Está en la página 1de 12

Progress in Organic Coatings 52 (2005) 339350

Comparative electrochemical studies of zinc chromate and


zinc phosphate as corrosion inhibitors for zinc
A.C. Bastosa , M.G.S. Ferreiraa,b , A.M. Simoesa,
a

Chemical Engineering Department, Instituto Superior Tecnico, Av. Rovisco Pais, 1049-001 Lisboa, Portugal
b Departamento de Engenharia de Cer
amica e do Vidro, Universidade de Aveiro, Aveiro, Portugal
Received 28 May 2004; received in revised form 9 September 2004; accepted 9 September 2004

Abstract
The anticorrosive performance of two inhibitive pigments, zinc chromate and zinc phosphate, was compared using electrochemical
impedance spectroscopy (EIS) and the scanning vibrating electrode technique (SVET) in pigment extracts in 0.1 M NaCl. It was observed
that zinc was protected from corrosion in both extracts. In tests using hot dip galvanised steel painted with an epoxy primer incorporating the
pigments, the SVET detected the anodic and cathodic distribution along the scribes, although no significant differences were observed among
the various primers. On the contrary, EIS was able to distinguish processes occurring on the metal surface exposed by the scribe in different
samples. For primers with anticorrosive pigment, a time constant at high frequencies was attributed to a layer of protective nature, probably
formed by metal ions from the substrate and inhibitive ions leached from the anticorrosive pigments.
2004 Elsevier B.V. All rights reserved.
Keywords: Anticorrosive pigments; Corrosion; EIS; SVET; Zinc; Phosphate; Chromate

1. Introduction
Decades of industrial practice have lead to the selection
of a few pigments with excellent anticorrosive properties in a
wide range of situations. Chromates have long been the first
choice for many applications, but they represent environmental and health threats that are leading to their replacement.
Among the new non-toxic anticorrosive pigments that today
line up as substitutes for traditional pigments, zinc phosphate
is probably the most widely accepted, and it is incorporated
in many paint formulations.
Although not fully understood, the inhibitive action of
chromate pigments is now well accepted to be based on the
chromate ions that leach out to solution, where they act in the
same way as other chromate soluble inhibitors. In solution,
chromate ions are present in the hexavalent form and become
reduced to the trivalent form to counterbalance the anodic
oxidation of the metal substrate. It is the trivalent chromium

Corresponding author.
E-mail address: alda.simoes@ist.utl.pt (A.M. Simoes).

0300-9440/$ see front matter 2004 Elsevier B.V. All rights reserved.
doi:10.1016/j.porgcoat.2004.09.009

that is usually found in the passive layer. On iron, films of


oxide spinels with Cr3+ and Fe3+ have been reported [13].
The Cr/Fe atomic ratio seems to depend on the pH, oxygen
and chromate content of the solution. Other authors describe
the film composition as a mixture of oxides and hydroxides
of iron and chromium [46]. Hydroxylated layers may be
present [2] and adsorbed Cr(VI) has also been observed [7].
On aluminium alloys, the action of chromate has also been
studied and an inhibiting effect on the oxygen reduction reaction at the copper precipitates was observed [8]. Studies of
chromate passive layers on zinc are scarce but a recent study
reports a passive film of Cr(III) with the absence of zinc [9].
Zinc yellow, being a mixed salt of zinc chromate, potassium
chromate and zinc hydroxide, has, in addition to CrO4 2 ,
two more ionic species of inhibitive nature, Zn2+ and OH .
It has been referred that for the same concentration of chromate, zinc yellow has a high inhibitive power compared to
other simple chromate salts [10]. OH leads to a rise in pH,
decreasing the corrosivity, particularly at the anodic sites,
whereas Zn2+ precipitates as Zn(OH)2 at the cathodic sites.
The protective layers formed on the metal prevent either the

340

A.C. Bastos et al. / Progress in Organic Coatings 52 (2005) 339350

release of metal ions from the surface towards the solution


or the oxygen or proton reduction, by, for instance, impeding
the flow of electrons needed for such reactions. A hydrophobic character has also been proposed for the Cr(III) oxide
layers, as well as a lower zeta potential of oxide layers when
chromate is adsorbed, hindering the adsorption of aggressive
anions such as chloride [11].
Phosphate has been used for decades as soluble inhibitor
and also in conversion coatings. The use of zinc phosphate as
pigment was proposed for the first time in 1965 [12]. Several
modes of action have been proposed, including the formation of soaps of zinc [13], in addition to an improvement of
the binder barrier properties [14]. Studies with soluble phosphates report phosphate layers at both the anodic [15] and
cathodic regions [16]. Tertiary phosphates precipitate with
iron and also with zinc, forming a protective layer that hinders the access of oxygen or the oxidation of the base metal. A
neutralizing ability, capable of inhibiting corrosion by controlling the pH, was also suggested [10]. Real exposure of
paints with zinc phosphate has led to very good results, contrasting with accelerated tests where the results were quite
poor, possibly due to the slow kinetics of the reactions or to
the low solubility of the pigment [17]. The pH of the solution seems to play a major role, since the solubility of zinc
phosphate is higher in acidic media [18].
In this work, two anticorrosive pigments were studied: the
classical basic zinc potassium chromate (zinc yellow) and the
new alternative, zinc phosphate. The pigments were tested
either incorporated in an epoxy primer or by preparing aqueous extracts. A two-component epoxy primer was selected
and three variations were studied: a commercial zinc phosphate primer, a modification of the former by replacing the
amount of phosphate by chromate, and a clear coat, based
on the primer formulation but without any pigments or extenders. Comparative electrochemical tests were performed
using electrochemical impedance spectroscopy (EIS) and the
scanning vibrating electrode technique (SVET). EIS is a wellestablished technique in corrosion research; it can work with
either coated or uncoated samples and its capability to differentiate the processes occurring in the electrochemical system
makes it suitable for this kind of study. The SVET measures
potential differences in solution, created by uneven ionic distributions. By a prior calibration, the measured potential values are converted to currents. The SVET technique was introduced to the field of corrosion in the early 1970s [19] and
some work has been published since then [2025].

mounted (Fig. 1a), whereas for the SVET measurements, zinc


samples were epoxy mounted, leaving exposed a rectangle of
1 mm2 (Fig. 1b). The specimens were polished with SiC grit
papers of grades 220, 500, 800 and 1000, washed in deionized water (Millipore), degreased with ethanol and dried with
compressed air.
2.2. Coated samples
Three primers were used: a two-component epoxy primer
with zinc phosphate (commercial product), an epoxy primer
with addition of chromate (modification of the commercial
primer by replacing the phosphate pigment by zinc yellow)
and the corresponding epoxy clear coat. The primers were
applied to hot-dip galvanised steel (HDG) by dipping the
substrates into the paint for 1 s and leaving them to dry.
The tests were performed 1 month after application of the
coating. Dry film thickness measured using an Elcometer 355
digital thickness gauge was 40 m for EIS measurements
and 20 m for SVET measurements. The smaller thickness
in the samples for the SVET study was chosen in order to
allow a closer approach of the tip to the metal surface at the
scribe. For the SVET measurements, squares of 2 cm 2 cm
were cut, scribed and glued to an epoxy sample holder and
the edges and sides isolated with epoxy adhesive, as shown
in Fig. 1c. For EIS, a scribe of 1 cm long was made using a
scalpel, across the coating and down to the substrate; a plastic
cylinder was then glued to the surface, exposing an area of
3.80 cm2 , as shown in Fig. 1d.
2.3. Exposure media
Pigment extracts were made based on existing literature
[26], by magnetic stirring 1 g of pigment in 500 mL of 0.1 M
NaCl for 24 h, followed by filtering twice. The pigments were
the same as used in the paints, i.e., zinc chromate known
as zinc yellow (K2 CrO4 3ZnCrO4 Zn(OH)2 2H2 O) and zinc
phosphate (Zn3 (PO4 )2 2H2 O). Table 1 shows some details of
the solutions used.
2.4. Electrochemical techniques
SVET measurements were made using Applicable Electronics Inc., equipment. Tape glued around the sample holdTable 1
Characterization of the exposure media (T = 20.0 C)

2. Experimental
2.1. Bare metal samples
Zinc 99.95% pure, in the shape of 1-mm thick foil
(produced by Goodfellow Ltd., UK) was used. For EIS,
1 cm 1 cm samples were electrically connected via a multithread copper wire with silver conductive adhesive and epoxy

0.1 M NaCl
Chromate extract
Phosphate extract
a
b
c
d

pHa

Conductivityb
(mS cm1 )

Concentration

6.20
6.55
6.43

9.44
10.90
9.46

|Cr|total = 1.1 102 Mc


|PO4 3 | = 4.8 105 Md

Metrohm 632 pH meter, electrode 6.0220.100.


Crison GLP 31 conductimeter.
Inductively coupled plasma.
Molecular absorption spectrometry.

A.C. Bastos et al. / Progress in Organic Coatings 52 (2005) 339350

341

Fig. 1. Samples used in the experiments. (a) Mounted bare zinc for EIS measurements in extracts, (b) mounting for SVET measurements of bare zinc in extracts,
(c) coated sample glued with epoxy for SVET measurements (d) cell for EIS measurements on scribed coated samples.

ers worked as solution reservoir. The microelectrode had a


platinum tip of 40 m and was made to vibrate at an average
distance of 200 m above the surface, with 40 m of amplitude. Each scan comprised 20 20 points and an optical
image was acquired before each scan. Although the SVET
equipment used can detect ionic currents in two directions,
parallel and normal to the surface, only results from the flux
normal to the surface will be presented. In general, for an
active metal immersed in an electrolytic solution, the diffusion of metal cations from the metal oxidation in the anodic
regions is detected as a positive current, whereas negative currents correspond to OH emerging from the cathodic regions
as a product from oxygen reduction.
EIS measurements were made using Gamry instrumentation. The three-electrode cell consisted of the sample as
working electrode, a platinum counter electrode and a Saturated Calomel Electrode (SCE) as reference. Measurements
were made inside a Faraday cage, at room temperature and the
solution was quiescent and exposed to air. Usually, a 50 kHz
to 5 mHz frequency range was swept with a sinusoidal potential signal with an amplitude 10 mV rms, superimposed to
the open circuit potential. Fitting was made using ZView 2.70
software and all the spectra were corrected to the geometrical
area of the working electrode.

Measurements of the open circuit potential (OCP) were


made with an AUTOLAB PGstat 20 apparatus and are plotted
against the potential of the SCE reference.

3. Results
3.1. Bare metal electrodes
SVET current mapping made on pure zinc immersed in
sodium chloride solution shows significant differences. In
the salt solution without pigment, activity occurred always
with the development of one single anode and one single
cathode (Fig. 2a). Each of these electrodes occupied nearly
half of the surface in the first minutes of immersion. With
time, the anodic area became more localized (Fig. 2b) until,
after approximately 1 day, a pit was observed on the surface
(Fig. 2c and d).
When the NaCl solution contained the pigment extracts
the current inhibition was obvious. The current values measured on the surface were significantly reduced and there
was only weak indication of cathodic or anodic activity, with
the pattern of the surface corresponding to scattered microscopic and weak activity, corresponding to currents below

342

A.C. Bastos et al. / Progress in Organic Coatings 52 (2005) 339350

Fig. 2. SVET maps of the ionic currents measured 200 m above pure zinc exposed to: 0.1 M NaCl (ac); zinc phosphate extract in 0.1 M NaCl (eg); zinc
chromate extract in 0.1 M NaCl (ik). Times of immersion are (top to bottom): 5 min, 1 h and 1 day. Scanned area: 1.25 mm 1 mm; Scale units: A cm2 .
Also shown are pictures of the surface after 1 day of immersion in each solution (d, h, l).

6 A cm2 with the phosphate (Fig. 2eg) and 2 A cm2


with the chromate (Fig. 2ik). Inhibition by the phosphate
required a longer time before its full action was felt, i.e.,
before the phosphate film was formed, since some anodic
activity was measured in the first few minutes of immersion. For longer exposure times the activity remained negligible with both extracts (Fig. 2g and k), as confirmed by
the inspection of the surface at the end of the test (Fig. 2h
and l).
The open circuit potential of pure zinc in 0.1 M NaCl
was practically constant and near the equilibrium potential
of zinc (Fig. 3). With the phosphate, the potential was only
slightly higher, although at the beginning it started from a
more anodic value and then rose during a period of approximately 3 h, which probably corresponds to the growth of
the protective film. In contrast, the chromate pigment has
led to a significant anodic polarization, reaching potentials

Fig. 3. Open circuit potential of pure zinc in 0.1 M NaCl, either uninhibited
or containing phosphate extract or chromate extract.

A.C. Bastos et al. / Progress in Organic Coatings 52 (2005) 339350

343

Fig. 4. Impedance spectra obtained for pure zinc at different times of exposure to 0.1 M NaCl, without inhibitor (a), or containing phosphate extract (b) or
chromate extract (c); (a1), (b1) and (c1) are Bode plots, whereas (a2), (b2) and (c2) are the corresponding Nyquist plots.

of 0.945 V. This value is within the region of stability of


chromium oxides/hydroxides, which are known to be the
major components of the passivating film. The fluctuations
observed in the phosphate curve, in opposition to the stable curve for the chromate, indicate differences in reactivity
between the two protective films, and agree with the current density measurements with the SVET. After a stationary
state is achieved, the order of the corrosion potential values
agrees with the SVET observations and follows the expected
order of relative protective efficiency of the pigments (chromate > phosphate > reference system).
The impedance response of pure zinc exposed to the three
solutions is presented in Fig. 4. In the absence of inhibitor, the
spectrum has two relaxation constants, one well developed
at high frequencies and a small one at lower frequencies.
The low-frequency process was clearly resolved in the first
minutes of immersion but disappeared after 3 h of exposure,
which means that it may be associated with the presence
of an air-formed film of oxides that becomes dissolved in
the chloride solution. Meanwhile, the high frequency pro-

cess became shifted to lower frequencies, which corresponds


to a rise of the double layer capacitance as corrosion progresses (possibly due to increasing active area). The spectra
obtained for pure zinc in NaCl solution is identical to the
spectrum modelled by Titz et al. [27] for a metal covered
with a porous oxide layer, in which corrosion proceeds under oxygen diffusion control. In the presence of the extracts
the spectra are simpler, with only one maximum in the phase
angle plot, although the slope in the impedance modulus at
the low frequencies suggests the presence of a second time
constant. The capacitance is smaller than described above,
revealing the inhibiting effect of the pigments, and does not
change with time. The low frequency resistance, which was
1 k cm2 in the NaCl solution, increased in the presence of
the extracts, confirming a slower corrosion rate. The Nyquist
plots, included in Fig. 4 for a better visualization of the results, also show the difference in behaviour among the three
solutions; in the absence of inhibitor the spectrum has two
semi-circles, the first of which is well-defined and reveals the
corrosion process, whereas in the presence of the extracts the

344

A.C. Bastos et al. / Progress in Organic Coatings 52 (2005) 339350

Fig. 5. General equivalent circuit and fittings for pure zinc in 0.1 M NaCl without inhibitor (a) and with zinc chromate extract (b). Fitting parameters for (a) Rs = 34.4  cm2 ; Y0 (Q1 ) = 1.62 105 F cm2 sn1 ; n(Q1 ) = 0.872; R1 = 1.31 103  cm2 ; Y0 (Q2 ) = 6.06 104 F cm2 sn1 ; n(Q2 ) = 1;
R2 = 631  cm2 ; 2 = 3.9 104 . Fitting parameters for (b) Rs = 26.2  cm2 ; Y0 (Q1 ) = 3.31 106 F cm2 sn1 ; n(Q1 ) = 0.903; R1 = 1.74 105  cm2 ;
Y0 (Q2 ) = 6.51 106 F cm2 sn1 ; n(Q2 ) = 0.500; R2 = 1.14 106  cm2 ; 2 = 1.0 103 .

semi-circle is always incomplete and reaches much higher


values in the scale.
For the equivalent circuit model fitting of the spectra, as
shown in Fig. 5, the capacitance was replaced by a constant phase element (CPE) with an impedance given by
Z = 1/Y0 (j)n . Provided the exponent n is close to 1, the values
of Y0 can be taken as a measure of the capacitance. Although
the spectra in Fig. 5 are apparently quite different, the fitting
was made using the same equivalent circuit, generically having a high frequency time constant, described by R1 and Q1
and a low-frequency time constant (R2 , Q2 ). The values of
these electrical components, as well as their physical meaning, change however between the two systems. In NaCl, the
high-frequency time constant corresponds to the charge transfer resistance and double layer capacitance associated with
the corrosion process, whereas the low-frequency process,
with a capacitance of approximately 600 F cm2 , can be
attributed to the mass transport across the porous air-formed
film of oxides. As described above, this low-frequency process ceased to be detected a few hours after immersion. The
spectrum in Fig. 5(b), which is representative of both the
situations of chromate and of phosphate extracts, can be described by the same circuit, although only R1 and Q1 are
clearly resolved; a closer observation shows in fact the presence of a secondary slope in the impedance modulus plot,
below 101 Hz.

The time evolution of the fitted parameters corresponding


to the three systems is presented in Fig. 6. The charge transfer
resistance, R1 , was in the range of 103  cm2 in 0.1 M NaCl,
104  cm2 in the phosphate extract and 105  cm2 in chromate. In the NaCl solution, the resistance decreased after the
first hours, possibly due to the activation of the zinc surface as
the air-formed film became dissolved. With the pigments R1
was higher from the first measurements, meaning that at least
a monolayer of film was formed in the first instants of immersion; the slight growth of R1 with time reveals no change of
the behaviour in the film, but rather the build-up of the film
on the surface.
The value of Y0 increased for zinc in 0.1 M NaCl to nearly
1 mF, a value that suggests some influence of mass transfer across the gel-like layer of corrosion products. The values of Y0 were almost constant both in phosphate and in the
chromate extracts, with values below that of a typical double
layer capacitance, and with a slight tendency to decrease with
time. The high values of the exponent n reveal that the CPE
corresponds to a nearly capacitive response, as expected. In
physical terms, Y0 can be either associated with the capacitance of a compact dielectric oxide film or it may correspond
to the double layer capacitance over the fraction of area left
uncovered by the passivating film, as discussed by Zeller and
Savinell [28]. According to this last interpretation, a fraction
of coverage of 99% can be estimated from the capacitance

A.C. Bastos et al. / Progress in Organic Coatings 52 (2005) 339350

345

Fig. 6. Fit parameters for the impedance spectra obtained for pure zinc at different times of exposure to 0.1 M NaCl only, to phosphate extract in 0.1 M NaCl
and to chromate extract in 0.1 M NaCl.

values for any of the two pigments, after 25 h of immersion.


Estimation of the inhibition efficiency by the charge transfer
resistance, R1 , gives an inhibition of 90% with the phosphate and 99% for the chromate.
3.2. XPS inspection
XPS analysis of pure zinc after immersion in the phosphate and chromate extracts revealed well-defined peaks of
phosphorous and chromium, respectively. Cr was present as
a mixture of Cr(III) and Cr(VI). Zinc was detected by a single Zn 2p3/2 peak, at 1022.7 0.3 eV, corresponding to ZnO
or Zn(OH)2 . The O 1s peak was measured at 532 eV and is
probably due to OH , which means that both the zinc and the
chromium are present mainly as hydroxides. In the samples
exposed to the phosphate extract, a P 2p3/2 peak appeared at
140.2 eV.
3.3. Scribed samples
EIS measurements were performed on samples of Hot Dip
Galvanized Steel (HDG), with the various paint formulations
and with a scribe reaching the zinc layer. After 1 h of immersion, the response was strikingly similar for all the coatings
(Fig. 7), and also identical to the spectra obtained for the
bare metal exposed to plain NaCl solution (Fig. 4a). The
main difference is related to the higher impedance values,
explained by the smaller ratio of the area exposed to the total
sample area, roughly estimated as 1:501:100. The properties of the primer film, typically observed at the high frequencies, were not detected, probably because of the high
impedance of the film; being in parallel with the scribe and
having a much higher impedance, the coating does not allow
any significant flow of current, and therefore it does not contribute to the impedance spectrum in the range of frequencies
measured.
After 24 h of free corrosion, the impedance was lowest in
the sample without pigment, which reveals the higher activity
of this system, and another time constant appeared at the high

frequencies. Fitting of the spectra was made using the equivalent circuit in Fig. 8, where the high frequency impedance
(1 k cm2 ) is due to the electrolytic solution at the scribe.
RHF and QHF describe the high frequency process, and Rct
and Qdl correspond to the charge transfer resistance and the
double layer capacitance, respectively. Finally, a component
describing the mass transfer impedance was considered in
some cases in order to account for a small tail appearing below 0.1 Hz. The high frequency process had a resistance of
510 k cm2 , as shown in Fig. 9, whereas the Y0 for this
process has a value of roughly 108 to 105 F cm2 . This
capacitance is too high for a polymer film, which generally
has values in the range of 1010 to 108 F cm2 [27,29].
Further, in the presence of chromate Y0 decreases with time,
a behaviour that is the opposite to the known evolution of
coating capacitance values (which tend to increase due to
water uptake). This high frequency time constant can however be explained by the formation of the passivating film, of
either chromium hydroxide or zinc phosphate. The formation
of this film starts in the first minutes of exposure, possibly
with the precipitation of a monolayer, and grows afterwards,
during the first 2448 h. The low frequency process gives
Rct = 2080 k cm2 and is maximum for the chromate. The
double layer CPE grows with time, revealing the growth of
the active area underneath the coating, i.e., delamination. The
ratio of Y0 with and without pigment is roughly 0.1, meaning that the loss of adhesion was weaker in samples with
pigment. The delamination rate, estimated from the CPE,
was minimum in the epoxychromate coating, followed by
the epoxyphosphate and finally by the clear coat. This was
the same rate that was determined by visual inspection of the
samples at the end of the test.
The SVET was also used to study the ionic current distribution above the scribed samples, for the three coatings
in 0.1 M NaCl. The evolution of the current maps and the
current density values have not show any significant differences among the specimens with the various coatings. As an
example, Fig. 10 shows SVET measurements of the scribed
HDG sample, coated with the clear epoxy paint. In the first

346

A.C. Bastos et al. / Progress in Organic Coatings 52 (2005) 339350

Fig. 7. Bode and Nyquist impedance plots of scribed samples of HDG, coated with epoxy primer with pigment, after 1 h (a) and 1 day (b) of exposure in 0.1 M
NaCl. (a1) and (b1): Bode plots; (a2) and (b2): corresponding Nyquist plots.

23 days of immersion there was a clear separation between


the cathode and the anode along the scribe (Fig. 10a) and
there were no signs of delamination in the micro-video image. For longer exposure times, however, the anodic activity
along the scribe increased and the signs of cathodic activity
disappeared from the map, as delamination progressed away
from the scribe (Fig. 10b).

4. Discussion
The inhibiting power of anti-corrosive pigments depends
upon a number of parameters, namely the pigment solubility,

the pH, the properties of the substrate and the composition of


the solution. The levels of inhibitor concentration achieved
in the extracts were much lower for the phosphate than for
the chromate, although in both cases the inhibiting effect has
proved to be sufficient to provide high inhibition efficiency
to zinc in 0.1 M NaCl. In both cases, inhibition was the result
of the formation of a layer of inhibitor on the surface. In the
chromate, the film was formed instantly after immersion. Inhibition by the phosphate was a slower process, as confirmed
by the SVET and the OCP measurements. The chromate was
also the more efficient inhibitor. On the bare metal, it led to
the highest charge transfer resistance and also to the development of a protective layer characterized by a capacitance

A.C. Bastos et al. / Progress in Organic Coatings 52 (2005) 339350

347

Fig. 8. Equivalent circuit and fitting of impedance spectra for the scribed samples of HDG in 0.1 M NaCl with chromate primer (a) and with
phosphate primer (b). Fitting parameters for (a) Rscribe = 775  cm2 ; Y0 (QHF ) = 3.05 108 F cm2 sn1 ; n(QHF ) = 0.844; RHF = 9.04 103  cm2 ;
Y0 (Qdl ) = 6.76 107 F cm2 sn1 ; n(Qdl ) = 0.630; Rct = 7.33 104  cm2 ; 2 = 5.2 104 . Fitting parameters for (b) Rscribe = 990  cm2 ;
Y0 (QHF ) = 1.20 105 F cm2 sn1 ; n(QHF ) = 0.551; RHF = 5.47 103  cm2 ; Y0 (Qdl ) = 6.79 106 F cm2 sn1 ; n(Qdl ) = 0.768; Rct = 2.73 105  cm2 ;
2 = 2.2 104 .

decrease of approximately two orders of magnitude, when


compared with the reference solution.
When these pigments are used in a coating, leaching to
the solution is the first step towards inhibition. Since this is
a slow process, inhibition can only be detected after a sufficient concentration of inhibitor is achieved near the metal.
The mechanisms of passivation by zinc chromate are usually
described by an oxidizing action of CrO4 2 . In the process
of film formation, Cr(VI) is reduced to Cr(III), while Zn is
oxidized to Zn(OH)2 , according to the reaction scheme
3Zn + 2CrO4 2 + 5H2 O 3Zn(OH)2 + Cr2 O3 + 4OH
Chromate is usually considered to act especially at the damaged areas of the coating, where the electropositive metals
are left exposed to the aggressive medium, and are therefore
available for providing free electrons for chromate reduction.
For the conditions of our experiments and for the duration of
the tests, we found none of the pigments had the power to
inhibit the corrosion occurring at the defect.
An interesting feature was related with the development
of a time constant at the high frequencies. Unlike the typical spectra observed with defective painted samples, the high
frequency process could not be attributed to the dielectric
properties of the coating, due to its high capacitance. On the
other hand, it cannot be simply due to the precipitation of corrosion products, or it would have to be even more relevant in

the clear coat. This process has then to be due to a passivating


layer. The SVET, however, has shown that the anodic activity
occurred at the scribe irrespective of the pigment. This passivating layer, therefore, can only be underneath the coating,
where it inhibits the cathodic reaction. Chromate pigments
are usually classified as oxidizing inhibitors, because of the
strong thermodynamic tendency of Cr(VI) to become reduced
to Cr(III). It is claimed to inhibit steel corrosion at concentrations above 104 M [30]. According to our results, the precipitation of this layer seems to occur preferentially underneath
the coating, as water permeates across the coating and leaches
some pigment, and then at the scribe, after diffusion of the
dissolved pigment. At the pH values expected underneath the
coating, chromate acts as mixed anodiccathodic inhibitor,
with an efficiency that is only slightly smaller than at neutral
pH [26]. On the other hand, at this pH the thickening of Zn
oxide layers is easier. Therefore, the film probably consists of
a mixture of chromium and zinc oxides and hydroxides. Both
zinc and chromium oxides are know to behave as semiconductors, for which the capacitance is inversely proportional
to the thickness:
C = 0

A
L

where is the relative permitivity, 0 the permitivity of vacuum, A the area and L is the film thickness. Provided the
dielectric properties of the layer remain constant, the drop

348

A.C. Bastos et al. / Progress in Organic Coatings 52 (2005) 339350

Fig. 9. Fit parameters of the impedance spectra of scribed painted HDG at different times of exposure to 0.1 M NaCl.

in capacitance means that the thickening of the film is more


important than its growth along the surface, i.e., as if a monolayer of Cr(OH)3 was formed as soon as the water permeated
across the thin polymer coating, through leaching of a small
amount of chromate. After that, the capacitance decreased as
the result of film thickening.
The low frequency semicircle in the impedance spectrum, which is actually related to the corrosion process,
gives the actual rate of delamination, and the ranking is
the same as observed with the pigments extracts, i.e., chromate > phosphate > clear coat.
4.1. SVET
SVET mapping at the scribes showed one large anode and
one large cathode at the scribe in the first days of immersion.
The fact that both the anode and the cathode are located at
the scribe shows that the scribe itself constitutes a complete
corrosion cell, without the need for further electrochemical
reactions to balance the electrons. As already observed by
Bierwagen and co-workers [23], for longer immersion times
the cathodic currents cease to be detected by the SVET. This
reveals that the cathodic reaction takes place away from the
scribe, i.e., underneath the coating, and is thus a distinct sign

of the delamination process. Since the cathodic and anodic reactions have to balance each other, then a large ratio between
the cathodic and the anodic areas will lead to a small density
of the cathodic current. The cathodic current is essentially
due to the reduction reaction of oxygen:
O2 + 2H2 O + 4e 4OH
and occurs at the metalcoating interface, thus leading to an
increase of pH (responsible for the rupture of coating metal
bonds and/or to saponification of the coating) but also to an
accumulation of anions in a confined region. Compensation
of this excess of anions can be achieved either by the outward flow of the hydroxyl anions themselves or, more easily
by the inward flow of cations. As pointed out by Leidheiser
and Wang [31], the high concentration of Na+ in the outer
solution makes them the most likely cations to migrate to
the front of the delamination. Migration of sodium cations
to the delamination front was confirmed by Stratmann, who
found an excess of sodium in the delaminated area but also
in a thin area considered to be beyond the delamination front
[32]. This migration of cations can occur in two ways: either across the coating or from the scribe. This last possibility, however, would mean that other cations would migrate
from the bulk into the scribe, generating a significant ionic

A.C. Bastos et al. / Progress in Organic Coatings 52 (2005) 339350

349

Fig. 10. SVET maps of ionic currents measured 200 m above a scribe in HDG coated with clear epoxy paint, after (a) 2 days and (b) 7 days of exposure to
0.1 M NaCl. Also presented are micro-video images of the corresponding sample area.

flow above the scribe, equivalent to that of Zn2+ ions flowing


away from the surface. Such a flow should be detected by the
SVET, in the same way as anodic currents are detected. In
contrast, for a flow of ions across the coating, the cross area
for the flow is much larger, leading to very small potential
gradients above the coating, which are necessarily difficult to
detect.

5. Conclusions
Protection of zinc by pigments of zinc chromate and zinc
phosphate was monitored by EIS, SVET and OCP, with good
agreement between the three techniques.
Evaluation of the electrochemical parameters for the bare
zinc in pigment extracts allowed the ranking of the decreasing
corrosion resistance in the extracts in 0.1 M NaCl as: chromate > phosphate > plain salt solution. Interpretation of the
impedance spectra of the scribed coatings has shown that the
equivalent circuit is identical to that of the bare metal in a
first stage, whereas the processes occurring underneath the
coating start to appear in a second stage. A high frequency
relaxation constant observed in the scribed, pigmented coatings, was assigned to the formation of a protective layer of
pigment underneath the coating. The decreasing rate of delamination, taken as the rate of growth of the double layer
capacitance, was the same, i.e., chromate > phosphate > clear
coat.

Mapping of the local ionic fluxes above the metal surface


has shown that zinc corrosion in NaCl occurs in a localized
way, whereas the presence of any of the two pigment extracts
inhibits the establishment of this local activity.

References
[1] J.E.O. Mayne, P. Ridgway, Br. Corros. J. 3 (1974) 177.
[2] E. McCafferty, M.K. Bernett, J.S. Murday, Corros. Sci. 28 (1988)
559.
[3] W. Meisel, E. Mohs, H.-J. Guttman, P. Gutlich, Corros. Sci. 23
(1983) 465.
[4] Z. Szklarska-Smialowska, R.W. Staehle, J. Electrochem. Soc. 121
(1974) 1146.
[5] A.I. Onuchukwu, J.A. Lori, Corros. Sci. 24 (1984) 833.
[6] S. Virtanen, M. Buchler, Corros. Sci. 45 (2003) 1405.
[7] H.S. Isaacs, S. Virtanen, M.P. Ryan, P. Schmuki, L.J. Oblonsky,
Electrochim. Acta 47 (2002) 3127.
[8] W.J. Clark, J.D. Ramsey, R.L. McCreery, G.S. Frankel, J. Electrochem. Soc. 149 (2002) B179.
[9] C. Gabrielli, M. Keddam, F. Minouflet-Maurent, K. Ogle, H. Perrot,
Electrochim. Acta 48 (2003) 935, 1483.
[10] M. Svoboda, J. Mleziva, Prog. Org. Coat. 12 (1984) 251.
[11] M. Kendig, R. Buchheit, Corrosion Inhibition of Al and Al Alloys by
Hexavalent Cr Compounds a mechanistic overview, in: Proceedings
of Corrosion/2000Surface Conversions of Aluminum and Ferrous
Alloys for Corrosion Resistance, NACE, 2000.
[12] J. Barraclough, J.B. Harrison, J. Oil Col. Chem. Ass. 48 (1965) 341.
[13] J.E.O. Mayne, Br. Corros. J. 5 (1970) 106.
[14] M. Beiro, A. Collazo, M. Izquierdo, X.R. Novoa, C. Perez, Prog.
Org. Coat. 46 (2003) 97.

350

A.C. Bastos et al. / Progress in Organic Coatings 52 (2005) 339350

[15] K. Aramaki, Corros. Sci. 43 (2001) 591.


[16] I.M. Zin, S.B. Lyon, V.I. Pokhmurskii, Corros. Sci. 45 (2003) 777.
[17] M.J. Austin, Inorganic anticorrosive pigments, in: ASTM Manual
17Paint and Coating Testing Manual, ASTM, Philadelphia, 1995,
Chapter 27.
[18] J.A. Burkill, J.E.O. Mayne, J. Oil Col. Chem. Ass. 71 (1988) 273.
[19] H.S. Isaacs, G. Kissel, J. Electrochem. Soc. 119 (1972) 1628.
[20] H.S. Isaacs, Y. Ishikawa, Applications of the vibrating probe to localized current measurements, in: R. Baboian (Ed.), Electrochemical
Techniques for Corrosion Engineering, NACE, Houston, 1986.
[21] K. Ogle, V. Baudu, L. Garrigues, X. Philippe, J. Electrochem. Soc.
147 (2000) 3654.
[22] F. Zou, C. Barreau, R. Hellouin, D. Quantin, D. Thierry, Mat. Sci.
Forum 289292 (1998) 83.
[23] J. He, V.J. Gelling, D.E. Tallman, G.P. Bierwagen, J. Electrochem.
Soc. 147 (2000) 3661.
[24] S.M. Powell, H.N. McMurray, D.A. Worsley, Corrosion 55 (1999)
1040.

[25] I.M. Zin, R.L. Howard, S.J. Badger, J.D. Scantlebury, S.B. Lyon,
Prog. Org. Coat. 33 (1998) 203.
[26] A. Amirudin, C. Barreau, R. Hellouin, D. Thierry, Prog. Org. Coat.
25 (1995) 339.
[27] J. Titz, G.H. Wagner, H. Spahn, M. Herbert, K. Juttner, W.J. Lorenz,
Corrosion 46 (1990) 221.
[28] R.L. Zeller III, R.F. Savinell, Corr. Sci. 26 (1986) 389.
[29] A.S. Castela, A.M. Simoes, Corr. Sci. 45 (2003) 1631.
[30] J.G.N. Thomas, The mechanism of corrosion prevention by inhibitors, in: L.L. Shreir, R.A. Jarman, G.T. Burstein (Eds.), Corrosion, vol. 2, Butterworth-Heinemann, Oxford, 1994, Chapter
17.3.
[31] H. Leidheiser Jr., W. Wang, Rate controlling steps in the cathodic delamination of 1040 m thick polybutadiene and epoxypolyamide coatings from metallic substrates, in: H. Leidheiser Jr.
(Ed.), Corrosion Control by Organic Coatings, NACE, Houston,
1981.
[32] W. Furbeth, M. Stratmann, Corros. Sci. 43 (2001) 207.

También podría gustarte