Вы находитесь на странице: 1из 10

Corrosion Science 52 (2010) 19481957

Contents lists available at ScienceDirect

Corrosion Science
journal homepage: www.elsevier.com/locate/corsci

The adhesion properties and corrosion performance of differently pretreated


epoxy coatings on an aluminium alloy
M. Niknahad a, S. Moradian a, S.M. Mirabedini b,*
a
b

Polymer Engineering Department, Amirkabir University of Technology, P.O. Box 15875-413, Tehran, Iran
Colour, Resin & Surface Coatings Department, Iran Polymer and Petrochemical Institute, P.O. Box 14965-115, Tehran, Iran

a r t i c l e

i n f o

Article history:
Received 11 August 2009
Accepted 6 February 2010
Available online 14 February 2010
Keywords:
A. Aluminium
B. EIS
B. IR spectroscopy
B. SEM
C. Polymer coatings

a b s t r a c t
The inuence of various blends of hexauorozirconic-acid (Zr), polyacrylic-acid (PAA) and polyacrylamide (PAM) pretreatment on the performance of an epoxy coated aluminium substrate was investigated
and compared to that of a so-called chromate/phosphate conversion coating (CPCC).
Adhesive-strength of epoxy coated substrates was evaluated using pull-off and tape tests. Salt spray,
humidity chambers and EIS were employed to characterize corrosion performance of coated substrates
with different initial surface pretreatments. Among the Zr-based formulations, PAA/Zr and PAA/PAM/Zr
showed the best adhesion strength, while the later revealed a good corrosion performance as well. However, CPCC pretreated sample was still superior in these aspects.
2010 Elsevier Ltd. All rights reserved.

1. Introduction
Aluminium and its alloys are widely used because of their some
what unique properties, such as lightweight, relative low toxicity
and a fair corrosion resistance [1,2] and the surface pretreatments
play an important role in the protection of aluminium and aluminium alloys [3]. Although the aluminium surface naturally covers
with an air-formed oxide lm, about 23 nm, it is an insufcient
barrier for moderately long-term corrosion prevention of the
underlying substrate. The mentioned oxide-lm covered aluminium is attacked in certain environments, even after being further
coated by an organic protective coating [4].
Consequently, chemical conversion coatings are applied to aluminium to improve corrosion resistance and/or to establish a base
for subsequent application of organic coatings. Conversion coatings
on aluminium and its alloys are generally performed in solutions
containing chromate and uoride ions [1]. The coating usually
develops in the presence of uoride ions, which involves a cathodic
reduction of Cr(VI) to Cr(III) together with the evolution of hydrogen. Chromium-based treatments have been used widely with
good anti-corrosive performance, but the toxicity and carcinogenic
nature of hexavalent chromium are well documented [5]. Thus,
several alternative treatments, including zirconium- or titaniumcontaining processes, or thin layers of polymeric materials, have
been introduced [1,613].

* Corresponding author. Tel.: +98 21 4458 0040; fax: +98 21 4458 0023.
E-mail address: m.mirabedini@ippi.ac.ir (S.M. Mirabedini).
0010-938X/$ - see front matter 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.corsci.2010.02.014

Chibowski [14,15] has studied the inuence of ionic strength,


pH and molecular weight of PAA on adsorption onto an aluminium
oxide surface. Adsorption of PAA on metal oxides having surplus
positive charges increases with increasing degrees of dissociated
carboxylic groups in a PAA chain. The ratio of COOH/COO decreases when pH values of the solution changes from 3 to 9. The
dissociated carboxylic groups COO in the polymer chain repel
each other and hence are responsible for stretching and conformation of the polymer chain.
van den Brand and co-workers [16] have studied the effect of
differently functionalised interfacial polymer layers on adhesion
strength of an epoxy coated aluminium substrate. A good adhesion
strength and durability has been found for polyethylene maleic
anhydride based systems, as a result of formation of cured interfacial region. PAA based system showed the presence of carboxylate
ion throughout the interfacial region [15,16]. This indicates that a
curing reaction between the epoxy coating and PAA had occurred
but had not been fully completed. Moreover, a hydrophilic, relatively cured region can absorb a considerable amount of water.
Such a system would not reveal a good adhesion strength and
durability in the presence of water [16,17].
The role of hexauorozirconate in the formation of conversion
coating indicates that the uoride ion is involved in the formation
of an AlZrOF based layer that increases the hydrophilicity of the
surface with increasing surface activity. These, in turn enhance the
interaction between polymers and the aluminium alloy surface
resulting in the formation of a polymeric conversion coating [18].
In our previous study, the surface free energy of differently pretreated 1050 Al alloy was investigated [19]. The results showed

1949

M. Niknahad et al. / Corrosion Science 52 (2010) 19481957

that hot water treated, PAA and CPCC pretreated samples provide
the highest surface free energies. Comparison of these results with
adhesion strength measurements showed that good wettability is
an essential factor in achieving relatively good dry adhesive joints,
however; this cannot be considered to be a sufcient reason for
high bond strengths achieved in practice.
In a further study [20], Zr and PAA were selected to provide a
chrome-equivalent corrosion protection as well as to facilitate
the adhesion of a polyester/epoxy powder coating. Experimental
results revealed that adhesion of an aluminium substrate in the
presence of a polyester/epoxy coating had indeed improved by a
PAA/Zr based pretreatment and a chromate-based equivalent corrosion performance was achieved [20].
The main purpose of the present study was to evaluate the
interaction between PAA, PAM and Zr on an 1050 aluminium alloy
surface. Additionally, the effect of different composition of PAA/
PAM/Zr pretreatments on the adhesion strength and corrosion protection performance of a clear epoxy coating on the aluminium alloy were also studied.

2. Experimental
2.1. Materials
Aluminium alloy (1050) was supplied by Arak Al Company as
0.5 mm thick sheets in H18 temper. PAA (average molecular
weight: 104,000) as a 63% solution in distilled water, and carboxyl
modied PAM (average molecular weight: 200,000) were both
supplied by ACROC Organic (Fisher Chemicals). Hexauorozirconic
acid was obtained from Aldrich Chemicals Company as a 45% solution in distilled water.
Aluminium oxide powder with an average particle diameter of
0.1 lm was purchased from Merck Company. Epikote 1001, epoxy
clear coat, and its hardener, Ardur115 polyamide were supplied by
Shell Company. All chemicals used in this study, were of analytical
grade having high percentages of purity.
2.2. Sample preparation
Preliminary surface cleaning was carried out in acetone using a
soft brush. For chemical etching, the specimens were immersed in
a 5 wt.% solution of NaOH for 3 min at 50 C followed by washing
with distilled water. For desmutting of etched aluminium surfaces,
the specimens were then placed for 1 min in a 50% v/v solution of
nitric acid in water were then subsequently washed with distilled
water, and dried for 1 h at 50 C. The samples were then treated by
immersion in a solution of different combined weight percentages
of PAA, PAM and Zr at 20 1 C for 3 min.
Taguchi statistical analysis method was employed to optimize
the number of pretreatment solutions shown in Table 1. The specimens were then allowed to dry under ambient temperature. Postheating treatment was carried out in an oven at 145 C for 20 min.
For comparison purposes, a chromate/phosphate conversion coating (CPCC) solution was also used; the treatment involved the Alodine procedure [1], with a specic immersion time of 5 min.

Table 1
Pretreated samples with different combinations of PAA, PAM and Zr.
Treatment no.

PAA

PAM

Zr

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16

0.0
0.0
0.0
0.0
0.5
0.5
0.5
0.5
1.0
1.0
1.0
1.0
2.0
2.0
2.0
2.0

0.0
0.5
1.0
2.0
0.0
0.5
1.0
2.0
0.0
0.5
1.0
2.0
0.0
0.5
1.0
2.0

0.0
0.05
0.1
0.2
0.5
0.0
0.2
0.1
0.1
0.2
0.0
0.05
0.2
0.1
0.05
0.0

The number of pretreatment solutions were used are emphasized (bold) in Table 1.

Table 2
Mixture compounds for FTIR analysis.
No.

Combination

1
2
3
4
5

PAA + Al2O3
PAM + Al2O3
PAM + Zr
PAA + Zr + Al2O3
PAA + PAM

30 C for 1 h, subsequently heated for 20 min at 145 C, rinsed with


distilled water, and nally dried at 30 C for 1 h.
FTIRATR spectra of treated specimens and ltered powders
were recorded on a FTIRATR spectrometer (Equinox 55, Bruker,
Germany). The measurement conditions being: collecting 8 scans
in the 4004000 cm1 range with 4 cm1 resolution. The mixture
compounds were mixed with KBr powder in a ratio 1:100 for FTIR
analysis.
2.4. Application of coating layer
The clear coating was sprayed on the prepared samples at ambient temperature and 50% relative humidity. Panels were allowed to
cure for at least 10 days prior to examination. The thicknesses of
the cured coating lms were measured using an Elcometer 345
instrument to an accuracy of 0.1 lm, with measurements carried
out on a set of ve replicate samples. The coating thicknesses varied from 37.5 to 41.3 lm.
2.5. Accelerated corrosion tests
Two accelerated test methods, namely: salt spray and humidity
exposure were used for each treated aluminium specimen. Two sets
of samples were then exposed to a standard salt spray chamber for
1000 h at 35 1 C according to ASTM B 117 and exposed to a humidity cabinet for 1000 h at 38 2 C according to ASTM D 2247, respectively. Visual assessments of the macroscopic surfaces were carried
out at various time intervals during the exposure time.

2.3. FTIR spectroscopy


2.6. Electrochemical impedance spectroscopy (EIS)
In order to study possible interaction between aluminium oxide
powder and pretreatment solutions, 1 wt.% of various mixtures of
the materials and 1 wt.% of alumina powder in distilled water (Table 2), were prepared. The solutions were then agitated using a
magnetic stirrer with a rate of 300 rpm at 20 1 C. After 24 h, aluminium oxide powder was ltered from the solutions, dried at

A three-electrode arrangement, including an Ag/AgCl reference


electrode, a platinum counter electrode, and the coated aluminium
specimen (exposed area 40  40 mm2) immersed in a 3.5 wt.%
NaCl solution was used. EIS was carried out at the open circuit potential, using an Auto Lab G12 instrument. The data were obtained

1950

M. Niknahad et al. / Corrosion Science 52 (2010) 19481957

as a function of frequency, using a sine wave of 10 mV amplitude


peak to peak; a frequency range of 30 kHz to 10 mHz was selected.
2.7. Adhesion measurements
Pull-off adhesion testing of the coating was performed according to the procedure described in ASTM D 4541. The dollies of
20 mm diameter were degreased by acetone and then glued to
the surface of the coated panels with two components epoxybased or a cyano-acrylate adhesive. After adhesive curing, a testing
apparatus was attached to the loading xture and was strained at
5 mm/min in an Instron machine (Universal Testing Instrument)
until the coating material had detached from the substrate. For
each test, seven replicate samples were employed, and the average
value is quoted. The adhesion strength of the coated specimens in
the presence of water (wet adhesion) was measured after one
week exposure of the specimens in a humidity cabinet.
Tape adhesion measurements were undertaken immediately
after the humidity test according to ASTM D 3359. Six, approximately parallel, score lines were made with a separation of
1.0 mm; a further six score lines were scribed perpendicular to
the original score lines. For individual specimens, 25 grids were
generated. Adhesive tape was placed on the grids, using a soft eraser; the tape was then removed with a rm, steady pulling action.
2.8. Scanning electron microscopy (SEM)/energy dispersive
spectroscopy (EDS)
A Philips XL 40 SEM was used to examine the morphology of the
treated aluminium specimens. In this analysis, the samples, measuring 10  10 mm2, were placed in vacuum chamber of the
instrument. Samples were xed onto an aluminium stud using a
double-sided cello-tape. The samples were examined at various
magnications and the images were recorded photographically.
3. Results and discussion
3.1. FTIR analysis
The FTIR spectra of different mixture compounds as described
in Table 2, within the spectral range of 4004000 cm1 are shown
in Fig. 1. The most predominant and characteristic IR reectance
bands are listed in Table 3. The presence of a peak at 1568 cm1
in the spectrum treated with PAM and PAA mixture indicates the
formation of carboxylate groups (Fig. 1A).
The present illustrate that there is an ionic interaction between
COO and Al3+ with an absorption located near 1557 cm1 [21].
Spectra of mixtures that have been shown in Fig. 1A and B revealed
possible chemical interactions of PAA and PAM molecules with alumina powder. The presence of COO functional group in the case of
a relatively thin deposited layer of PAA results from a partial ionization of COOH groups in the presence of water molecules as
follows:

ACOOH H2 O $ ACOO H3 O
In the FTIR spectrum of mixture of PAM and H2ZrF6, Fig. 1C, due
to the dipole moment in the carbonyl group and the resonance effect of nitrogen, an absorption band in the region 16001700 cm1
was observed. Additionally the CO stretch band at 1270 cm1 has
disappeared from the spectrum of mixture compounds with
H2ZrF6. The disappearance of this band and the presence of an
adsorption band at 830 cm1 occur when acrylic acid reacts with
a ourozirconate compound. Applying polymers in combination
with a zirconium compound has been claimed to cause cross-linking at the aluminium surface [1].

Additionally, the shifts of the absorption of NH2 groups from


1417 and 1463 cm1 to 1408 and 1454 cm1 most likely indicate
the formation of amine salts and provide evidence for electrostatic
interaction between COO and NH
3 . Despite weak absorption of
PAM onto Al2O3, it could form strong bonds with PAA. Formation
of a relatively thin and insoluble layer on aluminium surface was
revealed previously by FTIR and EDXA techniques [20].
3.2. Accelerated corrosion tests
Visual assessments of the pretreated aluminium alloy samples
during humidity and salt spray tests are summarized in Tables 4
and 5. Appearance of the samples after 1000 h exposure in humidity conditions and salt spray tests are illustrated in Figs. 2 and 3,
respectively. The data does reveal consistently that PAA/PAM gave
the lowest scribe ratings. CPCC treatment clearly provided outstanding performance. Equivalent performance was achieved with
PAA/PAM/Zr treatment.
3.3. Electrochemical impedance spectroscopy (EIS)
It has been shown that EIS plays an important role to monitor
and predict degradation of organic coatings [2225]. Fig. 4 shows
Nyquist plots for the impedance of the epoxy coatings on aluminium substrate with different pretreatments over various immersion times in 3.5% NaCl electrolyte. The plots usually show one
or two semi-circles, with two-time constants at frequencies of
about 200100 and 30.2 Hz, respectively.
From Fig. 4, the rst measurement (1-day) of the degreased
substrates clearly shows capacitive behaviour with a parallel resistive component in excess of 106 X cm2, most likely due to the polymer lm [2628]. With increasing immersion time (10 days), the
radius of the high frequency semi-circle decreased. The coating
resistance values decrease during rst few days of immersion, indicating the entry of electrolyte into the epoxy coating [29]. This is
this step of electrolyte penetration through an the coating, due to
water uptake, when molecules of pure water diffuse into the
micropores of the polymer net according to Flicks law.
It is supposed that decreasing in the coating resistance value
may be due to the penetration of water and movement of ionic
species among the coating layer, increasing the coating conductivity. Initially, the electrolyte penetrates through the coating layer,
and sets up conducting paths at different depths within the coating
[21].
Further, a second semi-circle in the impedance spectrum is observed at reduced frequencies, suggesting [27,28] that electrochemical reactions at the interface between the coating and the
metal surface are making progress. At this stage, penetration is
completed and the electrolyte phase meets the metal/oxide interface and a corrosion cell is activated. With increased immersion
time (20 days), the barrier properties of the coating decreased further; however, the radius of the second semi-circle also decreased,
suggesting an increase in the corrosion rate, possibly through the
presence of further pores in the coating or an increase in the area
exposed at the base of the existing pores or aws [27].
It can be seen from Fig. 4 that epoxy coating on aluminium pretreated by CPCC has greater resistance than both epoxy coating on
degreased aluminium and other treatments, indicating the benecial role of chromium on the overall coating resistance. The spectra
demonstrate essentially capacitive behaviour. Not much change
with exposure time was observed in the capacitive region of the
spectra. Cr(VI) may exist included in the conversion coating and
is reduced at aws [4] to Cr(III) to re-passivate any damaged
surface.
Plots (C and F) in Fig. 4 show relative poor barrier properties of
PAM/Zr and PAA/PAM treated samples. This is perhaps due to the

M. Niknahad et al. / Corrosion Science 52 (2010) 19481957

1951

Fig. 1. FTIR spectra of different mixture compounds.

Table 3
FTIR assignments.
No.

Wave number (cm1)

Functionality

1
2
3
4
5
6
7
8
9
10
11

33002200
17251700
1557
1456
1417
1275
1106
35002500
16001700
14161457
1262

OH (acid)
COO (acid)
Acid salt. Ionized form
CH2
CHCO
OCOH (dimmer)
OCOH (dimmer)
NH2 (two peaks)
CNH2
NH2
CN

hydration of aluminium oxide at the water-rich interface [21], with


the water being replaced by aluminium hydroxide and polymeroxide bonds. In the presence of under-lm corrosion, the coating
resistance is low due to the existence of anodic and cathodic sites,
ionic migration within the lm (electrolyte), local lm changes and
electro-osmotic movement of water [30].
Davis and co-workers [31] suggested that hydrated aluminium
oxide decreases the stability of adhesive joints in the presence of
water. The coating layer may detach from metal substrates at the
interface due to time dependant diffusion of both water and oxygen through the coating to the metal/coating interface increasingly. Water may solvate un-reacted PAA molecules, resulting in
a loss of adhesion [20]. The high hydroxyl ion concentration dissolves the aluminium oxide and the aluminium hydroxide possibly

1952

M. Niknahad et al. / Corrosion Science 52 (2010) 19481957

Table 4
Visual observations of samples during 1000 h humidity test.

Table 5
Visual observation of samples during 1000 h salt spray test.

Treatment
no.

Samples conditions during humidity test

Treatment
no.

Samples conditions during salt spray test

Very small blisters were observed from the second day of


testing. These expanded as time elapsed
Very small blisters were observed after 15 days of exposure.
These expanded with increased exposure time
After 15 days of testing only slight changes in samples colours
were observed. After 30 days slight brown stains were observed.
With further increase exposure time, light brown stains
developed over 5% of the samples. Blisters were not observed for
samples 10 and 13
Some dark gray stains were observed over the sample surface
area after the second day of testing. With further exposure time
gray and dark areas covered the exposed sample surface. After
15 days very small blisters were observed over the samples
surfaces
No visible changes observed

Corrosion products and very small blisters were observed after a


week and blisters expanded fast
Corrosion products and pitting corrosion was observed after a
week. After 10 days white corrosion covered the entire sample
After 10 days slight changes in the appearance were evident.
Beyond 20 days isolated black stains were observed with a gray
back ground over 10% of the sample area. After 25 days,
corrosion products were occasionally observed and black stains
developed on the surface. These changes occurred at a slower
rate for sample No. 10
From the second day of exposure black and gray stains
developed over 25% of sample surface area. After a week white
corrosion products were observed with a gray background.
Pitting corrosion with white products was observed after
10 days. After 20 days white corrosion products covered the
sample, with small pits evident. The number of pits increased
with time
The sample appearance did not change with exposure time

4
10 and 13

16

CPCC

4
10 and 13

16

CPCC

attacks the polymer at the interface between the polymer and the
substrate.
With increased immersion time, the radius of the high frequency semi-circle (Fig. 4C and F) is rapidly decreased, suggesting
that the barrier properties of the coating is progressively decreased, and the corrosion rate continues to increase to the end
of the test.
Fig. 4 reveals that PAA-based treated samples exhibit more tendencies to negative direction. When an aluminium surface treated
with PAA is immersed in the test solution, as mentioned previously, un-reacted functional groups on the PAA polymeric chain

may dissociate in water. Dissociation in water might change the


pH of the solution to lower values. Such increased H+ concentration
may increase the aggressive behaviour of the solution and, as a result, the corrosion rate may increase.
From Fig. 4D for PAA/PAM/Zr treated Al samples, the spectra are
dominated by capacitance of the coating, which remained constant
even after more than 50 days exposure to 3.5% NaCl solution. It is
clear that the impedance values recorded for PAA/PAM/Zr treated
samples are higher than the impedance values obtained for de-

Fig. 2. Visual appearance of samples after 1000 h humidity test.

M. Niknahad et al. / Corrosion Science 52 (2010) 19481957

1953

Fig. 3. Visual appearance of samples after 1000 h salt spray test.

greased and PAA treated samples. PAA/PAM/Zr pretreated samples


reveal relatively good corrosion performance over the rst 50 days
of immersion in the electrolyte.
It is suggested [12] that zirconium-based treatments may protect the surface against the corrosive environment, by inhibition
of anodic reactions over the Al surfaces; consequently the passivity
of the metal increases [32].
A further possibility is that when a polymer is present in the
conversion coating solution, it may act as a surfactant and modify
uniformity of the conversion coating. PAA molecules may assist
production of a more uniform conversion-coating layer on the
surface.
It is clear that the impedance values recorded for PAA/PAM/Zr
treated aluminium alloy samples are higher than the impedance
values obtained for PAA/Zr treated samples. This is attributed to
the interactions between PAA and PAM molecules (as it is evident
from FTIR spectra analysis) which leaves only a few un-reacted
PAA molecules in the interface, and also the presence of zirconium
compound that increases the inhibition of anodic reactions on the
Al surfaces and in so doing improves the corrosion performance.
An equivalent circuit model, proposed by many authors
[26,33,34], based on the Nyquist plots for epoxy coated samples
is also shown in Fig. 5, either chromate or zirconium-based conversion treatments. This model reveals the electrolyte resistance, Rs,

the coating capacitance, Cc, the coating resistance, Rpf, the charge
transfer resistance, Rct, and the double layer capacitance, Cdl.
3.4. Adhesion strength measurements
The adhesion of coatings may be inuenced by and are many
factors surface pretreatments play an important role in the corrosion protection of aluminium alloys [35]. The adhesion strength of
differently treated Al samples is shown in Fig. 6. As shown in Fig. 6
the pull-off adhesion test revealed that samples Nos. 10 and 13 had
higher dry adhesion values compared with samples Nos. 4 and 16.
The adhesive strength of CPCC treated samples had an average value of 140 kg/cm2. This performance generally correlated with the
formation of an anchor pattern through which mechanical interlocking takes place.
There are indications [19,20] that good adhesion between a
high energy surface and the coating layer can occur through intimate contact between the substrate and the molecules of the coating. CPCC, PAA/PAM/Zr and PAA/Zr treated substrates have higher
surface energy than other treatments. Furthermore, as a result of
chemical interactions between PAA molecules and aluminium
oxide layer on the surface (as it is evident from FTIR spectra analysis), dry adhesion strength of epoxy coating increased
additionally.

1954

M. Niknahad et al. / Corrosion Science 52 (2010) 19481957

Fig. 4. Impedance spectra of epoxy clear coated aluminium substrates during 20 days immersion in 3% NaCl electrolyte.

In the dry state, all coated specimens indicated no failed regions, proving good adhesive strength. However, it is understood
that the adhesive strength of the coatings in the dry state does

not reect durability and performance [20]; thus, the adhesive


strength in the presence of water and possible corrosion products,
i.e., wet adhesion, was determined. Initial, dry adhesion has been

1955

M. Niknahad et al. / Corrosion Science 52 (2010) 19481957

Fig. 5. Equivalent circuit for the epoxy coated aluminium alloy substrates.

Adhesion Strength (kg/cm 2)

200

Fig. 7. EDS spectra of A: sample No. 10 pretreated with PAA/PAM/Zr and B: CPCC
treated sample.

mechanical stress on the adhesive bonds. Once the stress passes


a critical value, the bonds between reacted PAA molecules and
the aluminium oxide break, and the detached surface area increases [20].
The results revealed that PAM/Zr treatment has little effect on
the adhesion of coating to the aluminium substrate in wet or dry
condition. This is perhaps due to weaker chemical interaction
and lower surface energy of PAM/Zr treatment comparison with
PAA based treatments. It is believed that in treatment No. 10, there
are two forms of chemical interaction, between PAA molecules
and aluminium oxide, and PAA molecules and PAM molecules.

Dry Adhesion Strength (kg/cm2)


Wet Adhesion Strength (kg/cm2)
Adhesive Remaining (%)

100

160

80

120

60

80

40

40

20

Adhesion Remaining (%)

recognized for a long time as a poor predictor of coating performance in the presence of water or electrolyte [36].
Fig. 6 reveals that in the wet stage (immediately after one week
exposure in humidity condition), the adhesion strength of all specimens are decreased in comparison with dry stage. The main reason for failure of most organic coatings is the loss of bonding (if
any) between the coating and the substrate after exposure to the
humid atmosphere. When a coated aluminium surface is exposed
to an aqueous environment, water molecules penetrate through
the coating layer, after a certain initial period, water molecules
are accumulated in the coating/substrate interface, within the aluminium oxide layer. Mechanical interlocking or/and chemical
interactions throughout the interface may affected by this phenomenon and as a result, adhesion strength decreasing will be
detected.
For PAA-based treated samples, adhesion strength decreasing
may arise from the nature of the chemical interaction between
the treatments and aluminium surface. FTIR analysis has conrmed the conversion of carboxyl acid functional groups to carboxylate ions at the PAA/aluminium oxide interface. However, only
about 40% of the acidic functional groups on the polymeric chains
may convert to the salt form. It is proposed that steric and/or conformational hindrance, associated with the polymer backbone, prevent reaction of all carboxylic acid groups in PAA with surface
hydroxyl functionalities to form carboxylate complexes [12].
It is believed that un-reacted groups may act as weak sites at
the interface, therefore, when aluminium surface, on which PAA
based treatment was deposited, is exposed to aqueous environment, dissolution may proceed leaving little ionic or complex
interaction with the metal surface. The nal poor wet adhesion
at the local site is the result of the presence of un-reacted PAA molecules on the metal surface. These molecules may dissolve in
water, creating considerable osmotic pressures, with subsequent
increase in the water volume at the interface. This concludes to

0
No. 1

No. 4

No. 10

No. 13

No. 16

CPCC

Treatment
Fig. 6. Pull off adhesion strength of various pretreated epoxy coats on aluminium and percentage of adhesive remaining after humidity test.

1956

M. Niknahad et al. / Corrosion Science 52 (2010) 19481957

Therefore, un-reacted PAA molecules are very limited in the


coating/substrate interface, as a result, decreasing water sensitivity
and increasing adhesion strength in dry and wet conditions.
Tape test were performed under dry and wet conditions. In the
dry state all coated specimens indicated no failed regions, 5B, and

implying good adhesive strength. In the wet state CPCC treated


samples reveal improved performance compared with other samples. For comparison, the percentage of remaining adhesion was
n
 100, where AR repredetermined using the expression: AR % 25
sents adhesion remaining and n is the average number of squares
of undetached coatings. The results of ve individual tape adhesion
measurements, in wet condition, are shown in Fig. 6. The percentage adhesion remaining varied widely, with CPCC, PAA/PAM/Zr and
PAA/Zr treatments giving the highest levels.
Typical scanning electron micrographs and EDS spectra have
been shown previously [20,21] to contain zirconium, oxygen and
aluminium for sample treated with PAA and H2ZrF6. This is also
gained in Fig. 7. EDS spectrum (Fig. 7A) and SEM micrograph
(Fig. 8C) of sample treated with PAA/PAM/Zr tend to suggest that
a conversion-coating layer had developed over the aluminium surface. The EDS spectrum in particular illustrates the presence of zirconium, oxygen and aluminium indicating a thin lm over the
macroscopic alloy surface.
Figs. 7B and 8A, B and D show the respective EDS spectra of
CPCC as well as SEM micrograph of Zr, PAA/Zr and CPCC for comparison purposes. EDS analysis of PZr treated specimens revealed
the presence of zirconium, oxygen and aluminium, suggesting that
a thin lm had formed over the macroscopic alloy surface.

4. Conclusion
FTIR analysis showed that there are two possible chemical
interactions, one between PAA molecules and aluminium oxide
and the other between PAA and PAM. The rst is a purely ionic
3
giving an absorption band
interaction between COO and Al
1
around 1557 cm , and the second is related to the interactions between COO and NH
3 groups which appears as a doublet in the region 14081454 cm1. FTIR and EDS spectra indicate the formation
of a relatively thin and insoluble layer of the pretreatment on the
aluminium surface. EIS tests showed that the combination of
PAA/PAM/Zr as the pretreatment demonstrated a much better corrosion performance than other formulations even after the accelerated tests, ranked second after CPCC pretreatment.
Although the polyacrylamide component does not noticeably
increase dry adhesion strength, however, it markedly improves
the corrosion protection performance of the PAA/PAM/Zr pretreatment compared to the PAA/Zr treatment.
Acknowledgements
The authors wish to acknowledge support for the research work
reports in this paper from the Amirkabir University of Technology
and Iran Polymer and Petrochemical Institute.
References

Fig. 8. SEM micrographs of variously pretreated aluminium alloy samples,


magnication: 1000, 20 kV.

[1] S. Wernick, R. Pinner, P.G. Sheasby, The Surface Treatment and Finishing of
Aluminium and its Alloys, vol. 1, sixth ed., ASM International, USA/Finishing
Publications Ltd., UK, 2001.
[2] S. Lin, H. Shih, F. Mansfeld, Corrosion protection of aluminum alloys and metal
matrix composites by polymer coatings, Corros. Sci. 33 (1992) 13311349.
[3] G. Goeminne, H. Terryn, J. Vereecken, Characterisation of conversion layers on
aluminium by means of electrochemical impedance spectroscopy,
Electrochim. Acta 40 (1995) 479486.
[4] G.E. Thompson, G.C. Wood, in: J.C. Scully (Ed.), Treatise on Materials, Science
and Technology, vol. 23, Academic Press, New York, 1983.
[5] C.J.E. Smith, K.R. Baldwin, S.A. Garrett, M.C. Gibson, M.A. Hewins, P.L. Lane, The
development of chromate-free treatments for the protection of aerospace
aluminium alloys, aluminium surface science and technology, ATB Metall.
XXXVII (1997) 266.
[6] L.E.M. Palomino, I.V. Aoki, H.G. de Melo, Microstructural and electrochemical
characterization of Ce conversion layers formed on Al alloy 2024-T3 covered
with Cu-rich smut, Electrochim. Acta 51 (2006) 59435953.

M. Niknahad et al. / Corrosion Science 52 (2010) 19481957


[7] M. Mohseni, S.M. Mirabedini, M. Hashemi, G.E. Thompson, Adhesion
performance of an epoxy clear coat on aluminum alloy in the presence of
vinyl and amino-silane primers, Prog. Org. Coat. 57 (2006) 307313.
[8] D. Raps, T. Hack, J. Wehr, M.L. Zheludkevich, A.C. Bastos, M.G.S. Ferreira, O.
Nuyken, Electrochemical study of inhibitor-containing organicinorganic
hybrid coatings on AA2024, Corros. Sci. 51 (2009) 10121021.
[9] S. Crips, H.J. Prosser, A.D. Wilson, An infra-red spectroscopic study of cement
formation between metal oxides and aqueous solutions of poly(acrylic acid), J.
Mater. Sci. 11 (1976) 3648.
[10] Y. Grohens, J. Schultz, PMMA conformational changes on c-alumina powder:
inuence of the polymer tacticity on the conguration of the adsorbed layer, J.
Adhes. Adhes. 17 (1997) 162167.
[11] M.A. Romero, B. Chabert, A. Domard, IR spectroscopy approach for the study of
interactions between an oxidized aluminium surface and a poly(propylene-gacrylic acid) lm, J. Appl. Polym. Sci. 47 (1993) 543555.
[12] M.R. Alexander, S. Payan, T.M. Duc, Interfacial interaction of plasma
polymerized acrylic acid and an oxidized aluminium surface investigated
using XPS, FTIR and polyacrylic acid as a model compound, Surf. Interface Anal.
26 (1998) 961973.
[13] P.D. Deck, M. Moon, R.J. Sujdak, Investigation of uoroacid based
pretreatments on aluminium, Surf. Coat. Int. 10 (1998) 478485.
[14] S. Chibowski, E. Opala Mazur, J. Patkowski, Inuence of the ionic strength on
the adsorption properties of the system dispersed aluminium oxide
polyacrylic acid, J. Mater. Chem. Phys. 93 (2005) 262271.
[15] S. Chibowski, M. Knipa, Studies of the inuence of polyelectrolyte adsorption
on some properties of the electrical double layer of ZrO2electrolyte solution
interface, J. Dispersion Sci. Technol. 21 (2000) 761783.
[16] J. van den Brand, S. Van Gils, P.C.J. Beentjes, H. Terryn, J.H.W. de Wit, Ageing of
aluminium oxide surfaces and their subsequent reactivity towards bonding
with organic functional groups, J. Appl. Surf. Sci. 235 (2004) 465474.
[17] J. van den Brand, S. Van Gils, P.C.J. Beentjes, H. Terryn, V. Sivel, J.H.W. de Wit,
Improving the adhesion between epoxy coatings and aluminium substrates, J.
Prog. Org. Coat. 51 (2004) 339350.
[18] D. Chidambaram, C.R. Clayton, G.P. Halada, The role of hexauorozirconate in
the formation of chromate conversion coatings on aluminum alloys,
Electrochim. Acta 51 (2006) 28622871.
[19] S.M. Mirabedini, S. Moradian, Relationship between adhesive strength and
surface energy, in: Sixth Iranian Seminar on Polymer Science and Technology
(ISPST, 2003), 1215 May 2003, Tehran, Iran.
[20] S.M. Mirabedini, The role of interfacial layer on the performance of powder
coated aluminium alloy, PhD Thesis, Corrosion and Protection Center: UMIST,
UK, 2000.

1957

[21] S.M. Mirabedini, G.E. Thompson, S. Moradian, J.D. Scantlebury, Corrosion


performance of powder coated aluminium using EIS, Prog. Org. Coat. 46 (2003)
112120.
[22] Yawei Shao, Cao Jia, Guozhe Meng Tao Zhang, Fuhui Wang, The role of a zinc
phosphate pigment in the corrosion of scratched epoxy-coated steel, Corros.
Sci. 51 (2009) 371379.
[23] Y. Huang, H. Shih, H. Huang, J. Daugherty, S. Wu, S. Ramanathan, C. Chang, F.
Mansfeld, Evaluation of the corrosion resistance of anodized aluminum 6061
using electrochemical impedance spectroscopy (EIS), Corros. Sci. 50 (2008)
35693575.
[24] F. Mansfeld, M.W. Kendig, Impedance spectroscopy as quality control and
corrosion test for anodized aluminum alloys, Corrosion 41 (1984) 490492.
[25] G.W. Walter, Laboratory simulation of atmospheric corrosion by SO2-II.
Electrochemical mass loss comparisons, Corros. Sci. 32 (1991) 13531359.
[26] S. Lin, H. Shih, F. Mansfeld, Corrosion protection of aluminum alloys and metal
matrix composites by polymer coatings, Corros. Sci. 33 (1992) 13311349.
[27] M. Kendig, F. Mansfeld, S. Tsai, Determination of the long term corrosion
behavior of coated steel with A.C. impedance measurements, Corros. Sci. 23
(1983) 317329.
[28] R. Ma, S.M. Mirabedini, R. Naderi, M.M. Attar, Effect of curing characterization
on the corrosion performance of polyester and polyester/epoxy powder
coatings, Corros. Sci. 50 (2008) 32803286.
[29] F. Mansfeld, M.W. Kendig, S. Tsai, Evaluation of corrosion behavior of coated
metals with AC impedance measurements, Corros. NACE 38 (1982) 478485.
[30] J.D. Scantlebury, The dynamic nature of underlm corrosion, Corros. Sci. 35
(1993) 13631366.
[31] G.D. Davis, T.S. Sun, J.S. Adhearn, J.D. Venables, Application of surface
behaviour diagrams to the study of hydration of phosphoric acid-anodized
aluminium, J. Mater. Sci. 17 (1982) 18071818.
[32] J. Nelson, Newhard Jr., in: H. Leidheiser Jr. (Ed.), Corrosion Control by Coatings,
Science Press, Princeton, 1978, pp. 225241.
[33] J.M. Sykes, 25 years of progress in electrochemical methods, Br. Corros. 25
(1993) 175183.
[34] G.W. Walter, A review of impedance plot methods used for corrosion
performance analysis of painted metals, Corros. Sci. 26 (1986) 681703.
[35] J.B. Bajat, V.B. Mikovic-Stankovic, Z. Kacarevic-Popovic, Corrosion stability of
epoxy coatings on aluminum pretreated by vinyltriethoxysilane, Corros. Sci.
50 (2008) 20782084.
[36] R.A. Dickie, in: K.L. Mittal (Ed.), Adhesion Aspects of Polymeric Coatings,
Plenum Press, New York, 1983, p. 319.

Вам также может понравиться