You are on page 1of 12

Progress in 0rganic Coatings, 23 (1994) 275-286 275

The study of coal tar epoxy protective coatings by


impedance spectroscopy

R. BabiC and M. MetikoS-HukoviC


Institute of Electrochemistry, Faculty of Chemical Engineering and Technology,
University of Zagreb, Savska 16, 41000 Zagreb [Croatia)

H. RadovEiC
Brodarski Institti, Marine Research and Naval Technologies, Zagreb [Croatia)

(Received February 1, 1993; accepted May 11, 1993)

Abstract
Steel panels coated wlth a coal tar epoxy composition were studied by impedance
spectroscopy in the dry condition as well as during exposure to synthetic seawater and
to 10% NaOH solution. In the temperature range from 0 to 20 “C the dry coating behaves
as an almost ideal capacitor with a dielectric constant ranging from 3.0 to 3.2. The coal
tar epoxy coating immersed in synthetic seawater showed capacitative behaviour over
a long period of exposure, in the frequency range 100 Hz to 100 kHz. However, the
thinner coatings ( - 50 pm) were subject to a permanent slow degradation and their
resistance decreased to approximately 107Rcm2 after 7 months of continuous immersion,
whereas the degradation process for thicker coatings (- 125 pm) was slower and the
resistance was still above lo8 Rem’ after 9 months of continuous immersion. The water
uptake showed an initial rapid rise to between 9 and 13%. The coating degradation in
a 10% NaOH solution is rather more rapid and after only one day of immersion a coating
resistance of lo6 Szcm’ and a water uptake of 30% was found.

Key wards: Impedance spectroscopy; Nondestructive testing; Paint film resistance; Coal
tar epoxy coating; Synthetic seawater

1. Introduction

Steel products are used in almost all industrial sectors of the economy.
Virtually all such products are coated with organic paints for aesthetic reasons
and corrosion protection. Since metallic corrosion is an electrochemical
phenomenon, the application of electrochemical methods is clearly appropriate
in developing procedures for assessing paints for their corrosion protection
characteristics. During the last decade or so, considerable success has been
achieved by the use of electrochemical impedance spectroscopy [ l-8 1, which
has the virtue of being a small signal perturbation technique.
Marine corrosion is an area of particular importance and in the last few
years a large number of papers have appeared dealing with the performance
of polymer-coated steel in chloride media [ 9-19 1. The paints studied have
included coal tar epoxy compositions [20-261 which provide excellent coat-
ings, the quality of which is attributed to the fact that they combine the
best qualities of coal tar (water resistance, brine resistance, chemical resistance

0300-9440/94/$07.000 1994 Elsevier Sequoia. All rights reserved


SSDI 0300-9440(93)00442-P
276

etc.) with the adhesion properties of epoxy resins [ 201. The protective
properties have been studied as a coating on a metal surface [20-241, or
as a detached paint film [25]. The physicochemical characteristics of the
coal tar epoxy film, especially permeability to water, oxygen, chloride and
hydroxyl ions, have also been investigated [20, 261.
It is generally agreed that the coated metal system may be described
in terms of the equivalent circuits presented in Fig. 1 [ 2 7-291, where C,
and R, represent the capacitance and the resistance of the coating respectively,
and R,,the electrolyte resistance. The term 2, describes the general impedance
of the metal/electrolyte interface, the magnitude of which depends on the
rate controlling step of any electrochemical reaction, that is, diffusion and/
or charge transfer.
A perfect protective coating has no free ions, oxygen or water present
in it, and shows zero mobility of ions and zero diffusion coefficients of
oxygen and water. Such a protective coating would prevent both atmospheric
or electrochemical corrosion of the underlying metal. The impedance of such
a coated metal when in contact with a corrosive solution would be entirely
capacitive, and would vary inversely to the frequency, with the capacity, C,,
given by

C, = e. EA/d (1)
where A is the surface area of the coated metal, d is the coating thickness,
co is the absolute permittivity (8.85 X lo-14),and E is the dielectric constant
of the film.
Real protective coatings generally have a finite ionic conductivity which
is due either to defects in the coating which contain an aqueous solution,
or to true ionic conductivity within the bulk of the coating.
When the coating is immersed in a chloride solution, the latent dis-
continuities or defects eventually give rise to an ionically conducting low
resistivity path perpendicular to the coated surface which may eventually
penetrate to the metallic substrate. Assuming a direct concentration depen-
dence of the degradation mechanism [30], the resistivity of these paths will
decrease as a function of the exposure time as water, oxygen, and anionic
and cationic species are transported through the coating. Figure 1(b) suggests
that the initial impedance properties of the coating and any changes during
exposure to electrolyte can be determined at relatively high frequencies,

RC RC
(4 @>
Fig. 1. Equivalent circuits for the impedance of coated metal electrodes. (a): Dry organic
coating as an ideal dielectric layer; (b): Electrode with defective organic layer.
277

while the corrosion reaction on the metal surface can be evaluated at relatively
low frequencies [lo, 311.
The principal objective of the present work was the study of the protective
properties of coal tar epoxy coating on steel in seawater. As painted structures
in seawater are very often cathodically protected, the influence of increased
alkalinity, caused by the electroreduction of oxygen on coating the properties
is of great importance. The study of the coal tar epoxy coatings was therefore
extended to include immersion in an alkaline solution. For a better under-
standing of the initial coating properties, a dry coating was also investigated.

2. Experimental

The investigations were performed using the following systems: carbon


steel/coal tar epoxy coating/brass; carbon steel/coal tar epoxy coating/synthetic
seawater (prepared according to DIN 50900); and carbon steel/coal tar epoxy
coating/lO% NaOH solution.
The surface of the steel samples was abraded with emery paper to a
400 metallograpbic finish, and before applying the paint it was degreased
with ethanol. The coal tar epoxy paint was applied with a brush on one side
of the steel sheet. The other side was masked by fixing the sample into a
plexiglass frame using Araldite. The working area of the electrodes in the
carbon steel/coal tar epoxy coating/brass system was circular, having a surface
area of 1 cm2. The investigation was performed on five samples with an
average coating thickness of 36, 41, 42, 43, and 44 pm. A sheet of the
same shape and size made of brass served in these measurements as the
counter electrode.
The working area of the samples which were investigated in the electrolyte
solutions was 10 cm2, and was rectangular. Six samples with an average
coating thickness of 49, 49, 51, 119, 125 and 129 pm were investigated
in synthetic seawater. The measurements in the 10% NaOH solution were
performed on samples with an average coating thickness of 63, 64 and
66 pm. A platinum sheet (10 X 5 cm) was used as the counter electrode.
The film thickness of the thinner coatings was obtained with one application
whereas for thicker coatings three successive applications were used, with
a drying period of 24 hours between each layer. Painted samples were kept
seven days in a desiccator at room temperature before immersion.
Measurements of impedance I2 I and of the phase angle 0 were performed
on a Hewlett-Packard 4800 A Vector Impedance Meter. The selected fre-
quencies were chosen using a Hewlett-Packard 53000A instrument. With thii
equipment measurement of 121 values above lo7 fi was uncertain.

3. Results and discussion

3.1. The system steel/coal tar epoxy coating/brass


The investigation of the dielectric properties of the coal tar epoxy coating
as a function of temperature was performed on samples with an average
278

thickness from 36 to 44 pm. Typical experimental data for IZI and 0 for
a 44 pm coating in a temperature range from 0 to 20 “C are represented
in Bode plots in Fig. 2. It can be seen that the coal epoxy coating behaves
in a purely capacitive manner with the log 121 versus log f slope ranging
from - 1 at 0 “C to -0.97 at 20 “C, and 0 between - 83 and - 89 degrees.
Thus, a coal tar epoxy coating can be considered as an almost ideal capacitor
and can be represented by the equivalent circuit shown in Fig. l(a). The
dielectric constant was calculated from the capacitance values at 35000 Hz:
at 20 “C its value is 3.2 and with decreasing temperature it decreases slightly
to 3.0 at 0 “C.

3.2. The system steel/coal tar epoxy coating/synthetic seater


The investigation of the protective ability of the coating in synthetic
seawater was performed on samples with a coating thickness between 49
and 50 pm during seven months of continuous immersion, and on samples
with coating thickness between 119 and 129 pm during nine months of
continuous immersion. At the beginning, the values of 121 and 0 were
measured each day and subsequently on every eighth day.
Typical measurements of IZI and 0 values for a coating thickness of
49 pm are represented in Fig. 3, and for a coating thickness of 125 pm in
Fig. 4, as Bode plots. It can be seen that the coal tar epoxy coating retains
almost purely capacitive behaviour at all frequencies investigated during the
whole of the time of immersion. For thinner coatings, the log IZI versus

Fig. 2. Bode plots for a 44 pm coal tar epoxy coating in steel/coating/brass system as a
function of temperature between 0 a.nd 20 “C.
279

.
?
8 $

I
i

*
-l
. - 2 hours
+ - 24 hours
x -3Odays
no- 3months

4.oi 2.0
0 m - 5 months
. - 7months

3.0 4.0 log(f/Hz) 5.0


Fig. 3. Bode plots for a 49 pm coal tar epoxy coating on a steel substrate as a function of
exposure time to synthetic seawater.

log f slope ranges from - 0.90 after 2 hours to -0.86 after 3 months of
continuous immersion, while the slope for thicker coatings ranges from - 0.96
after 2 hours to -0.90 after 9 months continuous immersion. However,
after a few months’ continuous immersion the thinner coatings show a
response departing from the log 121 versus log f linear dependence and
below lo3 Hz the phase angle, 0, drops below - 70”.
Since the system under investigation can be described in terms of the
equivalent circuit presented in Fig. 1, it is supposed that the values of C,
and R,, at any immersion time, represent the protective ability of the coating
at that time. The value of C, can be determined by extrapolating the impedance
data to infinite frequency:

c, = l/d” (2)
for o+ a,, and also from the value of the frequency corresponding to the
top of the high frequency semicircle in the Nyquist plot, (when this is well
defined).
The change in capacitance of a coating on exposures to a corrosive
environment is a widely used parameter for evaluating its protective properties
because of the good reproducibility of the measurements and that it depends
exclusively on deterioration on a microscopic scale. It does not, however,
give a direct indication of the corrosion rate of the underlying metal since
280

50
. - 2 hours
0 - 30 days
l - 9 months
t
I 1 I L
2.0 3.0 4.0 log( f/Hz) 5.fI
Fig. 4. Bode plots for a 125 pm coal tar epoxy coating on a steel substrateas a functionof
exposure time to synthetic seawater.

it is also affected by the amount of water absorbed in the coating. An increase


in the water content generally indicates loss of the protective properties of
the paint, for example reduced adhesion/cohesion and reduced electrical
resistance [32 1. According to Brasher and Nurse [33] the primary factor
determining the quality of marine painted steel was the amount of water
absorbed on exposure.
With the uptake of water the dielectric constant E of the coating increases;
this increase is quite sensitive to the uptake of water since the dielectric
constant of the water phase is more than 20 times that for the coating.
Assuming a linear relationship between the dielectric constant of the
polymer-water system and those of the pure components, a random distribution
of the water, and linear relationship between dielectric constant and capac-
itance, Brasher and Kingsbury [34] derived the formula:

% water uptake = [ 100 log (C,(t)/C,(O)]/log 80 (3)


where Cc(t) is the coating capacitance in time t, C,(O) the coating capacitance
at t = 0 (dry coating capacitance) and 80 corresponds to the dielectric constant
of water.
In many cases, Brasher’s equation gives better agreement with gravimetry
values than do the mixture equations derived by various authors [ 321 and
it is widely used for water absorption calculations. Poor agreement was
observed at high water contents [ 331.
281

The variation of percentage of water uptake as a function of immersion


time for thinner and thicker coal tar epoxy coatings in synthetic seawater
is represented in Fig. 5. The value of C,(O) was calculated using E= 3.2
which was determined for a dry coal tar epoxy coating at frequency 35 kHz.
The value of C,(t) was also calculated from IZI and 0 values at 35 kHz.
The calculated percent water uptake after an initial rapid rise is between 9
and 13% with only a minor dependence upon coating thickness. A similar
observation concerning the influence of coating thickness on water uptake
was noticed by Kendig and Leidheiser [35]. Initial rapid water uptake has
already been reported in the literature [ 12,191. For many coatings investigated,
Walter [ 12 ] found that the paint film/solution equilibrium had been attained
rapidly, almost 50% complete within a few miuutes of immersion.
An initial increase of capacity for thicker coal tar epoxy coatings in
synthetic seawater, followed by a long period of almost constant capacitance,
accompanied by a high R, value, was observed by Scantlebury and Sussex
[22]. However, the water uptake recorded seems rather large, bearing in
mind the good protective properties of the system, as is evident from Figs.
3 and 4. Sekine [20] showed that the water absorption depends on the ratio
of coal tar to epoxy and an increased water uptake in coal tar epoxy coatings
immersed in 3% NaCl solution was attributed by Scantlebury and Sussex
[ 211 to inadequate curing. Capacity increase during immersion, by a factor
between 1.5 and 3 (corresponding to a water uptake of 9 and 25%) have
been found by many authors [ 12, 14, 151, although a significant change in
log 121 versus logf dependence was not observed, indicating relatively good
coating performance properties. According to Walter [ 111, the relatively
steady values of paint film capacitance with time suggest that paint film

Fig. 5. Percentage water uptake for coal tar epoxy coating on a steel substrate: (a) in synthetic
seawater (Cl, n thin coatings, 0, 0 thick coatings); (b) in 10% NaOH solution.
282

capacitance is determined solely by the characteristics of the paint film and


not by the state of degradation of the coated metal. Consequently, it is now
widely believed that coating resistance is the parameter which is the best
measure of the deterioration of a coated-metal [4, 11 I. Bacon et al. [361
found that the electrical resistance of an immersed organic coating on a
metal substrate is a reliable and rapid measure of coating life. According to
Bacon et al. [36], Mayne and Mills [37] and Leidheiser [38] the systems
retain their corrosion protection while the coating resistance remains high
(lo8 to 10’ flcm’), but fail when the resistance is low (below lo7 Rcm2);
poor coatings were associated with measurements of lo6 Rem’.
The suggestion that the resistance of the polymer film was equal to the
pore resistance, R,,, was introduced by Mansfeld et al. (291 who suggested
that the observed decrease of R,, with exposure time is due to damage of
the coating and the formation of conductive paths.
In a Bode plot, according to the model in Fig. 1, (R, - R,J is equal to
121 when the phase angle, 0 drops to a minimum observed at an intermediate
but lower frequency than the near 90” phase shift associated with the coating
capacitance. From Bode plots presented in Pigs. 3 and 4, the value of R,
cannot be determined accurately since the frequency range used is not low
enough. However, it is evident that the& values of thicker coatings significantly
exceed lo8 s1cm2, even after nine months of continuous immersion, while
the R, values for thinner coatings can be evaluated around lo7 &m2 after
seven months of continuous immersion.
According to Haruyama et al. [39] a useful parameter which can be
obtained from the impedance spectrum is the breakpoint frequency. It is
defined on the Bode Magnitude plot as a point where, following the curve
from higher to lower frequencies, a transition occurs from a capacitive region
of slope - 1 to a resistive region of slope 0. On the Bode Phase plot it is
a point where, following the curve from higher to lower frequencies, the
phase shift drops below 45 degrees. This frequency, designated& following
the notation of Haruyama et al. [39] is a direct function of the proportion
of defect area. For defect areas of less than 1% of the sample area there
is a direct proportionality between breakpoint frequency and the ratio of
defect to total sample area. Hack and Scully [lo] recently calculated the
effect of defect area percentage on impedance magnitude and phase magnitude
behaviour of a coated steel equivalent circuit presented as in Fig. 1 (2,
being a parallel circuit of double layer capacitance and charge transfer
resistance), assuming that the total cell area is 10 cm2 and the electrolyte
is synthetic seawater. According to calculation the breakpoint frequency of
samples with a defect area percentage of 0.001 occurs at a frequency above
lo3 Hz (approx. 3 kHz); the corresponding coating resistance is above lo7
ficm2.
Since the breakpoint frequency of our samples with a coating thickness
close to 50 pm is less than lo2 Hz, even after seven months of continuous
immersion, this indicates that the coating resistance is above lo7 Rcm2,
which points to good protective properties.
283

However, the results presented in Pigs. 3 and 4 show that the coating
thickness is an important parameter as far as protective ability is concerned.
Very thin coal tar epoxy coatings (below 40 pm) in seawater, according to
Rowlands and Schuter [ 23 1 showed high capacitance and low resistance after
only a few days immersion, whilst coating layers 200-300 pm thick kept a
high resistance even after one year’s exposure. This can be explained by
the inhomogeneous nature of polymer films and its effect on ionic resistance.
A number of polymer films have been shown to be heterogenous in structure,
containing small areas, termed direct or D-type areas, which have a lower
ionic resistance than the rest of the film called inverse or I-type areas [37].
Differences between D- and I-type areas have been connected with differences
in crosslink density within the polymer film. Breakdown of painted metal
has been shown to occur at D-type areas of low resistance, where polymer
crosslinking density is presumably lower, leading to corrosion of the substrate.
The effect of high film thickness is to reduce dramatically the number of
D-type areas. D-areas are small and the chance of one D-area overlapping
another D-area is low; for two coat films it is almost nil [ 401. Thus, multicoat
films are more effective in providing good corrosion resistance than a single
film of the same equivalent thickness.

3.3. The system steel/coal tar epoxy coating/IO% NaOH


The investigation of the protective properties of the coating in a 10%
NaOH solution was performed on samples with a coating thickness of 63
to 66 pm, during a period of 10 days of continuous immersion.
Typical measurements of I2 I and 0 values for a coating thickness of
64 pm are represented in Pig. 6. In the frequency range investigated, purely
capacitive behaviour is only evident during the initial immersion time. The
log 121 versus logf slope is - 0.98 immediately after immersion and -0.95
one hour later. After a few hours of immersion the coating response departs
from capacitative behaviour below lo4 Hz where the resistive behaviour
becomes dominant. From Pig. 6 it can also be seen that the log 121 versus
log f curves in a frequency range showing resistive behaviour, gave lower
I.27 values during the first few hours of immersion and then, after a slight
increase, a stationary state was reached. The corresponding pore resistance
values determined by a graphical method [ 411 from the Nyquist plots (the
Nyquist plot after 24 hours immersion is presented as an example in Fig.
7) are represented in Pig. 7, and show a minimum value on an R, versus
time curve. Similar changes of pore resistance with time are well documented
in the literature. An increase of pore resistance with time for a poly(butadiene)
coating on steel in 1% NaOH was observed by McIntyre and Leidbeiser [42]
and was attributed to the blockage of pores or defects by corrosion products.
Perbe et al. [ 191 recently found a minimum on a R, versus time curve for
steel samples coated with a glycerolphthalate paint and immersed in a 3%
NaCl solution. They assumed that two phenomena occurred simultaneously,
but with variable rates, depending on the immersion time: (1) the increase
284

-90
0 0
.

.
-70
E
A - .
&I
. + al
‘5 .
A
. 1 $
.
-30
.

- 10

D
. 0
. A
. .
0

f
.

0 - 0 hour
l - 1 hour
l - 24 hours

A - 10 days
I I I t

2.0 3.0 4.0 log(f/Hz) 5.0


Fig. 6. Bode plots for a 64 pm coal tar epoxy coating on a steel substrate as a function of
exposure time to 10% NaOH solution.

in pore density, and (2) the partial sealing of the pores by corrosion products
or paint aggregates. The first phenomenon prevails for short immersion time,
whereas blockage of the pores, that is the second phenomenon, becomes
more prominent over a long immersion time, as can be seen in Fig. 7.
A value of pore resistance of between 1O6and 1O7ficm2 is often supposed
to represent the boundary between good and poor coating protective properties,
although there is evidence that corrosion beneath an organic coating can
occur at an appreciable rate at R, equal to lo6 Rcm2 [ 381.
The change in water absorption during immersion in NaOH solution,
calculated in the same way as for synthetic seawater is represented in Rig.
5. The Brasher and Kingsbury equation for calculating water uptake may
not be appropriate at these high levels of absorption because it assumes
random uptake uniformly over the whole surface which is more likely at the
lower levels. A rapid initial water uptake can again be observed followed by
a stationary state, which is in agreement with the change in pore resistance.
285

I I 1 I 1 -I
2 4 6 8 t/days 10

Fig. 7. Variation of the coating resistance with immersion time in 10% NaOH solution for a
64 @rn coal tar epoxy coating on a steel substrate. Nyquist plot for the same coating after
immersion time of 24 hours.

4. Conclusions

Steel samples coated with coal tar epoxy compositions were studied by
impedance spectroscopy (100 Hz to 100 kHz) in a dry condition as well as
during exposure to synthetic seawater and to 10% NaOH solution. From
these investigations the following conclusions can be drawn:
(1) the dry coating behaves as an almost ideal capacitor in the temperature
range from 0 to 20 “C. In this temperature range its dielectric constant varies
from 3.0 to 3.2;
(2) the coal tar epoxy coating immersed in synthetic seawater shows
a purely capacitative response over a long initial period of exposure. However,
after a few months immersion thinner coatings ( - 50 pm) depart from purely
capacitative behaviour below 1 kHz;
(3) the calculated percent water uptake after an initial rapid rise is
between 9 and 13%, with minor dependence upon the coating thickness.
The thiier coatings showed a slow permanent degradation, but their electrical
resistance after 7 months of continuous immersion was still above 1O7acrn’.
The thicker coatings ( - 125 pm) showed coating resistance above lo8 Rem’
after 9 months of continuous immersion;
(4) in 10% NaOH solution the degradation of the coal tar epoxy coating
is a more rapid process. The calculated water uptake after the initial rapid
rise on the first day stayed at around 30% during the 10 days of exposure.
The coating resistance versus immersion time showed a minimum (1 O6ncm2)
after approximately one day’s immersion followed by a slight increase up
to 1.3 X lo6 &m2 and no further change;
(5) the impedance spectroscopy parameters were shown to be useful
in the study of the coating degradation process.
286

References

1 M. Kendlg and J. Scully, Cowosio?k, 46 (1990) 22.


2 F. Mansfeld, Electrochim. Acta, 35 (1990) 1533.
3 G.W. Walter, Corros. Sci., 30 (1990) 617.
4 J.R. Scully, J. Electrochem. Sot., 136 (1989) 979.
5 U. Rammelt, G. Reinhard and K. Rammelt, .Z. Ebctroanal. Cti., 180 (1984) 327.
6 G.W. Walter, J. Electroanal. Chern., I18 (1981) 259.
7 M. Piens and R. Verb&, in H. Leldheiser, Jr. (ed.), Corrosion Control by Organzc Coatings,
NACE, Houston, 1981, p. 32.
8 Y. Sato, Prog. Org. Coat., 9 (1981) 85.
9 E. Frechette, C. Compere and E. Ghali, Corros. Scz, 33 (1992) 1067.
10 H.P. Hack and J.R. Scully, J. Electrochem Sot., 138 (1991) 33.
11 G.W. Walter, Corros. Sci., 32 (1991) 1059.
12 G.W. Walter, Corros. Sci., 32 (1991) 1041.
13 J.N. Murray and H.P. Hack, Corrosion, 47 (1991) 480.
14 C.T. Chen and B.S. Skerry, Corr-osion, 47 (1991) 598.
15 A. Amirudm and D. Tmerry, Br. Curt-OS. J., 26 (1991) 195.
16 J. Titz, G.H. Wagner, H. Spahn, M. Ebert, K. Juttner and W.J. Lorenz, Corrosion, 46 (1990)
221.
17 D.M. Drazic and V.B. Miskovic-Stankovic, Corros. Sci., 30 (1990) 575.
18 S. Feliu, J. Galvan and M. Morcillo, Corros. Scz., 30 (1990) 989.
19 N. Pebere, Th. Picaud, M. Duprat and F. Dabosi, Cowos. Sci., 29 (1989) 1073.
20 1. Sekine, in H. Leidheiser, Jr. (ed.), Corrosion Control by Organic Coatings, NACE,
Houston, 1981, p. 130.
21 J.D. Scantlebury and G.A.M. Sussex, in H. Leidheiser, Jr. (ed.), Proc. Ccrrrosiun Control
by Organic Coatzngs, NACE, Houston, 1981, p. 51.
22 G.A.M. Sussex and J.D. Scantlebury, 8th Znt. Congr. Metullic Corrosion, Mainz, Germuny,
1981, p. 1074.
23 J.C. Rowlands and D.J. Chuter, Con-OS. Sci., 23 (1983) 331.
24 J. Jargstorf and U. Rammlet, Plaste. Kautsch., 35 (1988) 31.
25 G.A.M. Sussex and J.D. Scantlebury, J. Oil Colour Che?m. Assoc., 66 (1983) 142.
26 H. Haagen and W. Funke, J. Oil Colour Chem. Assoc., 58 (1975) 359.
27 L. Beaunier, I. Epelboin, J.C. Lestrade and H. Takenouti, Surf: Technol., 4 (1976) 137.
28 J.D. Scantlebury and K.N. No, J. Oil Coluur Chem Assoc., 62 (1979) 89.
29 F. Mansfeld, M.W. Kendig and S. Tsai, Corrosion, 38 (1982) 478.
30 H. Leidbeiser, Jr., D.J. Mills and W. Bilder, m M.W. Kendig and H. Leidheiser, Jr. (eds.),
Corrosion. Protection by Organic Coatings, The Electrochemical Society Softbound Pro-
ceedings Series, Pennington, NJ, 1987, p. 31.
31 F. Mansfeld, S.L. Janjaquet and M.W. Kendig, Corros. Sci., 26 (1986) 735.
32 S.A. Lindquist, Corrosion, 41 (1985) 69.
33 D.M. Brasher and J.T. Nurse, J. Appl. Chem., 9 (1959) 96.
34 D.M. Brasher and A.H. Kingsbury, .Z. Appl. Cherry., 4 (1954) 2.
35 M.W. Kendig and H. Leidheiser, J. Ebctrochem. Sot., 123 (1976) 982.
36 R.C. Bakon, J.J. Smith and F.M. Rugg, Znd. Chem., 40 (1948) 161
37 J.E.O. Mayne and D.J. Mills, J. Oil Colour Chem Assoc., 58 (1975) 155.
38 H. Leidheiser, Jr., Prog. Org. Coat., 7 (1979) 79.
39 S. Haruyama, M. Asari and T. Tsuru, in M.W. Kendig and H. Leidheiser, Jr. (eds.), Corrosion
Protection by Organic Coatings, The Electrochemical Society Softbound Proceedings
Series, Pennington, NJ, 1987, p. 197.
40 D.J. Mills and J.E.O. Mayne, in H. Leidheiser, Jr. (ed.), Corrosion Control & Organic
Coatings, NACE, Houston 1981, p. 12.
41 L. Lemaitre, M. Moors and A.P. Van Peteghem, J. Appl. Ekctrochem., 13 (1983) 803.
42 J.F. McIntyre and H. Leidheiser, J. Electrochem. Sot., I33 (1986) 43.

You might also like