 16:1. In the next stage, the APTES-modified DGEBA precursors were added to different
Epoxy–silica
amounts of pre-hydrolyzed TEOS, i.e. 7.5, 12.5 and 17.5 wt%, as inorganic part of the resulting hybrid. The
Sol–gel
Hybrid coating mixtures were subsequently cured at room temperature by a cycloaliphatic amine based curing agent
Morphology to yield transparent epoxy–silica hybrid coatings. Microstructure assessment of the hybrid materials,
Electrochemical impedance spectroscopy before and after curing, was performed using FTIR and 29 Si NMR spectroscopies. The morphology of the
(EIS) epoxy–silica hybrid coatings has also been studied by scanning electron microscopy (SEM). The anti-
corrosive measurements on the resultant coatings were conducted based on electrochemical impedance
spectroscopy (EIS). The mechanical properties evaluation such as micro-hardness measurements and
pull-off adhesion tests of the cured samples were also carried out. The thermal properties of the cured
hybrid coatings were evaluated using thermogravimetric analysis (TGA). The results showed that the
concentration of APTES and pre-hydrolyzed TEOS play an important role in determining the morphology
as well as the mechanical and thermal properties of coatings. The EIS results corresponding to these effects
reaffirmed that the corrosion resistance of the hybrid coatings improved with increasing the inorganic
phase content.
© 2014 Elsevier B.V. All rights reserved.

1. Introduction pretreatment coating systems for the protection of metal sub-


strates with outstanding characteristics compared to their organic
Corrosion is a phenomenon through which a metal deteriorates counterparts [5–11]. These materials, denoted as organic–inorganic
as a consequence of its interaction with the surrounding envi- hybrids, typically contain co-continuous domains having dimen-
ronment that helps for an electrochemical reaction via oxidation. sions ranging in 5–100 nm [12–17]. In this regard, a wide variety
To protect or slow down the corrosion process, metals are often of applications were demonstrated for multifarious fields, e.g. rub-
coated to yield a protective barrier against corrosive environment bers, plastics, sealants, fibers, optical materials, medicals, and high
[1–4]. In recent years, explore and development of hybrid coatings thermal resistance materials [18–26].
grabbed the attention of many researchers and engineers providing The hybrid coatings are generally prepared by low-temperature
them the opportunity of designing appropriate primers or in situ sol–gel processes through in situ hydrolysis followed by conden-
sation of organometallic precursors like silicates, titanates and
aluminates, in an organic matrix [27]. Careful selection of a hybrid
∗ Corresponding author. Tel.: +98 21 22956209; fax: +98 21 22947537. coating allows combining the desirable properties of organic part
E-mail addresses: bakhshandeh-e@icrc.ac.ir, bakhshandehehsan@yahoo.com of system, i.e. toughness and elasticity with those of inorganic
(E. Bakhshandeh). phase that is characteristic of good hardness, chemical resistance,

http://dx.doi.org/10.1016/j.porgcoat.2014.04.005
0300-9440/© 2014 Elsevier B.V. All rights reserved.
1170 E. Bakhshandeh et al. / Progress in Organic Coatings 77 (2014) 1169–1183

and adhesion to the metal substrate via formation of covalent this way, EIS values are correlated this morphological features of
bonds [28–30]. Frings et al. [31] synthesized hybrid coatings based the resulting coatings.
on polyesters and tetraethoxysilane (TEOS) to meet protective
coatings exhibiting performance enhancement in prefinished steel
2. Materials and methods
and aluminum constructions. They have shown that increasing the
TEOS content in the coatings causes an increase in the hardness
2.1. Materials
and glass transition temperature (Tg ). Cho and Lee [32] stud-
ied the influence of in situ generated silica obtained from TEOS
A liquid DGEBA epoxy resin having EEW of 191 g/equiv. was
and found that the mechanical properties of polyurethane based
obtained from KZPC (Iran) and used as the organic phase precursor.
coatings were largely governed by both the presence and con-
The inorganic part of hybrid system, i.e. TEOS, was purchased from
centration of TEOS within the hybrid. Sailer and Soucek reported
Fluka (USA). APTES coupling agent as well as tetrahydrofuran (THF)
on dependency of thermo-mechanical properties on the type and
and ethanol solvents were all obtained from Merck (Germany).
amount of inorganic part in a series of hybrid coatings prepared
Hydrochloric acid (37 wt% in water) pH regulator was purchased
in the presence of an oxidizing alkyd [33]. It was also elucidated
from Aldrich (USA). A cycloaliphatic amine curing agent, known as
elsewhere that hybrid coatings based on polyurea/polysiloxane
Epikure F205, was obtained from Hexion (USA). Hydrogen equiv-
reveal higher adhesion to the metal substrate after addition of pre-
alent weight of this hardener was 102–104 g/mol, viscosity in the
hydrolyzed TEOS into the coatings [34]. Among family of hybrid
range of 500–700 mPa s, according to supplier. All other reagents
materials, epoxy–silica composites were examined by a num-
used in modification process were used as received.
ber of researchers [35–38]. In nature, epoxy resins are famous
because of their excellent properties, e.g. superior chemical and
corrosion resistance, good adhesion, and cureability at ambient 2.2. Instrumentation
temperature. Nevertheless, epoxy resin suffers from inferior ther-
mal and weathering stabilities, poor mechanical properties, limited Infrared absorption spectra were collected on a Perkin-Elmer
recoating times, unsuitable cutting and welding properties. This Spectrum One spectrometer. For carrying out this test, hybrid cured
puts limits on the selection of this resin for high-performance samples were grinded and amalgamated with dry IR-grade KBr in
applications [36–38]. To resolve this situation, it is suggested to a mortar to form pellets. Bruker DRX 500 Avance spectrometer
incorporate TEOS inorganic precursor into the epoxy binder to was used to record 29 Si NMR. The morphology of the samples was
achieve epoxy–silica hybrid materials [39,40]. There are sufficient observed using a Philips XL30 scanning electron microscope (SEM).
evidences for creation of a phase-separated silica-rich network rep- The turbidity of the mixtures containing modified DGEBA and pre-
resenting poor compatibility with epoxy resin while using TEOS. hydrolyzed TEOS was determined by 2100 AN Turbidimeter HACH.
In this context, the literature addresses different investigations The haziness of the cured hybrid films was measured by CE-7000A
to meet compatibility in the epoxy–silica hybrids systems via spectrophotometer (ASTM 1003). Rheological behavior of silane-
sol–gel process [13–16,39,40]. It is a typical that as the compati- functionalized epoxy resins was investigated utilizing MCR 300
bility between the organic and inorganic parts in a hybrid system Anton Paar. The micro-hardness was conducted on a HX-100 Vick-
increases, the solubility of the inorganic precursor in the organic ers Micro-hardness in accordance with ASTM-E384. The test output
matrix increases. Researchers have mostly been attempting to find was reported taking average of hardness of 3–5 points for each coat-
and examine different strategies for compatibilization assessment ing. In practice, it was a difficult task identifying diagonal indention
in the epoxy–silica system. Hence, they suggested functionaliza- of various hard brittle coatings due to the fractures around the
tion of epoxy by functional trialkoxysilane [35,36] or simultaneous indention point. Pull-off adhesion test was carried out utilizing
addition of this component with a coupling agent to guarantee Defelsko positest AT (ASTM-D4541), where the dollys, 20 mm in
chemical or strong physical interaction between epoxy and sil- diameter, were glued firmly to the surface of the coated panels
ica domains [39,40]. Though epoxy–silica hybrid coatings were using a cyano-acrylate adhesive. After complete curing of the adhe-
deemed to exhibit anti-corrosive potential, there is no direct report sive, the fixture was loaded and the joint was strained at a constant
aimed at this target. The idea of preparation of silane-modified rate of 5 mm/min using the pull-off testing equipment, until a plug
epoxy resins from pre-hydrolyzed TEOS was consequently devel- of the coating material was detached from the substrate surface.
oped and examined in this context. Since TEOS in the form of For each test sample, five replicas were employed and the average
monomer has a very low vapor pressure of about 1.5 mmHg at value reported. Electrochemical impedance spectroscopy (EIS) was
ambient temperature, therefore, it can easily evaporate during carried out in a three-electrode system made by Ivium Compactstat
room-temperature curing of the hybrid system. To prevent evap- (Netherlands) equipped with Ivium Equivalent Circuit Evaluator
oration of TEOS, we used its pre-hydrolyzed form and changed software, which enabled for determining the equivalent circuit and
the weight ratio of 3-aminopropyl triethoxysilane (APTES) to func- data analyzing. During EIS test, the frequency range was varies
tionalized epoxy resin based on diglycidyl ether of bisphenol A between 100 kHz and 10 mHz, while the perturbation voltage was
(DGEBA). The silane-modified epoxy resins were characterized by set to 10 mV. The saturated calomel electrode and graphite rod were
means of Fourier transform infrared (FTIR) and Epoxy Equiva- used as reference electrode and auxiliary electrode, respectively.
lent Weight (EEW) measurements. Three different concentrations The fabricated coatings were applied to substrates made of steel (ST
of pre-hydrolyzed TEOS as inorganic phase were added to the 37). Before casting, steel plates were cleaned by acetone and dried
APTES-modified DGEBA reactive agent. The mixtures were then to remove any kind of grease. The substrates were subsequently
cured at room temperature by a cycloaliphatic amine based cur- rinsed with distilled water to avoid contamination. In this manner,
ing agent to obtain transparent epoxy–silica hybrid coatings. The approximately 1 cm2 area of epoxy–silica hybrids with a thickness
chemical structure and morphology of the cured coatings were of 140 ±5 ␮m casted on metal and exposed to 3.5% NaCl electrolyte.
characterized using FTIR and scanning electron microscopy (SEM) The rest of surface was covered with a 75/25 beeswax/colophony
methods. Besides, mechanical properties of the coatings such as mixture. Thermal degradation of the coatings was investigated with
micro-hardness and adhesion strength were evaluated. Eventually, a Perkin-Elmer 6 thermogravimetric analyzer (TGA) at a heating
electrochemical impedance spectroscopy (EIS) and thermogravi- rate of 10 ◦ C/min. The specimens of about 6–10 mg were heated
metric analysis (TGA) were used to assess the corrosion resistance from room temperature to 600 ◦ C under nitrogen atmosphere and
and thermal behavior of the organic–inorganic hybrid coatings. In temperature-dependent weight loss was recorded.
E. Bakhshandeh et al. / Progress in Organic Coatings 77 (2014) 1169–1183 1171

Table 1
Experiments data of pull-off adhesion test for studied epoxy–silica hybrid coatings.

Samples Epoxy–silica Solvent (THF wt%) Pull-off adhesion


(psi)(±0.5)

Silane functionalized DGEBA wt% TEOS pre-hydrolyzed wt%

E16T7.5 92.5 (E16) 7.5 75 3.33


E16T12.5 87.5 (E16) 12.5 75 4.02
E16T17.5 82.5 (E16) 17.5 75 2.71
E8T7.5 92.5 (E8) 7.5 75 3.54
E8T12.5 87.5 (E8) 12.5 75 4.06
E8T17.5 82.5 (E8) 17.5 75 2.81
E4T7.5 92.5 (E4) 7.5 75 3.01
E4T12.5 87.5 (E4) 12.5 75 3.89
E4T17.5 82.5 (E4) 17.5 75 2.23
Pure epoxy 100 (unmodified epoxy) 0 75 5.20

2.3. Preparation of pre-hydrolyzed TEOS where WEP is the amount of the functionalized epoxy resin, A the
amine value of the curing agent (104 g/equiv.). Before incorporation
The preparation of pre-hydrolyzed TEOS precursor is achieved of curing agent, the turbidity of the samples was determined. The
via a stepwise procedure. First, TEOS (208.33 g, 1 mol) was dissolved turbid-to-clear transition of the mixtures took place 4 h after addi-
in ethanol (176.62 g, 3.84 mol) and distilled water (9 g, 0.5 mol) was tion of HCl solution. The FTIR, SEM, and haziness measurements
then added into the mixture under mechanical stirring. After the were employed for characterization of produced hybrid materials.
water dissolved, hydrochloric acid (0.001 molar) was added drop
wise and the reaction continued at 23 ◦ C for 48 h. It is no denying
3. Results and discussion
that pH is a key in controlling gel structure. It is documented that
an acidic media will speed up the hydrolysis reaction more effi-
3.1. Epoxy modification assessment
ciently compared to condensation reaction, for the latter involves
the attack of silicon atoms carrying protonated silanol species by
The functionalization of epoxy resins with APTES was used as
neutral Si OH nucleophiles [24]. The last step was to removal of
a route for compatibilization and control of phase separation phe-
the solvent at 1 mmHg and 50 ◦ C under vacuum to gain 144.58 g
nomenon in the epoxy–silica hybrid system. The epoxy resin has
pre-hydrolyzed TEOS. The yield of reaction based on TEOS and was
been to a specified extent functionalized to generate an epoxy sys-
estimated to be 69.4% [24]. The resulting precursor was character-
tem with the potential to cure with cycloaliphatic amine used in
ized by 29 Si NMR.
this work. Thus, some samples revealed insufficient compatibility
as a consequence of improper curing under stoichiometric ratio.
The modification reaction with the silane component was carried
2.4. Preparation of epoxy–silica hybrid coatings out varying molar ratio of epoxide groups to active hydrogen atoms
of amine group of APTES, as given in Table 2. By choosing different
A 25 wt% solution of silane-functionalized epoxy in THF was stoichiometric ratios of epoxide to amine groups; it was possible to
first prepared followed by addition of a mixture of pre-hydrolyzed track both the crosslink ability and applicability of resulting mate-
TEOS in water (3:1 molar ratio of water to TEOS). Organic–inorganic rial as a coating. The chemical interactions between the epoxy resin
hybrid coatings were formulated by addition of different amounts and APTES were studied by FTIR and EEW measurements. Fig. 1
of the pre-hydrolyzed TEOS to the silane-functionalized DGEBA, as shows the absorption spectra of pure APTES, DGEBA, and silane-
given in Table 1. The next step was the addition of HCl solution to modified DGEBA samples. The reaction was supposed to take place
bring the pH in the range of 2–3. The reactants were stirred for about in accord with Scheme 1 suggesting that functionalization of epoxy
6 h until the mixture became clear and then the curing agent was with alkoxysilane takes place through the reaction between the
added. The cycloaliphatic amine curing agent used in stoichiomet- amine groups of APTES and the epoxide groups of DGEBA. The
ric ratio to cure thin films through casting the mixtures on steel characteristic absorption bands for the NH and Si OEt (Et denot-
panels, and the films were conditioned at 25 ◦ C and 50% relative ing Ethyl) bonds of APTES appeared at 3381, 1605 and 1099 cm−1 ,
humidity to achieve cured films containing silica domains after two respectively. Also, the absorption peak at 916 cm−1 relating to
weeks exposure. It was found that dry to touch time for all samples epoxide groups is weakened during the functionalization reaction.
was approximately about 4–7 days from exposure. The stoichio- The presence of the peak of OH group near the 3400 cm−1 makes
metric amounts of curing agent (WCA ) were calculated evaluation evident that epoxide ring opening reaction between DGEBA and
the functionalized epoxy resin, as follow: APTES is occurred. The theoretically calculated and determined
EEW values of functionalized epoxides are listed in Table 2. As can
WEP × A be seen, there is a good agreement between theoretical and exper-
WCA = (1)
EEW imental investigations implying an increase in the values of EEW

Table 2
Experimental and theoretical EEW and turbidity measurements after functionalization of the epoxy resin.

Samples E16 E8 E4 E2

Epoxide groups (in DGEBA) 16 8 4 2


Active Hydrogen atoms (in APTES) 1 1 1 1
Measured EEW (g/mol) 208.2 232.6 288.4 –
Calculated EEW (g/mol) 217.5 240.3 295.1 Gel formation
Turbidity (NTU)a 1.81 1.37 1.03 –
a
Turbidity in NTU of the mixtures of APTES modified samples with 12.5 wt% of pre-hydrolyzed TEOS in functionalization of the epoxy resin.
1172 E. Bakhshandeh et al. / Progress in Organic Coatings 77 (2014) 1169–1183

Scheme 2. The proposed side-reactions taking place during sol–gel process.

seen, there is a viscosity upturn over increasing the concentration


of APTES within the system at the assigned region moving from
sample E16 to E8 and then E4. In fact, formation of silane bonds
in the system is of lesser probability in case of E16. It is obvious
that this system behaves nearly to a Newtonian fluid at the studied
shear rate interval due to shortage of inter- and intra-molecular
interactions. On the other hand, samples containing more APTES
units undergo shear thinning as a consequence of possible destruc-
tive effects at higher shear rates. It is to be noted that owing to very
high content of APTES in the sample E2, a gelled mass was formed
that made rheological measurement impossible.

3.2. Characterization of pre-hydrolyzed TEOS

The preparation of oligomeric precursors based on TEOS through


hydrolysis followed by co-condensation reactions have been the
subject of some investigations, as speculated in Scheme 3 [42–47].
Fig. 3 illustrates 29 Si NMR analysis of pre-hydrolyzed TEOS exhibit-
ing different chemical shifts at −89.98, −95.68 and −96.45 ppm
assigning to terminal group Si O Si (OEt)3 , cyclic ring Si O Si
Fig. 1. FTIR spectra of (A) DGEBA, (B) APTES, (C) E8, (D) E4, and (E) E2. (OEt)2 O Si and linear Si O Si (OEt)2 O Si, respectively. The
resonance splitting is considered to be the superposition of the
estimated for modified specimens compared to that of neat epoxy species with similar structure and different degree of polymeriza-
(191 g/mol). During functionalization of epoxy resin with APTES, tion.
some side reactions might also been possible, which are displayed
in Scheme 2 [27,41]. Thus, depending on resin to APTES molar ratio, 3.3. Compatibility analysis
the contribution of Si O Si bonds to gel formation could to some
extent be probable. When the pre-hydrolyzed TEOS was added to the APTES-
As shown in parts (c) and (d) in Fig. 1, with increasing APTES modified epoxy, a turbid mixture was obtained. Turbidity values
level in the modification reaction the intensity of Si O C bands of the mixtures of 12.5 wt% pre-hydrolyzed TEOS in different
(1080–1100 cm−1 ) decreases, while the one corresponding to
Si O Si bands (near 1030 cm−1 ) is associated with an increase. Vis-
cosity dependence of the silane-modified DGEBA samples is plotted
in Fig. 2. The high viscosity at low-shear rate region can be caused
by linear extent of the epoxy chains reacting with the amine group
of APTES and also by the formation of silica domains. As can be

Fig. 2. Viscosity as a function of shear rate and molar ratio of epoxide group to NH
Scheme 1. Chemical reaction between DGEBA and APTES. group in the APTES: 16/1, 8/1, 4/1.
E. Bakhshandeh et al. / Progress in Organic Coatings 77 (2014) 1169–1183 1173

Scheme 3. Preparation of pre-hydrolyzed TEOS exhibiting different chemical structures during sol–gel reaction.

29
Fig. 3. Si NMR spectra of (A) TEOS monomer and (B) pre-hydrolyzed TEOS.
1174 E. Bakhshandeh et al. / Progress in Organic Coatings 77 (2014) 1169–1183

Fig. 4. Optical micrographs of the mixtures of APTES modified-epoxy samples (A: E4, B: E8, C: E16) with 12.5 wt% pre-hydrolyzed TEOS (scale = 31.76 ␮m).

APTES-modified DGEBA samples, i.e. E16, E8, and E4, were are measured and reported in Fig. 6. As can be seen, the haziness
measured and expressed in nephelometric turbidity units, NTU of the films has increased as a result of a rise in the content of
(Table 2). As can be seen, the turbidity of the mixture has risen over pre-hydrolyzed TEOS. At a given concentration of TEOS; however,
decreasing the amount of APTES in the samples. Optical microscopy the higher APTES content leads to more compatibility between
observations, as represented in Fig. 4, reveal that the size of the dis- organic and inorganic phases followed by formation of smaller silica
persed phase (the pre-hydrolyzed TEOS) in the continuous phase domains, thereby lesser haziness of hybrid films.
(the APTES-modified DGEBA) varies by the APTES level. This reaf- Fig. 7 compares SEM micrographs of the hybrid samples con-
firmed turbidity measurements implying enhanced compatibility taining 12.5 wt% of pre-hydrolyzed TEOS in APTES modified-DGEBA
over APTES content. and the one relating to neat epoxy. Accordingly, the silica particles
smaller than 100 nm in size are uniformly dispersed throughout
3.4. Characterization of hybrid coatings the polymer matrix of the sample E8T12.5 (see Table 1). This
reveals good miscibility between organic and inorganic parts in
3.4.1. Cure assessment the nanocomposites. To make a greater sense of compatibility,
The FTIR analysis shows a very high reactivity between epox- distribution pattern of the silica domains is determined by Pixca-
ide and amine functional groups after two weeks exposure at vator Image Analyzer, as in Fig. 8. It is shown that silica particles
aforementioned condition (Fig. 5). This was evidenced by the dis- formed throughout the matrix have an average size in the range of
appearance of absorption band at around 916 cm−1 corresponding 20–40 nm. This places emphasis on development of a nanostructure
to epoxide groups. The Si O Si groups can also be detected in based upon proposed scheme in this investigation. Another impor-
the hybrid films at wavelength in the range of 1100–1200 cm−1 . tant aspect can be found comparing the SEM images of E4T12.5
To make a deeper sense on formation of covalent bond between and E16T12.5. Accordingly, larger silica domains in E16 before cur-
epoxy–amine network and silica domains, one can see Scheme 4. ing might have been the reason for morphology coarsening in case
E16T12.5 when compared with the other cured film. In the other
3.4.2. Haziness and morphology words, the final size of the silica domains depends on their initial
All cured samples formed a slightly hazy transparent film, but size achieved in the pre-hydrolytic stage, as evidenced by optical
different in haziness values. The haziness values of the hybrid films microscopic study.

Fig. 5. FTIR spectrum of E8T12.5 sample after curing reaction.


E. Bakhshandeh et al. / Progress in Organic Coatings 77 (2014) 1169–1183 1175

Fig. 6. Haziness of the cured films of samples; effects of APTES amounts (E16:16/1,
E8:8/1, S4:4/1) and pre-hydrolyzed TEOS (7.5, 12.5, 17.5).

3.4.3. Mechanical properties


To evaluate the adhesion strength of hybrid coatings, pull-off
adhesion test was carried out on the hybrid coatings as well as neat
epoxy coating applied on mild steel surface. The results of measure-
ments are given in Table 1. Accordingly, the hybrid epoxy–silica
coatings wholly show weaker adhesion strengths, approximately
from 22 to 58% lower in comparison with the neat epoxy coating. It
is known that the presence of APTES in epoxy matrix has beneficial
aspects regarding adhesion characteristics because of acceptable
Si O metal interface [27]. On the other hand, the presence of TEOS
Scheme 4. Chemical reaction between DGEBA–APTES and pre-hydrolyzed TEOS.
in the hybrid coating causes formation of a silica layer near the
surface of substrate, thereby metal-coating adhesion might be suf-
fered due to formation of microcracks within the cured films. As

Fig. 7. SEM image of epoxy–silica hybrids. (A) E4T12.5, (B) E8T12.5, (C) E16T12.5 and (D) pure epoxy.
1176 E. Bakhshandeh et al. / Progress in Organic Coatings 77 (2014) 1169–1183

Fig. 8. The size distribution of silica domains in epoxy–silica hybrid (E8T12.5).


Fig. 10. Micro-hardness of the hybrid coatings: effects of APTES amounts (E16:16/1,
E8:8/1, E4:4/1) and pre-hydrolyzed TEOS wt% (7.5, 12.5, 17.5).
can be seen in Table 2, manipulating the content of APTES does
not seriously affect the adhesion strength of hybrid coatings hav-
ing the same amount of TEOS (one can compare E4Tx with E8Tx that silica nanoparticles are evenly distributed on the surface of
and E16Tx, when x = 7.5, 12.5, and 17.5). Evidently, the adhesion the sample.
strength is enhanced between 30 and 40% with increasing concen-
tration of TEOS from 7.5 to 12.5 wt%, but followed a descending 3.4.4. EIS studies
trend at concentration of 17.5 wt%, regardless of APTES content. The Electrochemical impedance spectroscopy (EIS) is well-known
lowest value of pull-off data among hybrid coatings is devoted to for intensively prediction of corrosion resistance. Impedance indi-
sample E4T17.5. This was expected in view of population of micro- cates changes at the interface of coating and metal long before
cracks in this sample. Thus, very high concentration of Si O metal visual traditional exposure tests can indicate [45,46]. Employing
linkages could be a cause for adhesion strength depression in case EIS results, it is possible to deeply delve anticorrosive character-
of hybrid coatings with respect to neat epoxy coatings. istics of prepared coatings. We used Bode and Nyquist plots for
To place value on this speculation, we provided SEM micro- identification of coating performance and better description of the
graphs from the detached surface of E8T12.5 from the mild steel coating impedance behavior. The impedance spectra were ana-
surface exhibiting the highest pull-off value in this family of lyzed using impedance in low frequency (lZl0.01 Hz ) which is an
coatings (Fig. 9). This image clearly shows formation of silica layer indicator of coating condition under continuous immersion [47].
near the substrate. Besides, an electrochemical equivalent circuit model was suggested
The results of micro-hardness measurements are plotted in to correlate EIS variations with corrosive environments. Model
Fig. 10. The micro-hardness of the hybrid coatings appears to be (a) in Fig. 12 is simply illustrative of Rsol and Rpor respectively
dependent on the pre-hydrolyzed TEOS content. The maximum resistance caused by solution, and passive film, while Qcoat stands
attainable hardness in this series is 12.3 HV corresponding to the for the constant-phase element for the passive film. This circuit
case that 12.5 wt% of TEOS is used, i.e. E4T12.5 sample. By further was designed to fit all the experimental results from the hybrid
increasing in TEOS content from up to 17.5 wt%, the micro-hardness coatings after half an hour of immersion in the sodium chloride
values follow a descending trend. This can be ascribed to lack of uni- solution. Also, circuit (b) was used to fit the experimental data
formity and improper distribution of silicate domains throughout after 15 and 45 days of immersion in the salt solution. In this
the film. On the other hand, increasing the APTES content leads circuit, Rpor is the pore resistance of coating, Qcoat the constant-
to compatibility enhancement between the organic and inorganic phase element for the passive film, RlSi the resistance caused by
phases showing micro-hardness amelioration. Fig. 11 presents SEM silica layer possibly formed in the neighborhood of metal, QlSi the
images of the surface of E8T12.5. SEM micrograph clearly shows constant-phase element for silica layer capacitance, Rpol the

Fig. 9. SEM and graphic images of epoxy–silica hybrid coating, the detached surface of E8T12.5 from the mild steel surface.
E. Bakhshandeh et al. / Progress in Organic Coatings 77 (2014) 1169–1183 1177

Fig. 11. SEM and graphic images of epoxy–silica hybrid coating, the surface of E8T12.

charge-transfer resistance, and Qdl the constant-phase element for and water through the films. As more water diffuses into the film
double-layer capacitance at the coating-substrate interface. In both during immersion, the film resistance declines and finally water
circuits, Rsol is the electrolyte resistance. The typical impedance finds its path to the substrate and accelerate corrosion reactions.
spectra of the steel substrate coated with hybrid films after half an The proposed electrical equivalent circuits for the epoxy–silica
hour, 15 days and 45 days of immersion in 3.5 wt% NaCl solutions hybrid coatings corresponding to 15 and 45 days of immersion are
are compared in Figs. 13, 15 and 17. Fig. 13 shows the Bode plots shown in Figs. 16 and 18, respectively. As can be seen, the resis-
of hybrid coatings after half an hour of immersion in 3.5 wt% NaCl tance of coatings decreases by more than one order of magnitude
solution. with increasing the immersion time. However, each hybrid coat-
Employing the EIS data relating to epoxy–silica hybrid coatings ing exhibited similar trend by the time to failure (Figs. 15 and 17)
immersed for half an hour, the Rs (Qsol Rpor ) model was applied on as determined from low frequency depression in impedance with
equivalent circuit model to make a fit on single passive film pre- respect to immersion time. The higher amounts of pre-hydrolyzed
sented on the metal surface (see Fig. 14). The corrosion resistance TEOS and APTES obviously accelerated downward trend in the rate
of the hybrids on steel was evaluated by the coating resistance of coating resistance for the hybrid coatings immersed in NaCl solu-
(Rpor ) and impedance at low frequency region (lZl0.01 Hz ) after half tion (Fig. 19). This behavior could be attributed to higher contents
an hour of immersion in a 3.5 wt% NaCl solution. Accordingly to of silica content in the films. For drawing comparison between
Table 3, there is a good agreement between the experimental and experimental and fitted data, we provided Tables 4 and 5. In this
fitted values. regard, we simulated a parallel arrangement of a resistor and a
In a similar manner, Figs. 15 and 17 are obtained after 15 and capacitor for the silica layer near the substrate. The resistance of
45 days of immersion in a 3.5 wt% NaCl solution and revealed a growing silica layer near the substrate, Rlsi , increases with increas-
fall in the impedance values to 107 –108  and 105 –106 , respec- ing the TEOS content, whereas the capacitance Clsi , decreases. This
tively. In these plots, two time constants are obviously seen in the trend was expected taking into account growing silica domains
EIS spectra of the Bode plots over an immersion time increase, in the assigned substrates. A comparison between the impedance
demonstrating three corrosion regions. This may be assigned to spectra of different samples after 15 and 45 days of immersion
the formation of defects caused by penetration of ionic species test revealed that the resistances and capacitances of the hybrid
coatings decreased and increased respectively after 45 days of
immersion (see Figs. 15 and 17). These alterations are indeed simi-
lar to the results obtained after 15 days of immersion; thereby, the
rate of water uptake and the hydrolytic stability of coatings during
immersion in aqueous solutions can be determined from the evo-
lution of the capacitance of the hybrid films [48]. The water uptake
values summarized in Table 6 are obtained from Brasher–Kingsbury
equation (2):
log(Ct /C0 )
Water uptake = (2)
log 80
where Ct and C0 are the electrical capacitance of the coating during
and before the immersion, respectively [46]. The results indicate
that water uptake of hybrid coatings decreases as the silica con-
centration increases. These results confirm that by incorporation
of pre-hydrolyzed TEOS in the hybrid coatings the corrosion resis-
tances enhances, meanwhile, water uptake decreases. It can be
concluded that silica domains might have been formed as a barrier
layer inside the film, which prevent the penetration of electrolyte
and decrease the corrosion rate. This is confident with SEM obser-
vations in Fig. 9. The condensed SiO2 layer on the substrate was also
deemed to be able to block the electrolyte path leading to corro-
Fig. 12. (a) Randles equivalent circuit of intact coating (b) proposed electrical equiv-
sion inhibition. In this context, increasing the pre-hydrolyzed TEOS
alent circuit for corroded hybrid coated metals. content has positive impacts on anticorrosive characteristics.
1178 E. Bakhshandeh et al. / Progress in Organic Coatings 77 (2014) 1169–1183

Fig. 13. Bode plots for different samples after half an hour of immersion (A) E16TX, (B) E8TX, (C) E4TX; (♦) X = 7.5 wt%, () X = 12.5 wt% and () X = 17.5 wt%.

Fig. 14. Fitted Bode and Nyquist impedance spectra of E8T12.5 epoxy–silica hybrid coatings after half an hour of immersion.

3.4.5. TGA analysis of incomplete polycondensation reaction that is residual silanols.


Thermogravimetric analysis was performed to investigate On the other hand, the wide interval of the weight loss within
the thermal-decomposition behavior of the epoxy–silica hybrid 200–600 ◦ C might be attributed to decomposition of polymeric part
coatings. The thermograms (Fig. 20) show a two-step degrada- of the coatings. Table 7 provides thermal properties of pure epoxy
tion mechanism. It goes without saying that the weight loss below resin and hybrid samples in terms of some important tempera-
200 ◦ C is attributable to evaporation of volatiles ingredients used tures expecting from TGA test. It is obvious that the initial thermal
in the hybrid system, which might have been remained because decomposition temperature (Td ) of the epoxy–silica hybrids is

Table 3
Selected circuit fitting results for elements as shown in Fig. 1(a), after half an hour of immersion.

Samples Qcoat (F/cm2 ) Rpor lZl at 0.01 Hz Phase angles at 104 Hz n


−10
E16T7.5 2.2 × 10 1.77 × 10 9
1.81 × 109
89.01 0.95
E16T12.5 2.0 × 10−10 3.81 × 109 3.91 × 109 89.52 0.94
E16T17.5 1.8 × 10−10 5.42 × 109 5.45 × 109 89.76 0.96
E8T7.5 2.3 × 10−10 1.58 × 109 1.59 × 109 89.67 0.94
E8T12.5 2.0 × 10−10 4.106 × 109 4.2 × 109 89.14 0.93
E8T17.5 1.0 × 10−10 4.55 × 109 4.65 × 109 89.24 0.96
E4T7.5 7.0 × 10−11 3.017 × 109 3.03 × 109 88.56 0.94
E4T12.5 6.2 × 10−11 4.448 × 109 4.45 × 109 88.83 0.95
E4T17.5 2.3 × 10−11 2.073 × 109 2.73 × 109 88.97 0.95
E. Bakhshandeh et al. / Progress in Organic Coatings 77 (2014) 1169–1183 1179

Fig. 15. Bode plots for different samples after 15 days of immersion, (A) E16TX, (B) E8TX, (C) E4TX; (♦) X = 7.5 wt%, () X = 12.5 wt% and () X = 17.5 wt%.

Fig. 16. Fitted Bode and Nyquist impedance spectra of E8T12.5 epoxy–silica hybrid coatings after 15 days of immersion.

higher than that of the pure epoxy. The pure epoxy sample starts to attributable to the degree of compatibility between phases in the
decompose at 320 ◦ C, while the hybrids were possibly decomposed hybrid coatings, while the ash content accounts for concentra-
between 325 ◦ C and 342 ◦ C. These observations can be attributed tion of inorganic silica domains remained after TGA test. Since
to the strong interaction between the polymer chains and silica increase of TEOS in the system causes more incompatibility, Td
domains, which prevents thermal decomposition of the hybrids could be suffered at high TEOS contents. Ash content of epoxy and
[41,49]. It should be mentioned that fluctuation in Td would be hybrid samples are measured at 1000 ◦ C to evaluate contribution of

Table 4
Selected circuit fitting results for elements as shown in Fig. 15(b), for studied samples after 15 days of immersion.

Samples Rpor Qcoat RlSi Qlsi Rp Qdl lZl at 0.01 Hz


−10 −6 −7
E16T7.5 3.5 × 107
3.2 × 10 2.0 × 107
3.0 × 10 8.0 × 106
3.0 × 10 4.91 × 107
E16T12.5 4.2 × 107 2.5 × 10−10 2.2 × 107 1.0 × 10−8 1.0 × 107 1.0 × 10−6 6.52 × 107
E16T17.5 5.0 × 107 2.1 × 10−10 8.6 × 107 5.0 × 10−8 5.2 × 107 2.0 × 10−7 2.21 × 108
E8T7.5 2.4 × 107 3.3 × 10−10 2.2 × 107 9.5 × 10−9 8.8 × 106 1.0 × 10−8 5.41 × 107
E8T12.5 2.5 × 107 2.7 × 10−10 4.0 × 107 3.0 × 10−8 2.0 × 107 1.0 × 10−8 9.74 × 107
E8T17.5 7.3 × 107 2.1 × 10−10 7.0 × 107 5.0 × 10−8 1.0 × 107 1.0 × 10−8 1.64 × 108
E4T7.5 2.0 × 107 1.0 × 10−10 1.2 × 107 1.1 × 10−8 2.0 × 107 2.9 × 10−8 4.91 × 107
E4T12.5 2.5 × 107 2.1 × 10−10 5.0 × 107 1.0 × 10−7 4.0 × 107 1.0 × 10−8 7.53 × 107
E4T17.5 7.1 × 107 1.0 × 10−10 7.5 × 107 2.0 × 10−7 8.2 × 107 1.0 × 10−7 1.73 × 108
1180 E. Bakhshandeh et al. / Progress in Organic Coatings 77 (2014) 1169–1183

Fig. 17. Bode plots for different samples after 45 days of immersion, (A) E16TX, (B) E8TX, (C) E4TX; (♦) X = 7.5 wt%, () X = 12.5 wt% and () X = 17.5.

Fig. 18. Fitted Bode and Nyquist impedance spectra of E8T12.5 epoxy–silica hybrid coatings after 45 days of immersion.

Table 5
Selected circuit fitting results for elements as shown in Fig. 15(b), for studied samples after 45 days of immersion.

Samples Rpor Qcoat RlSi Qlsi Rp Qdl lZl at 0.01 Hz

E16T7.5 1.02 × 105 2.1 × 10−9 1.0 × 104 2.0 × 10−3 1.0 × 104 2.0 × 10−3 1.19 × 105
E16T12.5 1.30 × 105 1.6 × 10−9 8.1 × 104 1.1 × 10−3 3.0 × 104 1.0 × 10−4 2.10 × 105
E16T17.5 5.31 × 105 1.0 × 10−9 8.6 × 104 1.1 × 10−4 3.2 × 104 1.0 × 10−5 7.55 × 105
E8T7.5 1.12 × 105 7.4 × 10−9 4.2 × 104 1.2 × 10−6 2.0 × 104 1.0 × 10−4 1.09 × 105
E8T12.5 6.12 × 106 3.2 × 10−9 1.0 × 106 2.0 × 10−6 2.0 × 105 1.0 × 10−5 6.92 × 106
E8T17.5 7.21 × 106 1.5 × 10−9 1.3 × 106 4.1 × 10−7 1.0 × 105 1.0 × 10−6 8.81 × 106
E4T7.5 1.12 × 106 8.5 × 10−10 5.2 × 105 2.5 × 10−7 3.0 × 105 2.0 × 10−5 1.02 × 106
E4T12.5 2.42 × 106 6.5 × 10−10 1.0 × 106 1.5 × 10−7 5.0 × 104 2.0 × 10−5 3.42 × 106
E4T17.5 6.10 × 106 2.1 × 10−10 2.1 × 106 1.0 × 10−7 1.2 × 105 2.0 × 10−5 1.05 × 107

Table 6
Water uptake values as a function of TEOS wt% for hybrid coatings.

V.F.H2 O Sample

E16T7.5 E16T12.5 E16T17.5 E8T7.5 E8T12.5 E8T17.5 E4T7.5 E4T12.5 E4T17.5

After 15 days 0.223 0.051 0.033 0.211 0.147 0.045 0.116 0.084 0.017
After 45 days 0.515 0.475 0.391 0.793 0.633 0.618 0.570 0.537 0.509
E. Bakhshandeh et al. / Progress in Organic Coatings 77 (2014) 1169–1183 1181

Fig. 19. Impedance in low frequency (lZl0.01 Hz ) changes during immersion time (day) as determined by circuit fitting over 0.5 h and 15, 45 days of immersion. (A) ExT7.5, (B)
ExT12.5, (C) ExT17.5; (♦) X = 16, () X = 8 and () X = 4.

inorganic silica domains to the coatings. All hybrid specimens have the content of APTS coupling agent causes a shift in Td and coating
ash contents in the range of 2.33–6.83 wt% in the order that TEOS possessing higher compatibility are formed. For example, increas-
content is increased. Regarding the compatibility between organic ing Td from 332 to 342 moving from E16T7.5 toward E8T7.5 can be
and inorganic parts of the sample, it can be inferred that increasing considered.

Fig. 20. TGA curves of the hybrid coatings. (A) E16TX, (B) E8TX, (C) E4TX; X = 7.5, 12.5 and 17.5.
1182 E. Bakhshandeh et al. / Progress in Organic Coatings 77 (2014) 1169–1183

Table 7 [3] G.P. Bierwagen, R. Twite, G. Chen, D.E. Tallman, Atomic force microscopy, scan-
The initial degradation temperature (Td ) and weight loss percent of epoxy–silica ning electron microscopy and electrochemical characterization of Al alloys,
hybrids. conversion coatings, and primers used for aircraft, Prog. Org. Coat. 32 (1997)
25–30.
Samples Td (◦ C) Total decrease of [4] R.L. Twite, G.P. Bierwagen, Review of alternatives to chromate for corrosion
wt% at 1000 ◦ C protection of aluminum aerospace alloys, Prog. Org. Coat. 33 (1998) 91–100.
[5] X.F. Yang, D.E. Tallman, V.J. Gelling, G.P. Bierwagen, L.S. Kasten, J. Berg, Use
E16T7.5 332 97.67
of a sol–gel conversion coating for aluminum corrosion protection, Surf. Coat.
E16T12.5 325 96.29
Technol. 140 (2001) 44–50.
E16T17.5 328 94.98 [6] D.G. Shchukin, M. Zheludkevich, K. Yasakau, S. Lamaka, M.G.S. Ferreira, H. Möh-
E8T7.5 342 97.52 wald, LbL nanocontainers for self-healing corrosion protection, Adv. Mater. 18
E8T12.5 341 96.12 (2006) 1672–1678.
E8T17.5 333 94.83 [7] M. Sarwar, S. Zulfiqar, Z. Ahmad, Preparation and properties of polyamide–
E4T7.5 329 96.88 titania nanocomposites, J Sol–Gel Sci Technol. 44 (2007) 41–46.
E4T12.5 341 95.99 [8] N.N. Voevodin, J.W. Kurdziel, R. Mantz, Corrosion protection for aerospace alu-
E4T17.5 327.1 93.17 minum alloys by modified self-assembled nanophase particle (MSNAP) sol–gel,
Pure epoxy 320 100 Surface Coat. Technol. 201 (2006) 1080–1084.
[9] A. Conde, A. Duran, M. de Damborenea, Polymeric sol–gel coatings as protective
layers of aluminium alloys, Prog. Org. Coat. 46 (2003) 288–296.
[10] M.L. Zheludkevich, R. Serra, M.F. Montemor, I.M. Salvado, M.G.S. Ferreira,
4. Conclusions Corrosion protective properties of nanostructured sol–gel hybrid coatings to
AA2024-T3, Surf. Coat. Technol. 200 (2006) 3084–3094.
Hybrid coatings based on silane functionalized diglycidyl [11] M. Fir, B. Orel, A.S. Vuk, A. Vilcnik, R. Jese, V. Francetic, Corrosion studies
and interfacial bonding of urea/poly(dimethylsiloxane) sol/gel hydrophobic
ether of bisphenol A (DGEBA) epoxy resin and pre-hydrolyzed
coatings on AA 2024 aluminium alloy, Langmuir 23 (2007) 5505–5514.
tetraethoxysilane (TEOS) were prepared and comprehensively [12] C. Sanchez, F. Ribot, Design of hybrid organic–inorganic materials synthesized
characterized. In the studied hybrid systems, we changed the con- via sol–gel, New J. Chem. 18 (1994) 1007–1047.
[13] L. Mascia, L. Prezzi, Developments in organic–inorganic polymeric hybrids:
tent of 3-aminopropyl triethoxysilane (APTES) coupling agent to
ceramers, Polym. Sci. 3 (1995) 61–66.
make organic and inorganic parts of hybrid system compatible. [14] L. Mascia, T. Tang, Curing and morphology of epoxy resin silica hybrids, J. Mater.
It was found that increasing the APTES content causes a viscos- Chem. 8 (1998) 2417–2421.
ity upturn at low shear rates. Confidently, FTIR spectra illustrated [15] L. Prezzi, L. Mascia, Network density control in epoxy–silica hybrids by selec-
tive silane functionalization of precursors, Adv. Polym. Technol. 24 (2) (2005)
that the shear thinning behavior of the silane-functionalized epoxy 91–102.
resins can be attributed to the formation of silica domains. The [16] L. Mascia, L. Prezzi, B. Haworth, Substantiating the role of phase bicontinuity
morphological studies placed value on this finding indicating the and interfacial bonding in epoxy–silica nanocomposites, J. Mater. Sci. 41 (2006)
1145–1155.
presence of a slightly hazy transparent hybrid film with nanoscale [17] L. Mascia, L. Prezzi, G. Wilcox, M. Lavorgna, Molybdate doping of networks in
dimensions. The haziness measurements showed that the com- epoxy–silica hybrids: domain structuring and corrosion inhibition, Prog. Org.
patibility between the organic and inorganic phases depends on Coat. 56 (2006) 13–22.
[18] M.Z. Rong, M.Q. Zhang, Y.X. Zheng, H.M. Zeng, K. Friedrich, Improvement of
the APTES content, as sample E4 (epoxide/amine ratio of 4:1) tensile properties of nano-SiO2 /PP composites in relation to percolation mech-
revealed smaller silica domains among all prepared samples. On anism, Polymer 42 (2001) 3301–3304.
the other hand, the haziness of hybrid films has obviously risen [19] R.P. Singh, M. Zhang, D. Chan, Toughening of a brittle thermosetting polymer:
effects of reinforcement particle size and volume fraction, J. Mater. Sci. 37 (4)
with increasing the amount of pre-hydrolyzed TEOS in the formu-
(2002) 781–788.
lations. It was also observed that adhesion strength of samples [20] Q.X. Zhang, Z.Z. Yu, X.L. Xie, Y.W. Mai, Crystallization and impact energy
enhances in the range of 30–40% by weight over increase of of. polypropylene/CaCO3 nanocomposites with nonionic modifie, Polymer 45
(2004) 5985–5994.
TEOS from 7.5 to 12.5 wt%, while there was a fall assigned to
[21] Z. Ahmad, H.U. Rehman, S. Ali, M.I. Sarwar, Thermal degradation of
17.5 wt% of TEOS. The maximum micro-hardness was 12.3 HV poly(vinyichioride)-stabilization effect of dichlorotin dioxin, Polym. Mater. 46
with devoted to sample containing 12.5 wt% TEOS, namely E4T12.5, (2000) 547–559.
and further increase in TEOS content up to 17.5 wt% caused a [22] S.H.R. Lu, W. Chun, J. Yu, X. Yang, Characterization of the mesoporous
SiO2 –TiO2 /epoxy resin hybrid materials, Polym. Sci. 109 (2008) 2095–2102.
descending trend being observed. The EIS experimental data were [23] S.H.R. Lu, H.L. Zhang, C.X. Zhao, X.Y. Wang, Preparation and characterization of
analyzed considering two equivalent electrical circuits. The fit- epoxy–silica hybrid materials by the sol–gel process, J. Mater. Sci. 40 (2005)
ted values corresponding to equivalent circuits were used to 1079–1085.
[24] C.L. Chiang, C.C.M. Ma, Synthesis, characterization and thermal properties of
evaluate the evolution of corrosion protection properties for the novel epoxy containing silicon and phosphorus nanocomposites by sol–gel
different hybrid coatings under study. Analysis of EIS data sug- method, J. Eur. Polym. 38 (2002) 2219–2224.
gests that the corrosion protection of epoxy–silica coatings is [25] G. Schottner, Hybrid sol–gel-derived polymers: applications of multifunctional
materials, Chem. Mater. 13 (2001) 3422–3435.
attributable to formation of an intermediate silica layer at the [26] J. Sunirmal, L. Mi Ae, C.B. In, H.K. Chang, S. Sang, Non-hydrolytic sol–gel synthe-
coating-substrate interface, which prevents the penetration of elec- sis of epoxysilane-based inorganic–organic hybrid resins, Mater. Chem. Phys.
trolyte and decreases the corrosion rate. We also served water 112 (2008) 1008–1014.
[27] C.J. Brinker, G.W. Scherer, Sol–gel Science, Academic Press, London, 1990.
uptake quantities to prove anticorrosive nature of fabricated hybrid
[28] Ch. R. Wold, M.D. Soucek, Viscoelastic and thermal properties of lin-
coatings. Thermogravimetric analysis evidenced that thermal sta- seed oil-based ceramer coatings, Macromol. Chem. Phys. 201 (2000)
bility of epoxy–silica hybrid coatings is highly dependent on the 382–392.
[29] H. Nia, W.J. Simonsick Jr., B, A.D. Skaja, J.P. Williams, M.D. Soucek,
concentration of APTES coupling agent as well as inorganic phase
Polyurea/polysiloxane ceramer coatings, Prog. Org. Coat. 38 (2000)
content. 97–110.
[30] N.N. Voevodin, N.T. Grebasch, W.S. Soto, F.E. Arnold, M.S. Donley, Potentio-
dynamic evaluation of sol–gel coatings with inorganic inhibitors, Surf. Coat.
Acknowledgment
Technol. 140 (2001) 24–28.
[31] S. Frings, H.A. Meinema, C.F. Van Nostrum, R. Van, der. Linde, Organic–inorganic
The authors would like to acknowledge with gratitude Pars Pam- hybrid coatings for coil coating application based on polyesters and
chal Chemical Co. who kindly supported this work financially. tetraethoxysilane, Prog. Org. Coat. 33 (1998) 126–130.
[32] J.W. Cho, S.H. Lee, Influence of silica on shape memory effect and mechan-
ical properties of polyurethane–silica hybrids, Eur. Polym. J. 40 (2004)
References 1343–1348.
[33] R.A. Sailer, M.D. Soucek, Linseed and sunflower alkyd ceramers, Prog. Org. Coat.
[1] Z.W. Wicks Jr., F.N. Jones, S.P. Pappas, D.A. Wicks, Organic Coatings: Science and 33 (1998) 36–43.
Technology, 3rd ed., Wiley-Interscience, New York, 2007. [34] H. Nia, A.H. Johnson, M.D. Soucek, J.T. Grant, A.J. Vreugdenhil,
[2] D.A. Jones, Principles and Prevention of Corrosion, Prentice Hall, Upper Saddle Polyurethane/polysiloxane ceramer coatings: evaluation of corrosion
River, NJ, 1996, pp. 104–105. protection, Macromol. Mater. Eng. 287 (2002) 470–479.
E. Bakhshandeh et al. / Progress in Organic Coatings 77 (2014) 1169–1183 1183

[35] T.T.X. Hang, T.A. Truc, T.H. Nam, V.K. Oanh, J.B. Jorcin, N. Pebere, Corrosion [43] J.D. Basil, C.C. Lin, in: C.J. Brinker, D.E. Clark, D.R. Ulrich (Eds.), Better Ceram-
protection of carbon steel by an epoxy resin containing organically modified ics Through Chemistry I, Materials Research Society, Pittsburgh, PA, 1988, pp.
clay, Surf. Coat. Technol. 201 (2007) 7408–7415. 49–57.
[36] J.L. Massingill, P.S. Sheih, R.C. Whiteside, D.E. Benton, D.K. Morissearnold, Fun- [44] J.D. Basil, C.C. Lin, in: C.J. Brinker, D.E. Clark, D.R. Ulrich (Eds.), Better Ceram-
damental studies of epoxy resins for can and coil. II. Flexibility and adhesion, J. ics Through Chemistry II, Materials Research Society, Pittsburgh, PA, 1986, pp.
Coat. Technol. 62 (1990) 31–39. 585–590.
[37] M.T. Rodríguez, J.J. Gracenea, S.J. García, J.J. Saura, J. Suay, A critical appraisal [45] J.R. Scully, D.C. Silverman, M.W. Kendig, Coatings on metals, in: J.R. Scully,
of the potential of self-healing polymeric coatings, Prog. Org. Coat. 50 (2004) D.C. Silverman, M.W. Kendig (Eds.), Electrochemical Impedance: Analysis and
123–131. Interpretation, ASTM, Philadelphia, 1993.
[38] S.J. Garcia, J. Suay, Epoxy powder clearcoats used for anticorrosive purposes [46] D.D. Macdonald, M.C.H. McKubre, Corrosion of materials, in: E. Barsoukov, J.R.
cured with ytterbium. III. Trifluoromethanesulfonate, Prog. Org. Coat. 57 (2006) Macdonald (Eds.), Impedance Spectroscopy Theory, Experiment, and Applica-
319–331. tions, 2nd ed, John Wiley & Sons Inc., Hoboken, NJ, 2005.
[39] L. Matejka, K. Dusek, J. Plestil, J. Kriz, F. Lednicky, Formation and structure of [47] S.K. Poznyak, M.L. Zheludkevich, D. Raps, F. Gammel, K.A. Yasakau, M.G.S.
the epoxy–silica hybrids, Polymer 40 (1999) 171–181. Ferreira, Preparation and corrosion protective properties of nanostructured
[40] L. Matejka, J. Plestil, K. Dusek, Structure evolution in epoxy–silica hybrids: titania-containing hybrid sol–gel coatings on AA2024, Prog. Org. Coat. 62 (2)
sol–gel process, J. Non-Cryst. Sol. 226 (1998) 114–121. (2008) 226–235.
[41] Z.H. Huang, K.Y. Qiu, Y.J. Wei, Properties of hybrid materials incorporating tetra- [48] A.S.L. Castela, A.M. Simões, M.G.S. Ferreira, E.I.S. evaluation of attached and free
butyl titanate and tetraethoxysilane with ethylene–propylene nonconjugated polymer films, Prog. Org. Coat. 38 (2000) 1–7.
diene terpolymer (EPDM–ENB) via sol–gel process, Polym. Sci. Part A: Polym. [49] Z. Ahmad, M.I. Sarwar, J.E. Mark, Dynamic–mechanical thermal analysis of
Chem. 35 (12) (1997) 2403–2409. aramid–silica hybrid composites prepared in a sol–gel process, J. Appl. Polym.
[42] L.W. Kelts, N.J. Armstrong, in: C.J. Brinker, D.E. Clark, D.R. Ulrich (Eds.), Better Sci. 63 (1997) 1345–1352.
Ceramics Through Chemistry III, Materials Research Society, Pittsburgh, PA,
1988, pp. 519–522.

Você também pode gostar