You are on page 1of 27

JCSE Volume 3 Paper 5 Page 1 of 27

ISSN 1466-8858

Google Search

Web www.jcse.org

American Public Univ.


99% of Employers Said They Would Hire One of Our Graduates Again.
www.apus.edu

Volume 3 Paper 5
Submitted 9th August 2000

The Protective Action of Organic Coatings on Steel: A Review


David Greenfield and David Scantlebury
Corrosion and Protection Centre, UMIST, PO Box 88 Manchester M60 1QD, UK
Email: scantlebury@manchester.ac.uk

Abstract

This review considers the past fifty one years of study in the field of corrosion prevention by organic
coatings. It concentrates mainly on the mechanisms associated with the paint binder. Areas covered
include the importance of the coating ionic resistance and the concept of D and I areas. The effect of the
environment on the ionic resistance is examined. Blistering and delamination is considered in the light of
movement of ions through coatings and along interfaces. Specific attention is paid to the metal paint
interface with emphasis on adhesion, underfilm contamination and surface tolerant coatings.

Keywords: anti-corrosion paints, ionic transport, blistering, cathodic disbonding, underfilm


contamination, surface tolerant coatings

Introduction

In this review, we are concerned with the effects of the polymeric binder and the anti-corrosion properties
of paints. Aspects of corrosion inhibition by anti-corrosion pigments have not been emphasized. Also
within the field of underfilm corrosion and blistering, we have not considered the the highly important
and topical area of filiform corrosion.

Inhibition of Corrosion by Barrier Coatings

Any discussion which considers the prevention of aqueous metallic corrosion, usually takes as its starting
point the electrochemical model of the corrosion of mild steel in a neutral electrolyte with the four
important processes involved linked in series namely; the anodic reaction, the cathodic reaction and the
conductive pathways for ions and electrons. Inhibition of the anodic or cathodic reactions or an

http://www.jcse.org/volume3/paper5/v3p5.php?commentmode=2 6/20/2013
JCSE Volume 3 Paper 5 Page 2 of 27

interruption of the ionic flow will cause the rate of corrosion to be significantly reduced.[1] Thus, all
corrosion prevention measures are aimed at removing or suppressing one of the three elements. One of
the most convenient, and certainly the oldest method of protecting a substrate from the detrimental effects
of the environment, is to coat it with a polymeric barrier paint to isolate it from its surroundings.[2],[3]
This barrier might be due purely to the properties of the polymer, or to the inclusion of inert pigments
that act to increase the length of the diffusion path through the coating.

The inclusion of inert pigments into the film can increase the effectiveness of the coating providing the
formulation is not too heavily loaded. Too high a concentration of pigment is liable to reduce the
effectiveness of the coating. The Critical Pigment Volume Concentration has been shown to be a
transition point where a number of significant features of the coating change.[4],[5] The loading of a
coating with pigment particles is made more complex by the effect of the distribution of pigment particle
size.[6] A wide range of pigment particle sizes favours a high Pigment Volume Concentration. David –
is PVC the same as CPVC – define it!

The nature of this barrier has been the subject of much work over many years by workers in the field of
protective organic coatings. The original assumption was that organic coatings act as a barrier to water
and oxygen from the environment[7]. Scientific investigations of the subject over the last half century
have determined that the limiting factor in the protective mechanisms of barrier coatings is frequently
their resistance to the flow of ionic current.[8]

Electrolytic Resistance of Organic Coatings

Figure 1 : Schematic of resistance behaviour of


coatings in immersed metals after ref [9]

The importance of the electrolytic resistance of organic


coatings has long been appreciated. Bacon, Smith and
Rugg,[9] after measuring the resistances of over 300
coating systems, determined a direct correlation between
these resistances and the ability of the coating to protect
the underlying steel from corrosion. Three general
classifications, based on a sustained resistance value,
were established during this investigation: good, fair and
poor - see Figure 1. The classification that a coating was
allocated was dependant upon its long term resistance
value upon immersion in solution “Most coatings, if
continuous, have resistances in the neighbourhood of
Log R = 9 during the first 5 to 10 minutes of immersion,
followed by a decrease in resistance which may vary
considerably in steepness and duration”, the units of R were Ohms for one sq. cm. Their subsequent
behaviour determined which category they should be placed into. All coatings were found to exhibit an
initial decrease in resistance, which varied in terms of rate and duration. For a good coating, this initial
decrease was followed by an abrupt recovery to around the original value. The subsequent resistance of
the coating remained either essentially constant or fluctuated within the high resistance region. A coating
designated as fair either levelled off or displayed a slight increase in resistance. However, a subsequent
sharp drop in resistance was a pre-cursor to coating failure, which occurred within six months of the
exposure period. The resistance of a poor coating continued to decrease resulting in failure within sixty
days. A coating that maintained a resistance of 108 Ohm cm2 provided good corrosion protection while
one whose resistance fell below 106 Ohm cm2 did not.

http://www.jcse.org/volume3/paper5/v3p5.php?commentmode=2 6/20/2013
JCSE Volume 3 Paper 5 Page 3 of 27

Although this work identified the relationship between the ionic resistance of the coatings and their
ability to protect, the permeability of other corrosive agents other than ionic species was not considered.
Therefore, the results were presented as providing a measure of the likely protection that could be
expected from a particular system rather than proposing a mechanism for the protective action of organic
coatings.

The Effect of the Environment on the Performance of Coatings

A further important result from the work of Bacon, Smith and Rugg identified the effect of the
environment on the protective merit of coated submerged steel. One of the variables taken into
consideration was the level of dissolved salts in the test solution. It was found that the resistance of the
coating fell with increasing salt concentration and from this result it was concluded that the dissolution of
environmental water into the coating was more important than the uptake of salt from the solution by the
coating. That the level of dissolved salts of the exposure environment had an effect on the properties of
coatings had been reported previously by Kittelberger and Elm [10]. This work showed a relationship
between the level of dissolved salts in the solution, expressed as osmotic pressure, and the water uptake
of the films under investigation, expressed as percentage weight gain. It was shown that the nature of the
solute did not affect the results; rather it was the osmotic pressure that determined the level of imbibition
is this a word? of water into the coating. When the coatings were exposed to a salt solution, after an
initial high rate of water absorption, the curve levelled off and eventually an equilibrium was attained.
Where the films were exposed to distilled water, however, no equilibrium was reached and the film
continued to gain in weight. It was concluded that the degree of water absorption was a function of the
activity of the water in solution, due to the differences in osmotic pressure between the exposure solution
and that within the film, see Figure 2.

Figure 2: The relationship between osmotic pressure and water uptake after ref [10]

Permeability of Organic Coatings to Corrosive Species

In order that a paint coating should offer protection from corrosion, it should interact with one or other of
the processes involved in the overall series of corrosion reactions. This was considered by Mayne[11],
who determined that the permeability of organic coatings to both water and oxygen is so high that the rate
at which they arrive at the interface of the coating and the steel is greater than that required for corrosion
to proceed and so could not be rate determining. Therefore, Mayne reasoned that the method by which an
unpigmented coating protects steel from corrosion was by the introduction of a path of high electrolytic
resistance in between the metal and the environment; the term resistance inhibition was coined for this

http://www.jcse.org/volume3/paper5/v3p5.php?commentmode=2 6/20/2013
JCSE Volume 3 Paper 5 Page 4 of 27

mechanism.

Selective Permeability of Organic Coatings

Further work by Mayne [12] showed that, upon immersion in water or an aqueous solution of electrolyte,
most organic coatings acquire a negative charge. According to Mayne, the acquisition of this charge has
the effect of creating a selectively permeable membrane, which is preferentially permeable to cations;
that is, a film that has gained a negative charge due to immersion may be regarded as a large poly-anion.
An alternative view is expressed by Corti et al [13], who claim that a charged film would slow down the
passage of counter ions (those having the opposite charge) whilst having no effect on the movement of
co-ions through the film.

These contradictory views stem from the proposed mode of transport of ions across the film. According
to Corti, the rate of permeation across the film is affected by the presence of small imperfections or pores,
which extend through the film and have cross sections distinctly larger than the free areas normally
present between the atomic groups in the membrane matrix. Mayne’s model, on the other hand, infers
that the passage of ions is through the bulk matrix of the film. If the ionic conduction takes place through
pores, it is reasonable to claim that a negative charge on the polymer would impede the passage of cations
due to electrostatic attraction between the ions and the pore walls. However, if the movement of ions
through the coating is considered to be through the bulk of the film, the proposal that a negative charge
on the film would act as a repulsive force towards anions seems plausible.

Transport of cations through polymer films has been observed elsewhere[14] with the passage of iron
ions through a range of coatings on steel immersed in NaCl. These coatings were continuous and free
from defects and the ionic transport was considered to have taken place through the bulk polymer. Of the
coatings tested, a “practical anticorrosive coal tar epoxy” was found to retard the process. The inference
of this work is that the conduction of ions through the film takes place through the bulk polymer. The
work reviewed in the following section proposed mechanisms to account for this.

Mechanisms of Through Film Transport of Ions


Ionic conduction through the bulk of the polymer matrix was studied by Maitland and Mayne.[15] This
work, which mainly considered the properties of an unattached, unpigmented pentaerythritol alkyd
varnish, found that the resistance of the film was affected by the state of the exposure solution and
identified two distinct processes, which took place in succession of each other: The Fast Change and The
Slow Change. The Fast Change was observed to take place within a few minutes of immersion where the
films attained a steady resistance value, Ro. This was followed by The Slow Change, which took place
over a number of weeks or months.

The Fast Change was shown to be a reversible process and was independent of both the nature of the
solute in the exposure solution and the state of the film in respect of the slow change. The controlling
factor in the process was seen to be the osmotic pressure or water activity of the solution; this process is
shown in Figure 3

http://www.jcse.org/volume3/paper5/v3p5.php?commentmode=2 6/20/2013
JCSE Volume 3 Paper 5 Page 5 of 27

Figure 3: The Fast Change; Comparison of the effect of 3.5N KCl with isotonic sucrose solution
after ref. 15

The Slow Change was considered to be controlled by the concentration of the electrolyte in solution.
This was confirmed by experiment, whereby the reaction was accelerated by an increase in temperature.

Maitland associated this phenomenon with an ion exchange process where cations from the electrolyte
were exchanged with hydrogen ions from carboxyl groups within the polymer. The ion exchange
behaviour was found to be affected by the pH of the system or the concentration of potassium ions in
solution. An increase in either the pH or the potassium concentration resulted in a greater rate of ion
exchange, causing a fall in the resistance of the coating.

Work on the effect of the Slow Change on the reversibility of the Fast Change was carried out by Cherry
and Mayne,[16],[17] who found that the coatings under investigation altered their mode of conduction at
a given value of pH, the value of which varied according to the particular coating. Working on the
assumption that the slow change was an ion exchange process, it was postulated that the resistance of the
film should be a function of the ratio of the activities of the metal cation from the electrolyte (in this case
potassium) and that of the hydrogen ions. The term rK was defined, where:

[1]

Films were conditioned in solutions of various rK values and then the temperature coefficient of
resistance for each film was measured. It was found that there was a small range of rK values, specific to
each polymer, where the mode of conduction ceased to run counter to that of the solution and
commenced to follow it. The point where this change in the conductive behaviour of the film took place
was named the reversal point and was seen to indicate a change in the mechanism of ion transport from
one of activated diffusion to that of aqueous conduction. This conduction was considered to take place
through virtual pores.

The model whereby the electrolytic conduction takes place through pores is supported by Miskovic et al
[18] who, as a result of their AC impedance examination of electrodeposited epoxy, proposed a model
whereby electrolytic conduction through a coating is dependant only on conduction through macro-

http://www.jcse.org/volume3/paper5/v3p5.php?commentmode=2 6/20/2013
JCSE Volume 3 Paper 5 Page 6 of 27

pores. These pores were examined microscopically and were determined to have radii between 3 and
7.2µm.

Modes of Ionic Conduction Through Polymer Films

The question of the possibility of conduction through pores was addressed by Kinsella.[19] This work
identified different modes of ionic conduction through polymer films. Two distinct forms of conduction
were identified. The first was I, or inverse type conduction, where the resistance of the film coincided
with that of the solution; this type of conduction was in line with that identified by Bacon, Smith and
Rugg [9].

The second mode of through-film transport that was identified was named D, or direct type conduction
where the film’s resistance ran counter to that of the solution’s. Initially, the films that exhibited D type
conduction were thought to be due to coating deficiencies such as dust wicks passing across the film.
However, investigation of the temperature coefficients of resistance for these films showed them to be
higher than that of aqueous conduction. In addition, the films were shown to be preferentially cation
permeable and thus considered to be continuous.

Figure 4: Schematic plot of I and D type


conduction, after ref [20]

A series of papers were produced as a result of


this work, which addressed various different
factors affecting ionic conduction in polymer
films. The first,[20] examined the influence of
the electrolyte on the resistance of films. It was
found that the films investigated followed the
general trend exhibited in Figure 4. However,
when an I Type film was exposed to a solution
of very low electrolyte concentration, and
therefore a high water activity, the film
displayed D Type characteristics over a small
range of concentrations, at the lower end of the
scale. Once the electrolyte concentration reached
a critical level, the film began to exhibit more
typical I Type characteristics. This led to the oft
quoted conclusion that the difference between D
and I Type conduction was “more one of degree rather than of kind”.

Scantlebury examined the distribution of D and I Type areas across a given film.[21] It was found that
the films investigated were all very inhomogeneous and had a distribution of both D and I Type areas
over the film. Two conclusions, relating to the mechanisms responsible for the different modes of
conduction were drawn from this investigation. Firstly, that D Type conduction could not be attributed to
the presence of pores unless they were of molecular dimensions.

The second conclusion that was drawn from this study, was that the resistance of both D and I Type areas
bore a relationship to their hardness values, as determined via microhardness tests. D Type areas were
found overall to have a lower hardness value than that obtained for I Type. The hardness of the film was
equated to its crosslinking density. This result was reinforced by swelling experiments, which showed
that D Type areas swelled to a greater degree than I Type. Therefore, it was concluded that D Type
conduction was a function of lower crosslink density in the film.

http://www.jcse.org/volume3/paper5/v3p5.php?commentmode=2 6/20/2013
JCSE Volume 3 Paper 5 Page 7 of 27

This conclusion reinforced Kinsella’s view [20] that the initial drying conditions had a marked effect on
the final properties of the film. Other work on crosslink densities of polymer films has shown that the
crosslink density appears only to affect the ionic conductivity of the film.[22] The study was concerned
with the effect of crosslink density upon water permeation through films and concluded that the
permeation of water was not noticeably affected by alteration of the crosslink density. The final two
papers in the series by Mayne and co-workers considered the influence of temperature[23] and finally the
effect on the barrier properties of a coating due to the inclusion of inert pigments.[24]

Evidence that D Type areas are detrimental to the protective properties of organic coatings was produced
by Mills, [25] who showed a direct relationship between the presence of D Type areas in the film and the
occurrence of underfilm corrosion. With regard to the term “virtual pores” put forward by Cherry17 to
explain the mode of conduction in an apparently continuous film, Mills suggests the term “conductive
polymer phase” as more appropriate.[26] This term is intended to describe more accurately the apparent
properties of D Type films.

Mills further observed that the proportion of D Type areas was a function of the thickness of the film.
[27] It was found that the incidence of D Type areas decreased significantly once the film’s thickness
was increased to around 75μm, this finding was used to conclude that D Type areas were spherical and
approximately 75μm diameter.

Breakdown Mechanisms of Coatings

Blistering and Delamination

Amongst the most common forms of failure found in organic coatings are those of blistering and
delamination. Much debate has ensued as a result of the research into these failure modes and a number
of governing mechanisms have been put forward. The mechanisms involved in the two modes of failure
are similar, but it is unclear whether they are the same phenomena.

Within each of these failure modes, there are sub-classifications, which may or may not be important for
a particular system. The components of a system are the substrate, the coating and the exposure
environment. Factors affecting the performance of a system include surface preparation, coating
application, cure regime and film integrity.

Factors Affecting Blister Initiation

The blistering mechanism of paint coatings upon exposure to aqueous environments was addressed by
Gay,[28] who found that all the systems examined displayed four common features:

1. The higher the osmotic pressure of the immersion liquid, the smaller the amount of blistering.
2. The fluid in blisters formed on seawater immersion is almost invariably alkaline and its chloride
content is lower than that of the seawater.
3. The steel under such blisters is usually bright and free from corrosion.
4. Areas of blistering are frequently associated with adjacent areas of corrosion.

In the majority of cases, before any blistering occurred, some corrosion was observed at weak points in
the coating and no substantial amount of blistering was found where corrosion was not present. The
suggested sequence of events leading to the formation of blisters is as follows.

The paint film imbibes water from the solution, possibly containing some dissolved salts.
Eventually, sufficient liquid containing chloride ions passes through to the underlying metal and

http://www.jcse.org/volume3/paper5/v3p5.php?commentmode=2 6/20/2013
JCSE Volume 3 Paper 5 Page 8 of 27

activates primary corrosion sites at the interface.


As the corrosion proceeds at the anodic areas under the film, hydroxyl ions build up at the cathodic
sites.
The alkaline environment at the cathodic sites weakens or destroys the adhesion of the film whilst
producing osmotically active substances at the paint metal interface.
The presence of these substances at the interface is favourable to either osmotic or endosmotic
passage of water through the film from the immersion liquid.

The principle of endosmotic transfer of water through the coating was supported by Mayne [29] who
concluded that the contribution of osmotic transfer of water was much smaller; 6% in the case of the
coatings studied.

Figure 5: Stages in the development


of an osmotic blister.

Subsequent work on blistering has


produced three individual
classifications

Osmotic Blistering
Anodic Blistering
Cathodic Blistering

Osmotic Blistering

This form of failure is one


consequence of a contaminated
substrate.[30], [31], [32]. Soluble
salts at the interface can form a
concentrated salt solution that, due to
decreased water activity, acts to draw
water through the coating, which
behaves as a semi-permeable
membrane, from the exposure
environment. The osmotic pressures
exerted can be expected to range
between 2500 and 3000 kPa
compared with a mechanical
resistance to deformation of a coating
ranging from 6 to 40 kPa[33] In addition, the pressure within the blister becomes greater than the
atmosphere on the free side.[34] So, provided the film is not ruptured in the process, the conditions exist
for a continually expanding blister.

Figure 5 schematically represents the stages in the formation of an osmotic blister. In Figure 5A a layer
of interfacial water in contact with soluble salts is shown, in B the salts at the interface have formed a
concentrated solution and finally C represents the stage when the osmotic pressure, produced by the
concentration gradient across the coating, has drawn water through from the outside environment and
formed the blister. The foregoing is attractive due to its simplicity; however, it has been shown29 that
electroosmosis plays a predominant role in this form of blistering with osmosis only contributing to about
6%.

http://www.jcse.org/volume3/paper5/v3p5.php?commentmode=2 6/20/2013
JCSE Volume 3 Paper 5 Page 9 of 27

Gowers and Scantlebury[35] were able to measure currents flowing between different areas of the
substrate under artificial blisters. They found that contaminating the surface of the substrate with NaCl
promoted the corrosion process, the current flow and the blister formation.

One suggestion that has been made to reduce the tendency for a coating to suffer osmotic blistering
hinges on the formulation of the film. In a solvent borne coating, it is normal to find that a combination
of different solvents is employed. According to Storfer and Yuhas, hydrophilic solvents can facilitate and
accelerate several of the generally recognized blister-formation mechanisms, particularly osmotic
blistering.[36] An inverse relationship was determined between the solubility of water in the last solvent
in the formulation to evaporate (the tail solvent) and the blister resistance of the coating. Hydrophilic
solvents are more prone to cause blistering.

Anodic Blistering

This mode of failure was addressed by Koehler,[37] who considered liquid filled blisters to be anodic in
nature. Experiments were conducted an on epoxy phenolic resin exposed to 0.01N calcium chloride.
Calcium was chosen as the cation because it does not produce an alkaline environment. No deliberate
faults or holidays were introduced into the coating. The samples were anodically polarised and blisters
appeared after seven days. The substrate in the centre of the blister was dished and corroded while the
periphery was bright, indicating cathodic detachment. The explanation of these results was that during
the test, chloride ions passed through the coating at thin points and created an acid environment under the
coating within the blister. The cathodic reaction was postulated as being the reduction of the ferric oxide
on the steel to the soluble ferrous state. This type of failure requires a low pH environment to proceed.
Such low values may be found in many canned foods and drinks.

Cathodic Blistering

It is to this form of blistering that the majority of the literature is directed. Cathodic blistering is the
result of an alkaline environment under the coating caused by the cathodic reaction, associated with
corrosion that occurs at a damaged site of the film.[38], [39], [40] Early work by Kittelberger [41]
identified a relationship between bare areas of steel and the occurrence of blistering on immersed panels;
it was shown that an exposed area of of the coated specimen was enough to exert a significant
influence on the blistering of the coating. Figure 6 shows the proposed failure mechanisms, which may
occur when a film containing a fault is exposed to a corrosive medium.

The fault may take the form of mechanical damage to the coating or may be an inherent fault of the
coating i.e. pores or holidays. The primary pre-requisite for this form of failure is that the substrate
should support a cathodic reaction. In the case of a neutral or alkaline environment the cathodic reaction
would be the reduction of oxygen. Tests with cathodically polarised steel coated with polybutadiene,
exposed to a NaCl solution,[42] developed blisters containing a solution of high pH after 7 days, which
coalesced in 18 days; no intentional damage was made to the coatings in this experiment.

http://www.jcse.org/volume3/paper5/v3p5.php?commentmode=2 6/20/2013
JCSE Volume 3 Paper 5 Page 10 of 27

Figure 6: Possible consequences of a damaged coating.

As no defects were introduced into the coating material, the question of the route taken by the reactants is
raised. Mayne12 showed that the transmission of water and oxygen through the film was in excess of
that needed for corrosion and attributed the protective mechanism to the film’s high ionic resistance. If
we accept this hypothesis, some pathway must exist through the film to allow the sodium ions to the
interface in order to produce the alkaline environment. These pathways could be due to pores. However,
an alternative theory is proposed by Leidheiser,[43] who in a cautionary technical note suggests that
above a given concentration, alkali cations may have a deleterious effect on the coating, which leads to
morphological changes hence introducing conductive pathways to the interface.

Cathodic Delamination

Akin to cathodic blistering, cathodic delamination is also the result of alkalinity at the interface. Again,
this alkalinity is the result of cathodic activity under the coating. It is associated with faults, either
inherent or induced, in the coating. Cathodic polarisation may be a consequence of either corrosion at the
point of damage or the application of cathodic protection. Resulting from experiments carried out by
Smith and Dickie on primer failure,[44] it has been shown that under impressed cathodic conditions,
corrosion inhibitive pigments play no part in the reduction of disbonding. Under such conditions, the
performance of the system reflects the resistance of the primer resin system to alkali displacement.
During salt spray exposure, it was found that the corrosion inhibitive pigments did exhibit a degree of
control of the rate of failure by inhibiting the anodic reaction.

The Mechanism Of Delamination

A number of explanations have been put forward whereby the alkaline environment under the film affects

http://www.jcse.org/volume3/paper5/v3p5.php?commentmode=2 6/20/2013
JCSE Volume 3 Paper 5 Page 11 of 27

the integrity of the metal - polymer interface, or perhaps more properly the interface between the oxide
and the polymer. Koehler,[45] showed that this form of failure, which he called “halo detachment”, only
occurs when there are alkali metal cations available in the environment to act as counter ions to the
cathodically generated OH-.

One explanation is that dissolution of the air-formed oxide layer is responsible for the loss of adhesion.
Various techniques have been adopted to confirm this view. Ritter[46] carried out ellipsometric studies
on a range of coatings in tandem with an investigation of uncoated steel exposed to saturated NaOH. The
coatings studied were collodion, polystyrene and two proprietary acrylics, all of which were cured at
room temperature. Additionally, a heat-cured acrylic was examined. The argument put forward was that
delamination is a direct result of oxide dissolution and this is supported by the presence of iron ions in the
interfacial region of the polymer in the heat cured specimens. The oxide layer present on the metal under
these circumstances, is thicker than the normal air-formed layer, as a result of the curing regime.

Further work carried out using this technique[47] concluded that the high pH generated under the coating
was conducive to oxide growth and surface roughening of the substrate. This conclusion was the result
of tests carried out on uncoated substrates in a high pH environment. Taken in conjunction with the
findings cited in the previous paragraph, the inference of this could be that the thicker oxide layer,
developed due to the high pH generated under the film, could cause the oxide film to become
mechanically weaker and result in a cohesive failure.

Castle and Watts,[48] in an XPS investigation of cathodic delamination, concluded that dissolution of the
oxide layer was not a significant factor in the process. It was found that the oxide within the disbonded
region was reduced only in localised patches and the disbonded area extended well beyond this.

In a similar vein, Wiggle et al[49] considered different pre-treatments and the effect upon them of
cathodically produced hydroxide. Both zinc and iron phosphate pre-treatments along with bare metal was
included in the study. Initial tests were carried out to determine the porosity of the conversion coatings,
which yielded results of 0.2% for the zinc and 23% for the iron phosphate; these porosity values were
linked to the poor resistance of the iron phosphate to anodic undermining. Experiments were carried out
under both anodic and cathodic polarisation as well as at open circuit. Cathodic polarisation trials
resulted in dissolution of the zinc phosphate and hence loss of adhesion of the coating. Knaster and Parks
[50] also identified dissolution of zinc phosphate under paint coatings on pre-treated steel but considered
this to be a secondary reaction. The contention in this case is that delamination of coatings from
phosphated panels, due to a defect, is caused by oxygen depolarisation along the phosphate-paint
interface and hydrogen evolution at the delamination front that results in a force, which disrupts the
adhesion of the polymer to the phosphate layer.

Dissolution of phosphate conversion coatings on steel was also examined by Sommer and Leidheiser.
[51] The variables they considered were the effect of the pH and also the alkali cation involved. The pH
effects were investigated by exposing test panels to 0.01, 0.1 and 1.0-M solutions of NaOH for ten minute
intervals up to one hour. The importance of the cation involved was determined with 0.1-M solutions of
hydroxides of Na, Cs, K and Li. It was found that increasing pH values from 11.5 to 13.5 resulted in a
higher rate of dissolution of the phosphate ions from the conversion coating.

With regard to the cation type, the investigators were surprised to find that Na was the most effective of
the cations in the dissolution of the coating (1.6, 2.6, & 4.1 times K, Cs & Li respectively). Additionally,
the cation type not only affected the rate but also the mode of dissolution. With Na, Cs and K, the
phosphate ions were dissolved more rapidly than the zinc ions, whereas in the case of Li, the zinc was the
one to be dissolved at a higher rate.

http://www.jcse.org/volume3/paper5/v3p5.php?commentmode=2 6/20/2013
JCSE Volume 3 Paper 5 Page 12 of 27

Dissolution of the conversion layer is also suggested by Dickie,[52] after an investigation of the failure of
an epoxy coating. This mechanism is proposed to account for the slower delamination and longer delay
time observed with alkali resistant coatings.

An alternative view to the dissolution of the substrate theory is that the alkaline solution under the film
attacks the coating itself. Koehler’s investigation [53] of oleoresinous and polybutadiene coatings on
steel concluded that delamination was due to saponification of the polymers. In his discussion, he states
that disbonding may occur at pH 11.7, which is well below that required to dissolve the oxide layer. That
higher pH’s may form under the coating at cathodic sites is not disputed, however it is suggested that this
may take place subsequent to the loss of coating adhesion. Furthermore, it is proposed that regardless of
any other phenomena a film of water is required under the coating for disbonding to proceed.

Attack on the polymer is also a mechanism supported by Hammond et al [54] who examined three
different epoxy systems on steel with a coating range of 10 - 30 µm. In their examination of the
interfacial composition of delaminated surfaces, using XPS, predominant traces of polymer were in
evidence on the metal surface relative to iron oxide; this was in conjunction with little or no traces of iron
in the interfacial polymer. Their conclusion from this was that the failure mechanism was due to resin
degradation and that the locus of failure was within the polymer, resulting in a failure that was cohesive
in nature.

Another XPS investigation was undertaken by Castle and Watts [55], who looked at fusion-bonded epoxy
coated mild steel. The results of this study suggested that the failure was essentially adhesive in nature.
One of the variables in the experiments carried out was surface texture, ranging from grit blasted (Ra 3.80
µm) to a mirror finish polish with 1µm diamond paste (Ra 0.15 µm). It was found that as the Ra values
increased, there was a slight tendency toward cohesive failure of the oxide layer. However, although
some iron was transferred to the polymer, the amount was a trivial component of the locus of failure and
was consistent with cohesive failure of the oxide covering at protruding asperities.

Transport Pathways for Aggressive Species

The nature of the delamination mechanism is not the only area of debate within the overall subject of
cathodic delamination. Whilst there is agreement as to the species involved in the process, it is still open
to conjecture how those elements arrive at the delamination front. Considering a coated steel substrate,
immersed in an electrolyte of neutral or near neutral pH, the half-cell reaction responsible for the
delamination process is generally agreed to be oxygen reduction. This reaction generates OH- at the
cathodic site and is responsible for the alkaline environment at the delamination front. The elements
required for the process to proceed are water, oxygen and free electrons. The electrons may be generated
by either an anodic reaction or through the application of cathodic protection. Additionally, in order to
maintain electroneutrality, some form of counter ion is required; the high pH values recorded at the
reaction zone indicate that these counter-ions are alkali cations rather than protons.[56]

If we assume that cathodic delamination is a result of a damaged coating, there are two possible routes
that the reactants for the cathodic reaction may take. The two alternatives are either through the coating
or along the polymer-metal interface. An extensive review of the delamination process was carried out
by Leidheiser et al [56]. The results of the experiments carried out indicated that water was transported
to the reaction zone through the coating. It was suggested that a certain fraction of this could be in the
form of a cation such as H3O+, as the cathodic nature of the front may favour the transmission of water
associated with an ion possessing a positive charge. The supply of oxygen to the cathodic site was found
to be largely through the coating, with a small contribution from interfacial transport, in the case of the
epoxy coating studied. Other work on the subject of the supply of reactants to the delamination front

http://www.jcse.org/volume3/paper5/v3p5.php?commentmode=2 6/20/2013
JCSE Volume 3 Paper 5 Page 13 of 27

with pigmented coatings,[57] suggests that supply of oxygen along the interface may be the rate-
controlling factor.

The major unresolved factor in the overall picture is the route taken by the charge-neutralising cations.
Leidheiser’s study identified a linear relationship between the diffusion coefficient in aqueous solutions
of the different cations, used in his experiments, and the delamination rate. This suggested that the
cations were supplied along the interface. However, as delamination rates showed a strong dependence
on film thickness,[58] through film migration of cations could not be discounted.

An important feature identified in Leidheiser’s work was an incubation time needed before delamination
commenced. This was dubbed the “delay time” and the factor controlling this parameter was considered
to be cation ingress into the coating, due to an observed correlation between ionic mobility and the delay
time.

Interfacial migration of cations is supported by the work of Castle and Watts55. A number of factorsdrew
them to this conclusion. Firstly, EPMA examination of a cross section of the polymer showed Na ingress
into the coating from both the interface and the bulk solution; however there was no Na detected in the
bulk of the polymer, suggesting that Na+ did not fully penetrate the film. Also, the disbondment velocity
(Dk) was related to the surface profile of the substrate and the parameter referred to as tortuosity (τ) as
follows:

[2]

A tortuous interface is one whereby the route that must be taken by a species in order to diffuse from one
point to another, is made more difficult by roughening the substrate, thereby making the effective path
longer. It follows from this that a rougher surface should result in a lower lateral degree of delamination,
if the controlling mechanism is the interfacial migration of a species.

Castle and Watts’ results for a steel/epoxy system produced a constant disbondment velocity (Dk*) of 0.4
± 0.08 mm day-1. This again was felt to point towards interfacial migration as being important to the
process.

The effect of the substrate upon the delamination process was discussed by Skar and Steinsmo [59] in an
investigation of the importance of through film ionic migration. Their experiment, which aimed to clarify
the confusion as to the route taken by the cationic species, showed a strong linear relationship between
the dry film thickness and the disbonding rate. The explanation offered for this result considered three
alternative routes by which the counter-ions may reach the delamination front:

1. Along the interface.


2. Through the detached coating.
3. Through the attached coating ahead of the delamination front.

These alternative routes are illustrated in Figure 7. The first option was discounted due to the strong
dependence on film thickness. Of the remaining two alternatives, it was felt that the second was the most
likely.

http://www.jcse.org/volume3/paper5/v3p5.php?commentmode=2 6/20/2013
JCSE Volume 3 Paper 5 Page 14 of 27

Figure 7: Suggested ionic migration routes from ref. [59]

The logic behind this conclusion stems firstly, from work carried out by Mayne and Mills,25 whose
investigation of detached films showed a notable difference between the ionic resistances of attached,
compared to detached, films. Secondly, it was proposed that the conduction mechanism of the detached
film exhibited “D” type behaviour, as defined by Kinsella,[60] due to the high pH under the coating.
Whilst the second proposition may well be valid, the first is dubious. The conclusions drawn by Mayne
and Mills, were indeed that detached films showed lower resistances, as quoted, but not on ordinary mild
steel substrates - where the detached films showed similar resistances to those attached to the steel. It
was when the coatings were applied to inert substrates (platinum and steel passivated with a zinc
chromate pigment) that the quoted difference in resistance was found.

Working with both pigmented and unpigmented chlorinated rubber coatings, Sharman [57] identified a
number of features worthy of note. Referring to the delay time defined by Leidheiser, he concluded for
the system under investigation that it was essentially independent of cation mobility. Furthermore, it was
found that the iron oxide pigmented systems showed longer delay times than the clear varnish coated
substrates; this was unexpected as the former had higher ionic permeabilities. This situation was
considered by pre-soaking the pigmented coatings in water prior to testing; this resulted in a reduction in
the delay times from 5-7 hours to 1-2 hours. This result pointed to ingress of either water or oxygen as
the factor responsible for the duration of the delay time.

With regard to the species responsible for the delamination process and the route taken, different
mechanisms were seen to be responsible for the two systems. For the unpigmented coating, interfacial
cation migration was considered to be the determining factor as the delamination rate constant increased
in a linear fashion with cation mobility. With the pigmented coating, which displayed a greater through-
film ionic transport rate, coupled with lower oxygen permeability, interfacial migration of oxygen was
suggested as a likely rate determining mechanism; this was suggested to be virtually independent of
cation mobility. A further result, which reinforced the hypothesis of interfacial migration of a species as
the rate-determining factor, was that the rate of disbonding eventually became zero. The explanation for
this was that the interfacial pathways became blocked with corrosion products.

Armstrong and Johnson,[61] working with unpigmented chlorinated rubber coatings subjected to
cathodic protection, concluded that the corrosion current passed through faults in the form of cracks or
crevices in the film. It is proposed that degradation of the film due to OH- was essentially mechanical in
nature. This argument is supported by a number of factors. Free standing films exhibited conductive
behaviour only after long exposure to concentrated OH-; this was lost after the films were dried in a
dessicator. It was felt this result supported the hypothesis of conduction through cracks or crevices rather
than through the bulk film. FTIR (Fourier Transform Infra Red) spectra of membranes exposed to OH-

http://www.jcse.org/volume3/paper5/v3p5.php?commentmode=2 6/20/2013
JCSE Volume 3 Paper 5 Page 15 of 27

displayed no detectable chemical change from the untreated film.

The underside of the films (that in contact with the metal), which was subjected to the alkaline
conditions, was found to contain holes that were not visible on the upper face of the coating. This is
thought to be due to pockets of high solvent or plasticiser concentration; it is suggested that the OH-
opens up such defects by a solvent cracking process, which reinforces the assertion that the breakdown is
mechanical rather than chemical.

The Substrate

In an ideal world, the coatings that we applied to protect against corrosion would be a perfectly
homogeneous mix and be free of defects of any kind. Also, the surface to which the coatings were
applied would be in an ideal condition, free of any form of contamination and with an appropriate finish.

This utopian situation does not occur in the real world and the substrates are commonly far from pristine.
There are occasions where adequate preparation of the surface to be coated is either technically
unfeasible or prohibitively expensive.[62] In such circumstances, there is a need for a coating that will
perform well on a contaminated substrate. Mayne’s12 model of underfilm corrosion, considered earlier,
was applicable to a clean surface. In the presence of contaminants on the surface, resistance inhibition
does not come into play, as all the required elements for corrosion to proceed are present at the interface.

Effects of Contamination
Surface contamination is conventionally considered to be due to such factors as salt deposits or
corrosion. Indeed, this is the source of a great deal of concern and has been the basis of much work[63],
[64] [65] [66] [67]. Additionally, it has been demonstrated that, even on substrates cleaned to laboratory
standards, minor inclusions such as sulphides can have a detrimental effect on the performance of an
organic coating.[68]

An established method for dealing with the problem of painting over steel that is rusted, contaminated
with salts or, as is usually the case, a combination of the two, has been to coat with an oil-based coating
pigmented with red lead (Pb3O4). Environmental concerns have limited the use of lead compounds[69]
and forced formulators to look to alternative means of addressing the problem.[70] In the pursuit of a
viable alternative, the red lead formulation is still used as a benchmark by which to compare the various
contenders [71], [72], [73]. The exact mechanisms by which red lead protect a compromised substrate are
unclear, but its efficiency is undisputed.

Rust, in itself, is not seen as a problem;[74] rather it is the contaminants that the rust contains. Mayne
[75] identified that sulphates were embedded into adherent rust close to the metal surface; this rust layer
was not easily removed by wire brushing. These sulphate “nests” are felt, to a large degree, to be
responsible for the formation of blisters at anodic sites [71]. Such blisters lead to failure, due to local
bulging and eventual cracking of the coating.

Mayne’s experiments[76] were carried out in Cambridge, over the period 1945 to 1953. The
contamination of his specimens was determined to be due to the burning of fossil fuels, this was justified
by a seasonal variation in the degree of contamination giving higher levels in the winter months. One
could possibly argue that with the advent of smokeless zones in Britain that, in non-industrial urban
areas, this problem should be greatly alleviated. Having said this, industrial areas still provide an
adequate level of contamination to require attention. The problem of sulphates in industrial areas
becomes one of chlorides in marine environments.

http://www.jcse.org/volume3/paper5/v3p5.php?commentmode=2 6/20/2013
JCSE Volume 3 Paper 5 Page 16 of 27

Surface Tolerant Coatings

In a search for environmentally acceptable coatings that may be applied over compromised metal
surfaces, attention has been focused on the field of “surface tolerant coatings”. Frondistou-Yiannas [73]
surveyed the range of these coatings and their reported modes of action. The mechanisms considered
were: conversion of rust to magnetite, conversion of iron oxides to other compounds and inertisation of
soluble salts. In addition to these, Thomas [74] also considered barrier coatings, the ability of the coating
to wet and penetrate the rust and anti-corrosive pigments. The field trials of the former investigator,
which were carried out in both marine and industrial areas, concluded that none of the systems tested was
expected to provide long-term protection in the aggressive environments used in the tests.

So called rust converters are mainly formulations based upon tannic acid. An extensive review of the
mechanisms involved in the action of tannin based coatings,[77] details the reaction mechanisms by
which it is claimed that tannins help to protect steel. The action of tannin-based treatments is explained
by the inhibition of the formation of magnetite, which if allowed to develop, would increase the area
available for oxygen reduction. An alternative name of “rust deactivators” is proposed for these
formulations, which may lead one to think that they have some effect upon the contaminants that are
invariably present. However, this statement is followed by the fact that tannins are unable to neutralise
surface contaminants such as sulphates.

In her review of rust conversion preparations, Thomas [74] states that the premise behind their use
consists of the stabilisation of the rust, so that the redox reaction between Fe(II) and Fe(III) is no longer
possible. In this case, in principle, corrosion would be prevented. It is pointed out, however, that tannic
acid reacts with Fe(III) within regions underlying thick layers of surface rust; therefore, it would be
necessary to wire brush the surface prior to treatment.

A further assessment of the action of tannic acid has been conducted by Morcillo et al.[78] Panels that
had been exposed outdoors for up to two years were thoroughly wire brushed and immersed in tannic
acid of various concentrations. Fresh panels were also immersed, to assess the effect on clean steel. A
film of ferric tannate was found to form on the surface of the rusty panels, this film was found to be
extremely porous. The results from these tests showed that tannic acid promotes rusting on clean steel
and does not significantly reduce that of rusted steel. Furthermore, it was suggested that due to the nature
of the ferric tannate film formed on the surface, that the tannic acid might attack the iron at the base of the
pores.

Perhaps the most poignant point comes from Frondistou-Yiannas’ paper where, all the rust converter
systems under examination did not make it to the field trials, as they failed the initial laboratory screening
due to blistering and intercoat disbonding.

An alternative method of dealing with contaminated surfaces proposed by Sykes[79] is the use of water-
borne paint. The contention here is that if the requirements of the coating fall into any of the following
criteria:

Removal of soluble salts from the surface.


Conversion of salts into safer species.
Electrochemical alteration of rust involving the underlying steel.

Then an aqueous medium, in the form of a water-borne coating probably offers the best chance of
success; providing a method to either dissolve or bind the soluble ions while the paint is drying.

Experimental work was carried out on a 120µm thick water-borne vinyl acrylic latex system applied over

http://www.jcse.org/volume3/paper5/v3p5.php?commentmode=2 6/20/2013
JCSE Volume 3 Paper 5 Page 17 of 27

pre-rusted steel. The substrates were coated in the as-received condition, wire brushed or grit blasted
prior to application of the coating. The formulation from which the coating was taken normally has its
pH adjusted to 4.5 - 5.0. During the investigation, the pH was taken down to 2.7. This lower pH was
found to aid in the removal of light deposits of rust resulting in a decrease in the degree of blistering of
the coating.

Effect of Surface Preparation Techniques

Another feature highlighted in this work was that dry grit blasting could exacerbate the problem of
coating rusty steel by exposing contaminants buried in the rust layer. This feature has been noted
elsewhere[80]. Research sponsored by Nuclear Electric determined that dry blasting not only exposes
buried salts but also spreads them.[81] The research was primarily aimed at the determination of an
acceptable level for residual salt contamination. Substrates were pre-contaminated with two salt
compositions as shown in Table 1, prior to being coated with a three-coat alkyd paint system.

Table 1: Composition of salt mixtures, after ref. [81]

Salt Concentration
(% by weight)
Marine Sodium Chloride 81
Magnesium Chloride 19
Urban Ammonium Sulphate 41
Sodium Nitrate 37
Sodium Chloride 12
Sodium Nitrate 10

Exposure tests were carried out on panels contaminated with progressively higher levels of salt
contamination. The exposure trial ran for 6 ½ years, after which pull-off adhesion tests were carried out.
With zero contaminant, the mode of failure was cohesive within the primer coat. As the salt levels
increased, progressively less primer remained adhered to the steel surface. At higher contamination
levels, there was a change from mixed to total adhesive failure of the primer. This change was abrupt and
corresponded to the values shown in Table 2

Table 2: Critical Contamination Levels Applicable to Different Environments.

Critical
Environment Chloride Level
Contamination Level
2 µg cm-2 Marine 81%
6 - 20 µg cm-2 Urban 12%

The conclusion drawn by this investigation was that the adhesion, and therefore the performance, of the
system was severely affected by salt contamination of the substrate, where the levels were in excess of 2
µg cm-2 when the salts were predominantly chlorides.

An epoxy coating used as a tank lining was also tested. In this case, both the marine and the urban salt
contamination resulted in coating failure at levels above 20µg cm-2. The marine salt, however, brought
about the failure in 6 days compared with 56 days for the urban salt composition.

The more aggressive nature of the chloride ion is attributed to higher osmotic pressures exerted.

http://www.jcse.org/volume3/paper5/v3p5.php?commentmode=2 6/20/2013
JCSE Volume 3 Paper 5 Page 18 of 27

Additionally, chlorides tend to be more soluble than other salts and of a lower formula weight. Hence, a
given salt burden of a chloride represents a higher molar concentration, which produces a higher osmotic
pressure.

As often, in the field of organic coatings, the literature provides what appears at first sight, to be
contradictory data. The critical chloride levels determined by Morcillo [82] for example was determined
to be 50µg cm-2. It is important that one realises that each different coating/metal system is likely to have
various parameters, including the chloride levels it can tolerate, that are unique to itself.

Another effect of contaminated surfaces found by Morcillo was the degradation of a chlorinated rubber
primer applied over contaminated rusted steel.[83] The chloride levels found under the films in this
investigation were too high to be accounted for by the initial contamination or the exposure environment.
An XPS analysis was carried out which indicated that the C-Cl bonds in the polymer underwent a
chemical transformation that yielded Cl- and accounted for the increased chloride concentration in the
rust at the interface. So the, initially present or subsequently formed, interfacial rust appeared to catalyse
the decomposition of the polymer.

Morcillo discussed a further aspect of the importance of proper surface preparation, in a paper that
considered the significance of the size and the type of abrasive used during pre-cleaning operations.[84]
One important point highlighted was that there is a critical surface profile, the value of which is
determined by the environment along with the type and thickness of the coating. As the coating increases
in thickness, the effect of the surface profile diminishes. The critical surface profile was found to be a
function of the aggressiveness of the environment – a more aggressive environment resulted in a lower
critical profile.

In conclusion, the surface to be coated is as important a part of the overall system as any other, so the
state and condition of the substrate prior to painting significantly affects the demands that are placed
upon the coating. Traditionally tried and tested methods for coating compromised steel are no longer
environmentally acceptable, therefore new, safe, ecologically sound alternatives are demanded. Should
the emerging water-borne paint technology provide a solution to the contamination problem, it would
also have the added advantage of addressing the impending volatile organic content legislation.

Adhesion

Within the debate regarding the important features of the protection offered by organic coatings, the
argument over the importance of adhesion is amongst the most contentious. It is often claimed (or at
least inferred) that the adhesion is of paramount importance, to the exclusion of all other features.
According to Bullett and Prosser “the ability to adhere to the substrate throughout the desired life of the
coating is one of the basic requirements of a surface coating, second only to the initial need to wet the
substrate”,[85] Funke describes adhesion as the most important and decisive property of a coating.[86]

In order to enter into the debate as to the importance of adhesion in the protective mechanisms of paint
systems, it is necessary to consider the process of adhesion itself. In particular, the principle of adhesive
failure needs to be addressed. For two surfaces to adhere to each other, be they of the same or of
different materials, there is a need for intimate contact on an atomic scale between them. In the case of a
liquid coating, the degree of intimate contact is governed by the ability of the coating to wet the
substrate. This wetting ability needs to be present not only in the static sense, where the contact angle
between the substrate and the liquid may be used to determine how efficient the wetting is; but also in a
dynamic sense, where the rate of wetting of the substrate is not countered by the rate of shear involved in
the application of the coating.[87]

http://www.jcse.org/volume3/paper5/v3p5.php?commentmode=2 6/20/2013
JCSE Volume 3 Paper 5 Page 19 of 27

According to Bikerman[88] a true adhesion failure is improbable, a view shared by others.[89], [90] If
one subscribes to this point of view, then it must be the case that all failures that involve detachment of
the coating from the substrate, must include some element of either cohesive failure of the coating or the
oxide layer on the surface of the steel. This view is reinforced by XPS work which has been carried out
by several workers, who have found that traces of polymeric material is normally found on the “metal”
surface of a metal/polymer interfacial fracture, which may have appeared to be a purely adhesive failure
from a visual examination.[54], [55], [91] Although the explanation of a cohesive polymer or oxide
failure may account for the observed failure modes, this is probably still an oversimplification.

The function of the adhesive joint is to effectively remove the interface from between the two different
materials. In order to do this, a gradual change must occur in the chemical composition of the coating
and the substrate in the interfacial area. Plato elucidated this concept in his explanation of how the
different elements in the universe could exist as one:

It is not possible for two things to be fairly united without a third for they need a bond between them
which shall join them both, that as the first is to the middle, so is the middle to the last, then since the
middle becomes the first and the last, and the last and the first become the middle, of necessity, all will
come to be the same. And being the same with one another, all will be a unity.[92]

The inference of this is that in a true adhesive joint, there is no clearly defined interface and that one
needs to see the interfacial area as a gradual transitory phase where the coating and the substrate affect
and are effected by each other. So, in the same way that the atomic structure of the metal gradually
changes from that of the bulk phase to that of the oxide on the surface – it is proposed that the properties
of the surface of the uncoated metal are changed when a polymeric coating is applied to it. In the same
way, Leidheiser suggests that the structure of the coating varies as one passes through the bulk of the film
to the interfacial region with the metal. This region, which extends into the polymer, is known as the
interphase. The size of the interphase is not known, but is thought to be in the nanometer range.[93]

Whilst loss of adhesion may occur under a number of situations, a major form of this type of failure is
loss of adhesion resulting from exposure to aqueous or humid environments.[94], [95], [96], [97] As a
result of work into loss of adhesion due to water ingress, it has been concluded that adhesion tests in the
dry state are of little use for applications in immersed or humid environments.[98], [99] Indeed, it has
been proposed that, in isolation, adhesion tests are insufficient to determine the ability of a coating to
control corrosion.[100]

It is true that the adhesion plays an important role in the protective mechanism of coatings, but one
should take care when claiming that it is the overriding consideration. When one considers the process of
cathodic delamination it is clear that, once the paint has become detached from the substrate, the
underlying metal is exposed to the environment and is no longer afforded any protection from the coating
system. One should, however, be careful to ensure that the whole process is viewed in context. Whilst
the loss of adhesion, resulting from the delamination process, effectively negates the protection afforded
by the coating, it is important to consider whether the original adhesion is the deciding factor in the
delamination process. Gowers and Scantlebury suggested that the beneficial role of the adhesion of a
paint coating is due to the impairment of the formation of a layer of electrolyte at the coating/substrate
interface, preventing ionic current flow and the spread of corrosion over the surface.[101]

A case in point is that of the use of silanes as adhesion promoters. There has been a considerable amount
of work carried out in this field and the results may, at first sight, appear contradictory. The fundamental
reason why the results of various investigations lead to different conclusions is the starting point and the
assumptions made by the investigators.

http://www.jcse.org/volume3/paper5/v3p5.php?commentmode=2 6/20/2013
JCSE Volume 3 Paper 5 Page 20 of 27

Some work has been carried out which has shown that the use of silane coupling agents leads to an
improvement in wet adhesion.[89] This result has been used to assert that this treatment will therefore
result in enhanced protection against such phenomena as underfilm corrosion or cathodic delamination.
This claim is based on the assumption that good adhesion to the substrate will necessarily lead to
improvement in the protection afforded.

The work of Marsh[102] et al has tested this assumption and has shown that although silanisation does in
fact provide better wet adhesion, the treatment afforded no improvement in resistance to delamination
and that underfilm corrosion was significantly worse on the silane treated specimens. It was pointed out
by Marsh that the use of the cathodic delamination test as a determinant of adhesive strength is
inappropriate, as the work showed that improved adhesion played no part in the rate of delamination.
Furthermore, the cathodic delamination test is therefore a measure of cationic mobility rather than
adhesive strength. Other work on salt spray exposure of coated panels has found that no adhesion loss
had occurred although severe underfilm corrosion had taken place.[103]

Good surface preparation is the key to good adhesion but the type, as well as the condition, of the
substrate has been found to have a strong influence upon the initial dry, and the subsequent wet, adhesion
of a metal/polymer system. Gosselin showed that the metal used as the substrate affected the strength of
the final dry bond and Arslanov found that the adhesive strength of a coating on aluminium initially
decreased upon exposure to water, but subsequently got stronger.[104]

One method that has been suggested by Funke to increase the adhesive strength of coated systems is the
application of thin layer technology.[105], [106] Here, it is suggested that the adhesion to the substrate
could be enhanced as a result of co-operative bonding between the intermediate thin polymer layer and
the main bulk of the coating.

The proposed thickness of these coatings is in the nanometer range. [107] While this is in conflict with
common coating practice whereby the base primer coat is thicker than the surface profile of the metal,
when viewed in conjunction with Leidheiser’s contention regarding the interfacial region,[93] and
Kumins’ model of restricted chain mobility within thin layers;[108] this may well be an avenue worthy of
investigation.

In conclusion, though the adhesion of a coating to the substrate is an important factor, which must be
taken into consideration when addressing the mechanisms of corrosion protection afforded by the
coating, the overall picture is much more complex and the overall mechanism that must be considered is
a combination of various contributory factors. Karyakina and Kuzmak have produced an extensive
overview of the processes of adhesion and other contributory factors, together with a review of
experimental techniques for the evaluation of coating properties.[109]

References
[1] Mayne J.E.O. Paints For The Protection Of Steel - A Review Of Research Into Their Modes Of
Action. Br. Corr. Journal. Vol. 5, May 1970, pp 160 - 111.

[2] Hare C.H. Barrier Coatings. J. Prot. Coat. Linings, Feb 1989, pp 59 - 69.

[3] Royston I. A Ha'p'orth O' Tar, SURCON '97, Birmingham, Sept 1997.

[4] Asbeck W.K. and Van Loo M. Critical Pigment Volume Relationships. Ind. Eng. Chem. Vol. 41, No.
7, 1949, pp 1470 -1475.

http://www.jcse.org/volume3/paper5/v3p5.php?commentmode=2 6/20/2013
JCSE Volume 3 Paper 5 Page 21 of 27

[5] Bierwagen G.P. Critical Pigment Volume Concentration As A Transition Point In The Properties Of
Coatings. J.Coat.Tech. Vol. 64, No. 806, 1992, pp 71 - 75.

[6] Bell S.H. The Structure Of Paints. JOCCA, Vol. 38, Oct 1955, pp 595 - 623.

[7] Thomas, N.L. The Barrier Properties of Paint Coatings Prog Org Coat, Vol. 19, 1991, pp 101-121

[8] Dickie R.A. Smith A.G. How Paint Arrests Rust, Chemtech, January 1980, pp 31 - 35.

[9] Bacon C.R. Smith J.J. and Rugg F.M. Electrolytic Resistance in Evaluating Protective Merit of
Coatings on Metals. Ind Eng Chem, Vol.40, No 1, 1948 pp 161-167.

[10] Kittelberger W.W. and Elm A.C. Water Immersion Testing of Metal Protective Paints: Role of
Osmosis in Water Absorption and Blistering. Ind Eng Chem Vol. 38. No. 7. 1946, pp 695-699.

[11] Mayne J.E.O. The mechanism of the protective action of an unpigmented film of polystyrene JOCCA
Vol. 32 No 352 1949, pp 481-487.

[12] Mayne J.E.O. The Mechanism of the Inhibition of the Corrosion of Iron and Steel by Means of Paint
Official Digest, Feb. 1952, pp 127-136.

[13] Corti H. Fernádez-Prini R. and Gómez D. Protective Organic Coatings: Membrane Properties and
Performance Prog Org Coat, Vol. 10, 1982, pp 5-33.

[14] Skoulikidis T. and Ragoussis A. Diffusion of Iron Ions Through Protective Coatings on Steel,
Corrosion, Aug 1992, pp 666-670.

[15] Maitland C.C. and Mayne J.E.O. Factors Affecting the Electrolytic Resistance of Polymer Films
Official Digest, September 1962.

[16] Cherry B.W. and Mayne J.E.O. The Resistance Inhibition of Corrosion in Unpigmented Systems
Official Digest Vol. 33 No. 435. 1961, pp 469-480.

[17] Cherry B.W. and Mayne J.E.O. Ionic Conduction Through Varnish Films 1st Int Cong on Mett Corr
London 1961, pp 539-544.

[18] Miskovic-Stankovic V.B. Drazic D.M. and Teodorovic M.J. Electrolyte Penetration Through Epoxy
Coatings Electrodeposited on Steel Corrosion Science, Vol. 37 No. 2, 1995, pp 241-252.

[19] Kinsella E.M. and Mayne J.E.O. Ionic Conduction in Polymer Films 3rd International Congress on
Metallic Corrosion, Moscow 1966, pp 117-120.

[20] Kinsella E.M. and Mayne J.E.O. Ionic Conduction in Polymer Films: I. Influence of Electrolyte on
Resistance. Br. Polym. J. Vol. 1. July 1969, pp 173-176.

[21] Mayne J.E.O. and Scantlebury J.D. Ionic Conduction in Polymer Films: II Inhomogeneous Structure
of Varnish Films. Br. Polym. J. Vol. 2, September 1970, pp 240-243.

[22] Muizebelt W.J. & Heuvelsland W.J.M. Permeabilities Of Model Coatings: Effect Of Cross Link
Density And Polarity. Polymeric materials for corrosion control. ACS symposium 322 1986, pp 110-114.

http://www.jcse.org/volume3/paper5/v3p5.php?commentmode=2 6/20/2013
JCSE Volume 3 Paper 5 Page 22 of 27

[23] Kinsella E.M. Mayne J.E.O. & Scantlebury J.D. Ionic Conduction In Polymer Films III Influence Of
Temperature On Water Absorption. , Br Polym J, Vol. 3, 1971, pp 41 - 43.

[24] Mayne J.E.O. & Scantlebury J.D. Ionic Conduction In Polymer Films IV The Effect Of
Pigmentation With Iron Oxide, Br Polym J, Vol. 3, 1971, pp 237 - 239.

[25] Mayne J.E.O. and Mills D.J The Effect of the Substrate on the Electrical Resistance of Polymer
Films. JOCCA Vol. 58, 1975, pp 155-159.

[26] Mills D.J and Mayne J.E.O. The Inhomogeneous Nature of Polymer Films and its Effect on
Resistance Inhibition. in Corrosion Control by Organic Coatings Ed H.Leidheiser Jr. 1981, pp 12-17.

[27] Mills D.J. PhD Thesis, Cambridge, 1973.

[28] Gay P.J. Blistering of paint films on metal, JOCCA, Vol. 32, No. 352, 1949, pp 488-498.

[29] Mayne J.E.O. The Blistering of Paint Film. Part II Blistering in the Presence of Corrosion. JOCCA
Vol. 31. No 12, 1950, pp 538-547.

[30] Leidheiser H. Jr. Corrosion of painted metals - A review Corrosion-NACE, Vol. 38, No. 7, 1982, pp
374 - 383.

[31] Appleman B.R. Painting over soluble salts: A perspective J. Prot. Coat. Lin. Oct 1987, pp 68 - 82.

[32] Hare C.H. Blistering Of Paint Films On Metal, Pt.1: Osmotic Blistering, J.Prot.Coat.Lin. Feb 1998,
pp 45-63.

[33] Funke W Toward a unified view of the mechanism responsible for paint defects by metallic
corrosion Ind. Eng. Chem. Prod. Res. Dev. Vol. 24, No. 3, 1985, pp 343 -347.

[34] van der Meer-Lerk L.A. & Heertjes P.M. Mathematical model of growth of blisters in varnish films
on different substrates, JOCCA. Vol. 62, 1979, pp 256 - 263.

[35] Gowers K.R. & Scantlebury J.D. Blistering Phenomena On Lacqured Mild Steel, Corrosion
Science, Vol. 23, No. 9, 1983, pp 935-942.

[36] Storfer S.J. & Yuhas S.A. Jr. Mechanism of Blister Formation in Organic Coatings, Mat Perf, Jul
1989, pp 35-41.

[37] Koehler E.L. Underfilm corrosion currents as the cause of failure of protective organic coatings
Corrosion Control by Organic Coatings, Ed. Leidheiser H. Jr. 1981, pp 87 -96.

[38] Schwenk W. Adhesion Loss For Organic Coatings Causes And Consequences For Corrosion
Protection. Corrosion Control by Organic Coatings, 1981, pp 103-110.

[39] Hare C.H. Non-Osmotically-Induced Blistering Phenomena On Metal, J Prot. Coat Lin, Mar 1998,
pp 17-34.

[40] Nguyen T. Hubbard J.B. & McFadden G.B. A mathematical model for the cathodic blistering of
organic coatings on steel immersed in electrolytes J. Coat. Tech. Vol. 63, No. 794, 1991, pp 43 - 52.

http://www.jcse.org/volume3/paper5/v3p5.php?commentmode=2 6/20/2013
JCSE Volume 3 Paper 5 Page 23 of 27

[41] Kittleberger W.W. Water Immersion Testing Of Metal Protective Paints. Influence Of Bare Metal
Areas, Ind Eng Chem, Vol. 34, No. 9, 1942, pp 943-948.

[42] Leidheiser H. Jr. Cathodic delamination of polybutadiene from steel - A review J. Adhes. Sci. Tech.
Vol. 1, No. 1, 1987, pp 79 - 98.

[43] Leidheiser H. Jr. Alkali metal ions as aggressive agents to polymeric corrosion protective coatings
Corrosion - NACE, Vol. 43, No. 5, 1987, pp 296 - 297.

[44] Smith A.G. & Dickie R.A. Adhesion failure mechanisms of primers Ind. Eng. Chem. Prod. Res. Dev.
Vol. 17, No. 1, 1978, pp 42 - 44.

[45] Koehler, E.L. The influence of contaminants on the failure of protective organic coatings Corrosion
- NACE, Vol. 33, No. 6, 1977, pp 209 - 217.

[46] Ritter J.J. Ellipsometric studies on the cathodic delamination of organic coatings on iron and steel
J. Coat. Tech. Vol. 54, No 695, 1982, pp 51 - 57.

[47] Ritter J.J. & Kruger J. Studies on the subcoating environment of coated iron using qualitative
ellipsometric and electrochemical techniques in Corrosion Control by Organic Coatings, NACE, Ed. H
Leidheiser Jr. 1982, pp 28 - 31.

[48] Castle J.E. & Watts J.F. Cathodic Disbondment Of Well Characterised Steel/Coating Interfaces
Corrosion Control by Organic Coatings, Vol. 1981, pp 78 - 86.

[49] Wiggle R.R. Smith A.G. & Petrocelli J.V. Paint adhesion failure mechanisms on steel in corrosive
environments J. Paint Tech. Vol. 40, No. 519, 1968, pp 174 - 186.

[50] Knaster M. & Parks J. Mechanism of corrosion and delamination of painted phosphated steel during
accelerated corrosion testing, J. Coat. Tech. Vol. 58, No. 738, 1986, pp 31 - 38.

[51] Sommer A.J. & Leidheiser H. Jr. Effect of alkali metal hydroxides on the dissolution of a zinc
phosphate conversion coating on steel and pertinence to cathodic delamination Corrosion, Vol. 43, No.
11, 1987, pp 661 - 665.

[52] Dickie R.A. Chemical Studies Of The Organic Coating/Steel Interface After Exposure To Aggressive
Environments, Critical Issues in Reducing the Corrosion of Steels, 1985, pp 379-396.

[53] Koehler E.L. The mechanism of cathodic disbondment of protective organic coatings - aqueous
displacement at elevated pH, Corrosion, Vol. 40, No. 1, 1984, pp 5 - 8.

[54] Hammond J.S. Holubka J.W. deVries J.E. & Dickie R.A. The application of X-ray photo-electron
spectroscopy to a study of interfacial composition in corrosion-induced paint de-adhesion Corrosion
Science, Vol. 21, No. 3, 1981, pp 239 - 253.

[55] Watts J.F. & Castle J.E. The application of X-ray photoelectron spectroscopy to the study of
polymer-to-metal adhesion J. Mat. Sci. Vol. 19, 1984, pp 2259 - 272.

[56] Leidheiser H. Jr. Wang W. & Igetoft L. The mechanism for cathodic delamination of organic
coatings from a metal surface Prog. Org. Coat. Vol. 11, 1983, pp 19 - 41.

http://www.jcse.org/volume3/paper5/v3p5.php?commentmode=2 6/20/2013
JCSE Volume 3 Paper 5 Page 24 of 27

[57] Sharman, J.D.B. Sykes, J.M. & Handyside,T. Cathodic disbonding of chlorinated rubber coatings
from steel, Corrosion Science, Vol. 35, No. 5-8, 1993, pp 1375-1383.

[58] Leidheiser H. Jr. & Wang W. Some substrate and environmental influences on the cathodic
disbonding of organic coatings J. Coat. Tech. Vol. 53, No. 672, 1981, pp 77 - 84.

[59] Skar J.I. & Steinsmo U. Cathodic disbonding of paint films - transport of charge Corrosion Science,
Vol. 35, Nos. 5 - 8, 1993, pp 1385 - 1389.

[60] Kinsella E.M. PhD Thesis Cambridge 1967

[61] Armstrong R.D. & Johnson B.W. An investigation into the cathodic delamination of unpigmented
chlorinated rubber films Corrosion Science, Vol. 32, No. 3, 1991, pp 303 - 312.

[62] Baxter I.K. Surface tolerant protective coatings, Step Into the 90’s, Transactions of the 1st Joint
Conference on Corrosion, Finishing and Materials, Broadbeach, Queensland, 1989.

[63] Gross, H Examination Of Salt Deposits Found Under German Painted Steel Bridge Decks, Materials
Performance, Vol. 22, No. Oct 1983, pp 28-33.

[64] Feliu S. Bastidas J.M. Galván J.C. Feliu S.Jr. Simancas J. & Morcillo M. Electrochemical
Determination Of Rusted Steel Surface Stability. , J. Applied Electrochemistry. Vol. 23, No. 2, 1993, pp
157 - 161.

[65] Feliu S. Galvan J.C. Feliu S.Jr. Bastidas J.M. Simancas J. Morcillo M. & Almeida E.M. An
Electrochemical Impedance Study Of The Behaviour Of Some Pretreatments Applied To Rusted Steel
Surfaces. Corrosion Science. Vol. 35, No. 5-8, 1993, pp 1351-1358.

[66] Morcillo M. The Detrimental Effects Of Water-Soluble Contaminants At The Steel/Paint Interface.
Panoramic View Of The Authors Research On The Topic, Preceedings of 12th International Corrosion
Congress. Sept 19-24 Houston, Texas. NACE INTERNATIONAL. 1993, pp 87 - 98.

[67] Neal D and Whitehurst T Chloride Contamination Of Line Pipe, Its Effect On FBE Coating
Performance , Mat. Perf. Feb 1995, pp 47-52.

[68] Costa I. Faidi S.E. & Scantlebury J.D. Substrate effects on the corrosion performance of coated
steels under immersed conditions Corrosion Control for Low Cost Reliability, 12th International
Corrosion Congress, NACE, Houston, Texas, 1993, pp 437 - 448.

[69] Mickalonis J.I. & Leidheiser H. Jr. Corrosion inhibition of steel by lead pigments Corrosion NACE,
Vol. 45, No. 8, 1989, pp 631 - 636.

[70] Bittner A. Advanced Phosphate Anticorrosive Pigments For Compliant Primers. J.Coat.Tech. Vol.
61, No. 777, 1989, pp 111 - 118.

[71] Scantlebury J.D. Organic Coatings Systems And Their Future In Corrosion Protection, Paper
presented at Eurocorr 82, Budapest 1982

[72] Thomas, Noreen L The Protective Action of Coatings on Rusty Steel. J. Protective Coatings &
Linings, Vol. 6, No. 12, 1989, pp 63 - 71.

http://www.jcse.org/volume3/paper5/v3p5.php?commentmode=2 6/20/2013
JCSE Volume 3 Paper 5 Page 25 of 27

[73] Frondistou-Yiannas S. Evaluation of rust tolerant coatings for severe environments J. Prot. Coat.
Lin. August 1986, pp 26 - 35.

[74] Thomas N.L. Coatings for rusty steel: Where are we now? Surface Coatings International (JOCCA),
Vol. 74, No. 3, 1991, pp 83 - 88, 97.

[75] Mayne J.E.O. The problem of painting rusty steel J. Appl. Chem. Vol. 9, Dec. 1959, pp 673 - 680.

[76] Mayne J.E.O. Current Views on How Paint Films Prevent Corrosion, JOCCA. Vol. 40, Mar, 1957,
pp 183 - 199.

[77] Landolt D. & Favre M. Environment-friendly coatings for steel based on tannins: A critical review
and new results Progress in the Understanding & Prevention of Corrosion, 10th European Corrosion
Congress, Eds. Costa J.M. & Mercer A.D. Inst. Of Mats. 1993, pp 374 - 386.

[78] Morcillo M. Feliu S. Simancas J. Bastidas J.M. Galvan J.C. Feliu S. Jr. & Almedia E.M. Corrosion
of rusted steel in aqueous solutions of tannic acid Corrosion, Vol. 48 No. 12, 1992, pp 1032 - 1039.

[79] Sykes J.M. Smith H.E.M. Moreland P.J. & Padget J.C. Protection of poorly prepared rusty steel
surfaces by water borne paint Advances in Corrosion Control by Organic Coatings II, Eds. Scantlebury
J.D. & Kendig M. 1994, pp 7 - 14.

[80] McKelvie A.N. Steel Cleaning Standards - A Case For Their Reappraisal, JOCCA, Vol. 60,1977,
pp 277-237.

[81] Allan S.J. May R. Taylor M.F. & Walters J. The effect of salts on steels and protective coatings
GEC Journal of Research, Vol. 12, No. 2 1995, pp 86 – 92

[82]. Morcillo M. Feliu S. Galvan J.C & Bastidas J.M. Some observations on painting contaminated rusty
steel J. Prot. Coat. Lin. Vol. 4, No. 9, 1987, pp 38 - 43.

[83] Morcillo M Simancas J. Fierro J.L.G. Feliu S. Jr. & Galvan J.C. Accelerated degradation of a
chlorinated rubber paint system applied over rusted steel Prog. Org. Coat. Vol. 21, 1993. Pp 315 - 325.

[84] Morcillo M. Bastidas J.M. Simancas J Galván J.C. The Effect Of The Abrasive Work Mix On The
Paint Performance Over Blasted Steel. Anti Corrosion Methods & Mats., Vol. 36, No. 5, 1989, pp 4 - 8.

[85] Bullett, T.R. & Prosser, J.L. Measurement Of Adhesion, Prog Org Coat, Vol. 1, 1972, pp 45 - 71.

[86] Funke W The Role Of Adhesion In Corrosion Protection By Organic Coatings. JOCCA, Vol. 68,
No. 9, 1985, pp 229 -232.

[87] Bullett T.R. & Rudram A.T.S. The Coating And The Substrate, JOCCA, Vol. 44, No. 11, 1961, pp
787-815.

[88] Bikerman J.J. The Science of Adhesive Joints 2nd Edition Academic Press, New York and London,
1968, pp 137 – 150.

[89] Walker P. Organo Silanes As Adhesion Promoters For Organic Coatings. J.C.T. Vol. 52, No. 670,
1980, pp 49 - 61.

http://www.jcse.org/volume3/paper5/v3p5.php?commentmode=2 6/20/2013
JCSE Volume 3 Paper 5 Page 26 of 27

[90] Zorll U. New Insights Into The Process Of Adhesion Measurement And The Interactions At
Polymer/Substrate Interfaces, JOCCA, Vol. 7, 1983, pp 1983.

[91] de Vries J.E. Holubka & Dickie R.A. X-ray Photoelectron Spectroscopy Study of Corrosion-
Induced Paint Adhesion Loss on Conversion-Coated Steel, Ind. Eng. Chem. Prod. Res. Dev., Vol. 22,
1983, pp 256-261.

[92] Plato Timaeus

[93] Leidheiser H. Jr. and Deck P.D. Chemistry of the Metal-Polymer Interfacial Region Science Vol.
241, No 2, 1988, pp 1176-1181.

[94] Leidheiser H. Funke W. Water Disbondment and Wet Adhesion of Organic Coatings on Metals: A
Review and Interpretation. , JOCCA, Vol. 70, No. 5, 1987, pp 121 - 132.

[95] Downey S.J. and Devereux O.F The Use Of Imperance Spectroscopy In Evaluating Moisture-
Caused Failure Of Adhesives And Paints, Corrosion, Vol. 45, No. 8, 1989, pp 675-684.

[96] Crossen J.D. Sykes J.M. Knauss D. Briggs G.A.D. & Lomas J.P. The influence of water on the
coating-metal interface, adhesion measurements & scanning acoustic microscopy. ,
*****CONFERENCE*************, Vol. ***, No. ***, 1994, pp 274-283.

[97] Nguyen T. Bentz D. & Byrd E. A Study Of Water At The Organic Coating/Substrate Interface ,
J.C.T., Vol. 66, No. 834, 1994, pp 39 - 50.

[98] Gosselin C.A. Effect Of Surface Preparation On The Durability Of Structural Adhesive Bonds,
Polymeric Materials For Corrosion Control. ACS symposium 322 1986, pp 180-193.

[99] Funke W. How Organic Coating Systems Protect Against Corrosion. Polymeric Materials For
Corrosion Control. ACS symposium 322 1986, pp 223-228.

[100] Troyk P.R. Watson M.J. & Poyezdala J.J. Humidity Testing Of Silicon Polymers For Corrosion
Control Of Implanted Medical Electronic Protheses. Polymeric Materials For Corrosion Control. ACS
symposium 322, 1986, pp 299-313.

[101] Gowers, K R & Scantlebury, J D An Electrochemical Investigation of the Effect of the Adhesion of
a Lacquer Coating on the Underfilm Corrosion. JOCCA. Vol. 4, No. 71, 1988, pp 114-121.

[102] Marsh J. Scantlebury J.D. & Lyon S.B. Silinisation – A Dangerous Surface Treatment for Steel?
CONFERENCE 1994 pp 243 – 253.

[103] Jin, X H Gowers, K R & Scantlebury, J D The Effect of Environmental Conditions on the
Adhesion of Paints to Metals. JOCCA. Vol. 3, No. 71, 1988, pp 78-81.

[104] Arslanov V.V. & Funke W. The Effect Of Water On The Adhesion Of Organic Coatings On
Aluminium, Prog Org Coat, Vol. 15, 1988, pp 355-363.

[105] Arslanov V.V. & Funke W. Improvement of the resistance to water of an adhesive joint between
polymers and aluminium by using thin adhesion layers, Prog Org Coat, Vol. 15, No. 4, 1988, pp 365-372.

[106] Funke W. Improvement Of Wet Adhesion Of Organic Coatings By Thin Adhesion Layer, Surface

http://www.jcse.org/volume3/paper5/v3p5.php?commentmode=2 6/20/2013
JCSE Volume 3 Paper 5 Page 27 of 27

Phenomena and Latexes in Waterborne Coatings and Printing Technologies, 1995, pp 115-122.

[107] Funke W. Thin-Layer Technology In Organic Coatings, Prog Org Coat, Vol. 28, No. 1996, pp 3-7.

[108] Kumins C.A. Physical Chemical Models For Organic Protective Coatings. J.C.T., Vol. 52, No.
664, 1980, 1981, pp 39 - 53.

[109] Karyakina M.I. & Kuzmak A.E. Protection By Organic Coatings: Criteria, Testing Methods And
Modelling. Prog. Org. Coat. Vol. 18, 1990, pp 325 - 388.

This paper has been printed from the Journal of Corrosion Science and Engineering, the online journal for
corrosion and its control, ISSN 1466-8858. Freely available on http://www.jcse.org/. Comments can be
added to papers on the online version.

© 2001 University of Manchester, Corrosion and Protection Centre.

http://www.jcse.org/volume3/paper5/v3p5.php?commentmode=2 6/20/2013

You might also like