Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Electrochimica Acta 51 (2006) 4505–4515

Interpretation of EIS data from accelerated exposure of coated metals


based on modeling of coating physical properties
B.R. Hinderliter, S.G. Croll, D.E. Tallman, Q. Su, G.P. Bierwagen ∗
Department of Coatings and Polymeric Materials, North Dakota State University, Fargo, ND 58105, United States
Received 23 June 2005; received in revised form 16 December 2005; accepted 20 December 2005
Available online 6 March 2006

Abstract
Electrochemical studies on the accelerated weathering of corrosion protective coatings systems in this laboratory have yielded time series data
in the low frequency portion of the spectrum that can be fit by a simple exponential decay function in time. A description of how this can be
empirically used for lifetime prediction for coatings has already been presented. Attempts have been made to use other parts of the EIS spectrum for
analyzing coating performance in accelerated exposure, but these have not yet yielded a predictive model for the effects of exposure on the structure
and performance of organic coatings. First, a simple model which considers only loss of coating thickness by UV-induced ablation is applied to
the interpretation of EIS time series results from coatings systems undergoing accelerated weathering. Next, more complete models which include
a diffusion front of electrolyte that moves with either Fickian or Type II diffusion kinetics are applied to the same problem. Incorporating these
models for degradation as a simple RC equivalent circuit representation of a coating gives predictions of EIS performance that have reasonable
agreement with experimental data. Further, using a two layer representation of coating changes due to water intrusion and known effective medium
theories for dielectric and resistivity changes due to water intrusion allow prediction of the changes in EIS spectra seen in immersion or other
accelerated testing of coating films. These results provide further evidence of the power of modeling physical processes in interpreting EIS data
and predicting coating protective lifetimes.
© 2006 Elsevier Ltd. All rights reserved.

Keywords: Electrochemical impedance spectroscopy (EIS); Organic coatings; Modeling; Corrosion protection; Accelerated testing; Lifetime prediction

1. Introduction The ability to predict a coating’s behavior gives confidence


that all relevant and necessary details of the system’s physi-
Electrochemical impedance spectroscopy (EIS) is commonly cal makeup are included. Understanding those specific coating
used to examine and rank the protective ability of corrosion properties that impact the measurement, and their quantitative
protective organic coatings [1–10]. The method, in general, is importance, allows a more scientific design of coatings for
used in a qualitative or semi-quantitative manner, due to large particular applications, improved predictive ability, and better
variances between replicates and the difficulty in producing understanding of coating failure when it occurs. The inability to
identical coating samples. Equivalent circuit models using pas- predict, and thus interpret, in a physically reasonable and quanti-
sive electrical engineering and physics circuit elements have tative way the EIS spectrum during coating degradation assures
commonly been used to help interpret EIS studies of coating that a great deal of work is left to be done, although some work
systems [11–14]. Data analysis procedures use nonlinear, com- has been attempted in using electrochemical methods to model
plex least squares fit of these equivalent circuit models to the EIS cyclic accelerated testing of coatings [18].
data. Apparent deviations of the circuit element fitted responses One of the problems for a coating scientist in using standard
have produced continuing debate over their physical interpreta- EIS data fitting and modeling techniques is that the tools of pre-
tion [15–17]. dicting and fitting the data are based solely on equivalent circuit
elements, and not based explicitly on solid state physics, elec-
trical engineering or metallurgical engineering. Two especially
∗ Corresponding author. Tel.: +1 701 231 8294. informative references that discuss coated metal characteriza-
E-mail address: gordon.bierwagen@nsdu.edu (G.P. Bierwagen). tion by EIS and the modeling of such data by standard circuit

0013-4686/$ – see front matter © 2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.electacta.2005.12.047
4506 B.R. Hinderliter et al. / Electrochimica Acta 51 (2006) 4505–4515

elements are in Refs. [19,20]. It can be very informative to exam-


ine coatings and their changes by EIS using analysis methods
that focus solely on the physical properties of the coating as
described by the simple Randles circuit description of a coating
[21]. An intact coating is frequently described by an equivalent
circuit that consists of a parallel plate capacitor with the elec-
trolyte solution as one electrode, the coating as the dielectric,
and the metal substrate as the second electrode. The Randles
equivalent circuit describes this simple behavior and consists of
a solution resistance in series with a parallel combination of a
capacitor and resistor. The inputs for describing the coating are
its resistive and capacitive properties.
Based on simple continuum theory [22], the capacitance and
resistance of a smooth coating are given as functions of the size Fig. 1. The diffusion of water into the coating changes the coating in the
of the material and intrinsic properties. The capacitance, C, for diffusion-affected zone while the deeper level of the coating is nominally
a parallel plate capacitor is given by unchanged and retains most of its original physical properties (labeled pristine
coating).
ε0 εR A
C= (1)
d The second model is associated with a coating where water
advances by diffusion into the coating. This model presented
where ε0 is the permittivity of free space (8.8542 × in this paper will be called the diffusion-affected model, and
10−12 C2 /J/m), εR the coating dielectric constant, A the area of describes the changes in the coating layer(s) by at least one
the parallel plate capacitor, and d is the coating thickness. The water-swollen layer (see Fig. 1) in which Fickian or Type II dif-
coating resistance per unit area is given by fusion processes control the thickness of the region as a function
ρd of time [27–29]. We will describe the coating changes for the
R= (2) second model by a Randles-like circuit with a water-polymer
A
composite resistance (see Fig. 2) in series with another Randles
where R is the resistance, ρ the electrical resistivity (a mate- circuit representing the unaffected portion of the coating.
rial constant), d the coating thickness, and A is the area. We
will describe and model data based solely on models for R and 2.1. Changes in coating behavior at low frequencies
C in which changes are only in the film thickness, film resis-
tivity or film dielectric constant. These variables will not be The ablation model addresses the impact of the changes in a
treated as fitting parameters for general R and C circuit elements, coating’s effective thickness as a function of weathering assum-
but as coating variables, based on the knowledge developed by ing the coating material is completely removed (no change in
coating and polymer scientists for changes in coatings due to material properties). The diffusion-affected model employs the
well-described processes such as film thickness loss due to UV assumption that the resistance of the water affected coating is
degradation, film swelling, or due to aqueous diffusion. One negligible in the low frequency regime compared to the pristine
can describe multi-layer coatings as Randles circuits in series coating.
to account for gradients in water and ion concentrations. The In previous publications we have noted that a log-linear decay
present model homogenizes the water and coating into a single of coating resistance (equivalent to low frequency impedance
material as the diffusion front advances. modulus) with time is observed in measurements of various
types of coating in accelerated exposure [10,17]. This behavior
2. Impedance analysis based on two models for coating can be modeled as a coating degrading only by surface ablation
performance due to exterior UV-exposure. Additionally, if the resistance of

The models for coating behavior applied in this paper are


based on the Randles circuit, a representation most often used
for a nominally intact coating [23]. Two models are investigated.
The first model is associated with a topcoat over a primer where
changes occur only in thickness by swelling or ablation. The
ablation model describes the effects of changes in a coating’s
effective thickness as a function of weathering, as a system dom-
inated by bulk changes of the coating. Monte Carlo simulations
and experimental measurements confirm that the average thick-
ness (or weight per area) of the topcoat undergoing mainly UV
degradation induced ablation decreases roughly linearly with Fig. 2. The regions of the coating are separated into components that model the
time [24–26]. diffusion-affected zone and pristine coating.
B.R. Hinderliter et al. / Electrochimica Acta 51 (2006) 4505–4515 4507

the coating after water intrudes is very small compared to the resistance and capacitance:
remaining intact coating, and the low frequency response is dom-    
inated by the remaining intact coating, again we can reproduce 1 ρd ε0 εR A 2
log(|Z|) = log(R) − log 1 + ω
the log-linear decay of the low frequency impedance modulus 2 A d
[10,17]. These two proposed models will be tested for the full
1
EIS frequency range to see which model produces the most rea- = log(R) − log(1 + (ωε0 εR ρ)2 ) (9)
sonable EIS spectra. 2
In the following derivation, considering the solution resis- Using the standard expansion ln(1 + x) = x − 1/2x2 + O(x3 ) for
tance (generally small) detracts from the focus on the coating small x, one gets the form for log|Z| given in Eq. (10). (Also one
components, and therefore will be neglected. notes that that the RC factor loses dependence of the size, either
The impedance of a capacitor in parallel with a resistor is area or depth, of the coating.) For low enough frequencies using
1 R(1 − jωCR) only the first term in the expansion:
Z= = (3)  
1
+ jωC 1 + (ωCR)2 ∼ ρd 1
R log(|Z|) = log = (ωε0 εR ρ)2 (10)
√ A 2 ln(10)
where, j = −1, and ω is the angular frequency. The phase
angle, θ, is given by The impedance is a weak function of the geometry at low fre-
quencies and a Bode plot of the impedance can reveal a small
θ = arctan(−ωCR) (4) slope at low frequencies (0.01–0.1 Hz), depending on the resis-
tivity and relative dielectric constant, diminishing at lower fre-
If the magnitude of the impedance is taken, then the log of both quencies when the second term in Eq. (10) becomes negligible.
sides of the equation gives the functional form that is visualized Monte Carlo predictions and/or diffusion front advance may
in a Bode plot: be incorporated into this low frequency approximation. To a
  first order, the loss of coating thickness (or weight per unit
R 2 2 2 1/2
log(|Z|) = log (1 + ω C R ) (5) area) [30,31] is linear with exposure time. Thus, for the abla-
1 + ω 2 C 2 R2 tion film thickness loss at a rate K1 (in units of thickness/time),
or we can write a simple expression for film thickness as function
of exposure time. Additionally, Monte Carlo models and statis-
log(|Z|) = log(R) − 21 log(1 + ω2 R2 C2 ) (6) tical calculations predict the roughness of the coating to vary
with the square root of exposure time in most cases. Then the
2.1.1. Continuum approximation based on homogeneous “average” thickness as a function of exposure time is then given
materials by Eq. (11):
Applying the continuum expression for the capacitance (Eq. √
d(t) = d0 − (K1 t + K2 t) (11)
(1)) and resistance (Eq. (2)) in Eq. (6) gives
      where d(t) is the effective thickness (or the portion of the coat-
ρd 1 ρd ε0 εR A 2 ing that has not been infiltrated by water), d0 the initial coating
log(|Z|) = log − log 1 + ω
A 2 A d thickness, and t is the time. K1 and K2 are the proportionality
  constants. These constants depend on the polymer, and its ero-
ρd 1 sion processes during weathering. K2 has units of (depth lost per
= log − log(1 + (ωρε0 εR )2 ) (7)
A 2 unit time)1/2 .
The diffusion-affected model describes Fickian (K2 ) and non-
which simplifies in the high frequency range to the capacitive
Fickian (Type II) diffusion (K1 ) which may also be considered as
response:
a means of effective thickness decrease [32]. If one assumes that
  the water swollen section of the film has a negligible resistance
ρd
log(|Z|) = log − log(ωρε0 εR ) with respect to the non-swollen “pristine” portion of the coating,
A
  the formulation (time dependence) of Eq. (11) is the same for
d diffusion driven resistivity loss of pristine material in the water
= log − log ω (8)
Aε0 εR swollen layer as it is for the ablation model. In this case, Type II
diffusion which has a constant velocity advances linearly with
Therefore, if the slope on a Bode plot varies from −1 for a rea- time [27,32–34], would be represented by the constant rate (K1 )
sonably flat layer, one of the material properties, the resistivity or term and Fickian diffusion, which advances according to the
the relative dielectric constant, of the material must be a function square root of time, would be represented by the second (K2 )
of the frequency. term. Thus, the functional form of the effective coating thick-
ness remains the same as a function of time, but the definition
2.1.2. Low frequency limit of K1 and K2 in Eq. (11) will be defined in terms of the diffusion
The simplest form for the low frequency limit is that constants and plasticizing rates (Type II) rather than the mate-
(ωRC)2  1, i.e. |Z| ≈ R. Using the same expressions for the rial loss parameters of film ablation. Thus, linear thickness loss
4508 B.R. Hinderliter et al. / Electrochimica Acta 51 (2006) 4505–4515

per unit time or advancement rate of Type II diffusion front is


described by K1 . K2 is the proportionality constant for decrease
in effective thickness of a coating due to roughening of the sur-
face, and is applicable to Fickian diffusion based advance of the
water front in the coating.
The low frequency resistivity of the water infiltrated coating
is assumed negligible compared to the pristine coating, i.e. the
resistance of the infiltrating solution is often 10 or more orders of
magnitude less than the pristine polymer. As long as the pristine
coating thickness remains a significant portion of the overall
coating thickness, the resistance of the water infiltrated layer can
be neglected. Thus, in the low frequency regime the Ablation and
Diffusion models give the same impedance response. When the
diffusion front advance leaves only a small amount of pristine Fig. 3. Experimental, low frequency impedance |Z| data is taken from acceler-
coating unaffected by water intrusion, the resistance of the water ated exposures of aircraft coatings. Used with permission from Ref. [12].
affected material can no longer be neglected. As the diffusion
front consumes most of the pristine coating, the resistance of negative; resulting in accelerated decrease in the impedance as
the water saturated coating needs to be added in series to the higher order terms in Eq. (12) become more significant. This is
resistance of the remaining pristine coating. exactly what we have seen in our studies of long term exposure
Incorporating Eq. (11) for thickness into the estimate of low of coatings, especially in ASTM D5894-96 exposure [17,35].
frequency impedance, Eq. (10), gives: Fig. 3, taken from Ref. [35], shows such experimental results.
   √  Once the resistivity ρ and dielectric constant εR are measured
ρd0 K1 t K2 t
log|Z| = log 1− + using an independent method for both the pristine coating and
A d0 d0 water saturated coating, the only variable left in fitting the model
1 to measured data is the advance rate of the Type II diffusion front
− (ωε0 εR ρ)2 (12) (or rate of ablation). Thus, the model is more physical, can quan-
2 ln(10)
titatively match experimental data and reduces the number of
The last term of Eq. (12) is dependent only on material proper- fitting parameters compared to present methods based entirely
ties, not exposure time. on fitted equivalent circuit components. The one fitting parame-
Expanding the first term in Eq. (12) for short times gives a ter is the water/electrolyte-swollen front advance rate, which is
linear time dependence of impedance for Type II diffusion or itself a physical quantity of the system.
ablation (K2 = 0): Eq. (14) is fit to the data in Fig. 3 and plotted in Figs. 4 and 5
   √  for both Fickian and Type II diffusion. Fig. 3, which shows a
ρd0 1 K1 K2 t
log|Z| = log − t+ plot of the linear expansion calculated from Eq. (14), seems
A ln(10) d0 d0
reasonable over most of the coating’s effective life. However, as
1
− (ωε0 εR ρ)2 (13)
2 ln(10)
or re-ordered as
   
ρd0 1
log|Z| = log − (ωε0 εR ρ)2
A 2 ln(10)
  √ 
1 K1 1 K2 1/2
− t− t (14)
ln(10) d0 ln(10) d0
The slope of the impedance at low frequencies should correlate
with the coating loss (or depth of coating penetration by signifi-
cant water) as a function of time determined by K1 . The intercept
on the plot should be defined by the geometry of the system, coat-
ing thickness, EIS area, and intrinsic dielectric properties of the
coating.
Thus, if one considers only surface ablation, with time con-
stant K1 , or Type II diffusion only, the plot of the log of the
impedance modulus versus exposure time is linear. With pre-
dominantly surface degradation via roughening, one will also
Fig. 4. Experimental low frequency impedance data taken from Fig. 3 for the
see a linear decrease with time of the log of the low frequency High Gloss Topcoat and fit to the low frequency impedance predictions based
impedance after a short transient due to the square root of time on advance of the water diffusion front. The dashed line represents the fitting
dependence. In both cases, the higher order terms are strictly based on Fickian diffusion and the solid line is based on Type II diffusion.
B.R. Hinderliter et al. / Electrochimica Acta 51 (2006) 4505–4515 4509

Fig. 6. The electrical resistivity and relative dielectric constants for the diffusion-
affected zone are modeled in three ways: first as percolation paths which act
as channels to deliver water to the interface and advance the diffusion-affected
zone. The second model assumes a Maxwell type two-phase system where water
fills voids within the coating and coating stresses valve the polymer between
inclusions to result in water movement. The third model combines inclusions,
holding most of the water, with a few channels that connect the water inclusions.
Fig. 5. Experimental low frequency impedance data taken from Fig. 3 for the
Self-Priming Topcoat and fit to the low frequency impedance predictions based
on advance of the water diffusion front. The dashed line represents the fitting interface resistance, the resistance of transport along the inter-
based on Fickian diffusion and the solid line is based on Type II diffusion. face, becomes negligible, the equivalent circuit becomes two
Randles circuits in series. Randles circuits in series model the
the coating begins to fail, the decrease in impedance accelerates Maxwell approximation for spherical water inclusions and are
as modeled in Figs. 4 and 5. The impedance of the initial coating the first asymptote. The second asymptote is when the interface
sets the intercept. The slope is simply the effective loss rate of the has a very large resistance as would be the case of percolation
coating, either by material loss or advance of the water saturation paths of water that are reasonably separated from one another.
front, divided by the full coating thickness. Effective medium theory puts the conducting material in
Thus, a linear loss in film thickness due to coating ablation a sphere of equivalent volume, surrounded by a matrix, usu-
or water saturation front advance from weathering without other ally the insulator. Thus, in the direct application of EMT to a
coating damage, is sufficient to give a log-linear plot of low polymer/water system, insulating material always surrounds the
frequency impedance modulus versus exposure time. The decay conducting sphere. The application of EMT has predominantly
constant of the latter plot can be equated to the effective loss rate been to solid inclusions, such as conducting solid spheres in
of the coating, either by material loss or advance of the water an insulating system, frequently a capacitor dielectric. In the
saturation front, divided by the full coating thickness. This decay evolution of weathering or continuous immersion of coatings,
constant can be used to predict effective coating lifetime if the perhaps small low-density regions within the coating are filled
low frequency impedance modulus of the coating at “failure” is with water. However, in the water transport process the water
specified by the user [36]. must diffuse from the surface to these inclusions and between
inclusions. This makes the Type II diffusion front more like
2.2. Use of effective medium theory (EMT) models of the a continuous saturated polymer with, perhaps, islands of pris-
swollen layer to model the entire EIS frequency spectrum tine coating. In such a case, the saturated region of the polymer
for the diffusion model may be considered the matrix material with enclosed unsatu-
rated polymer as the spherical inclusions. This view requires a
The above discussion represents the low frequency different description of the volume fraction of water and poly-
impedance of the EIS spectrum for both models. In order to mer. One would need to define, identify and measure a fully
simulate the entire EIS spectrum for the polymer coating in the saturated region of the polymer.
diffusion-affected model, the dielectric constant and the resistiv- For either the ablation or the diffusion model to be reason-
ity of the portion of the coating which contains water need to be able, it must not only match the low frequency asymptote (which
quantified. Two ways of looking at the water in saturated poly- they do), but must generate full spectra EIS for a considerable
mers are suggested. It can be viewed as a set of isolated inclu- portion of the coating lifetime. The parameters for the advance
sions that fill with water or interconnected diffusion/percolation of the diffusion front and ablation, generated in our modeling of
paths filled with water, like a conducting web, Fig. 6. Modeling the low frequency EIS must give realistic full EIS spectra for that
the electrical properties of mixtures of materials goes back to time. Ablation alone, however, gives an excessive spread in high
the time of Maxwell and Raleigh [37]. These types of models frequency impedance, if it alone is considered as the cause for the
are now classified under the name effective medium theories and decrease in low impedance decrease. The high frequency (capac-
are, in general, designed to cope with the low volume fraction itive) region of the Bode plot often overlaps for EIS spectra
material as inclusions, not as percolation paths. The equivalent taken at various times during weathering, though ablation clearly
circuit in Fig. 2 incorporates both of these cases as limits. As the decrease the modulus of the impedance at high frequencies. This
4510 B.R. Hinderliter et al. / Electrochimica Acta 51 (2006) 4505–4515

is due to the large decrease in capacitance with film thickness Eq. (20) gives the Maxwell approximation based on spherical
loss. However, if the change in impedance of the coating with inclusions [40]:
time is due to water infiltration, the water containing portion of
water + 2εpolymer + 2φ(εwater − εpolymer )
the polymer coating has an effective dielectric constant that lies polymer εR
εeffective
R = εR R
polymer
R R
between that of the original coating and that of water. This breaks εwater + 2ε − φ(ε water − εpolymer )
R R R R
the coating into two regions. The first, diffusion-affected zone (20)
has electrical properties of the composite water/polymer system
and these properties are likely to vary with water concentration. Effective conductivity can also be calculated by equations hav-
The dielectric constant of the water-affected, swollen coating ing the same form. The conductivity, the inverse of resistivity,
depends on how the water is distributed. This is where one can is inserted into the calculations in place of the relative dielectric
use an EMT to estimate the resistivity and relative dielectric con- constant.
stant. The water almost certainly exists as an inhomogeneous The effect of water within the polymer coating is much greater
distribution through the coating, which is beyond the simple on the resistivity than on the relative dielectric constant of the
EMT approximations [38]. Finite element methods are currently water saturated material. This is due to the relative magnitudes of
being employed to investigate these more detailed distributions, the two intrinsic properties. Resistivity differs by nine orders of
and will be presented at a later time. What follows are EMTs magnitude between water and pristine coating, while dielectric
employed to get a first order estimate of the relative dielectric constant differs by one order of magnitude.
constant and resistivity to be incorporated into the equivalent Estimates for resistivity and relative dielectric constant for
circuit model and verify the reasonableness of the model in pre- the diffusion-affected zone can be made. Water must transport
dicting the entire EIS spectra. evenly across the exposed coating area through the diffusion-
The relative dielectric constant and resistivity of the water affected zone (Zone II) towards the pristine polymer (Zone I)
affected zone need to be calculated in order to estimate interface in Fig. 6. Thus, the water diffusion front is parallel to
the EIS response. We present below, in Eqs. (15)–(20), the the plane of the coating. The volume fraction of water to use
EMTs that have been used to describe composites such as our in the equations above is the volume of absorbed water divided
[water + coating] systems. These are intended to be applicable by volume of the water affected zone, at any particular time.
at inclusion volume fractions less than 10%. These allow a test of the diffusion-affected model. The effective
The equation: medium theories give relative dielectric constants and resistiv-
ities as shown in Table 1. Although these estimations of the
polymer
εeffective
R = εR (1 − φ)3/2 (15) evolution of the coating and assignment of materials properties
are reasonable, more complete analysis is necessary.
where εR is the dielectric constant of the superscripted material
Predictions for entire EIS spectra require knowledge of how
and φ is the volume fraction of the secondary phase (water if the
the water is entrained in the polymer. These early results from
water is assumed to remain a separate and distinct material) has
modeling, employing EMTs, suggest that water tends to cluster
been suggested by Brueggman [37].
at low density locations within the coating, with some additional
One can apply various effective medium theories. If one
percolation paths connecting a portion of these water rich regions
assumes that the water is the spherically included phase and
as shown in Fig. 6.
that the saturated polymer is the continuous phase (neglecting
possible islands of pristine polymer) the following equations
results: 3. Application of EMT and equivalent circuit models
polymer
εeffective
R = εR (1 − φ) + εwater
R φ (16) The resistivity and relative dielectric constants derived for the
This effective medium theory equation is a linear combination of water/coating composite are applied to the equivalent circuits
dielectric constants and is equivalent then to adding capacitors described in the earlier section. The diffusion-affected model is
in parallel: separated into two cases; the first has the water advancing in
percolation paths, while the second has water valving between
1 1−φ φ “Maxwell” inclusions. It is noteworthy that positron annihilation
= polymer + water (17)
εeffective
R εR ε R spectroscopy results are fitted well by assuming that the free vol-
ume is presented as spheres [41]. Emphasis will be placed on
this Eq. (17) is equivalent to adding the capacitance in series
Type II diffusion (K1 parameter) since it is common in polymer
(which implies that the resistivities would be added in parallel).
systems, though it will be seen below that Fickian diffusion
One can also develop a generalization of the above equations as
(K2 parameter) also produces reasonable results. Both cases
polymer n n 1/n described below assume a net advance of the diffusion front
εeffective
R = [(εR ) (1 − φ) + (εwater
R ) φ] (18)
with time and use the water front advance rates predicted by the
with n = 1 giving Eq. (16) and n = −1 generating Eq. (17). fits in Figs. 4 and 5. The coating imbibes water from the exter-
An approximation often used in the analysis of EIS data from nal electrolyte, then the interface between the saturated layer of
coatings was developed by Brasher and Kingsbury [39]: coating and the intact coating moves inward towards the sub-
polymer strate with exposure time, according to Type-II diffusion. In the
log(εeffective
R ) = log(εR )(1 − φ) + log(εwater
R )φ (19) figures that follow, no account is taken of the ablation of the
B.R. Hinderliter et al. / Electrochimica Acta 51 (2006) 4505–4515 4511

Table 1
The range of resistivities and dielectric constants for various EMT approximations
Resistivity ( cm) Dielectric constant

0.05a 0.1a 0.2a 0.05a 0.1a 0.2a

Assuming that the volume fraction is the water phase as inclusions in polymer
Eq. (15) 3.0E + 04 1.0E + 04 3.7E + 03 8.6 9.3 11.0
Eq. (16) 6.6E + 03 3.3E + 03 1.6E + 03 8.4 8.8 9.8
Eq. (17) 1.9E + 11 1.8E + 11 1.6E + 11 11.7 15.4 22.8
Eq. (19) 7.3E + 10 2.6E + 10 3.5E + 09 9.0 10.1 12.7
Bruggeman’s equation 1.7E + 11 1.5E + 11 1.0E + 11 8.9 9.9 12.0
Maxwell’s equation 1.7E + 11 1.5E + 11 1.1E + 11 8.9 10.0 12.3
Assuming that the water effected phase is the matrix and unhydrated polymer the inclusions
Eq. (20) 9.7E + 03 4.8E + 03 2.3E + 03 10.6 13.3 19.0

Resistivity ( cm): water, 330; polymer, 2.00E +11. Dielectric constant: water, 82; polymer, 8.
a Water volume fraction.

coating, or of swelling of the saturated layer, nor is there a gra- For the pristine coating, representative coating properties are
dient in water concentration with depth in the diffusion-affected applied with resistivities approximately 1012  cm and relative
zone. These contributions can be added easily when data are dielectric constant of five to eight. The coating thickness is taken
available to justify that level of detail. Thus, both models below as 50 ␮m and the EIS cell area is assumed to be 5 cm2 . The resis-
assume a coating of fixed thickness, composed of two homoge- tivity and dielectric constant of the water-modified coating are
neous materials as shown in Fig. 1. The first region is a saturated very model dependent, especially with regards to the resistivity,
polymer that grows linearly in thickness. The second region is see Table 1. Future experimentation will provide further under-
the unaffected polymer, which loses thickness linearly as it is standing of coatings and their degradation mechanism.
consumed by the advance of the water diffusion front. The dif- Fig. 7 shows changing EIS spectra calculated by assuming
ference in the two cases below is how they view and account that the saturated coating fails by having conducting electrolyte
for the resistivity and relative dielectric constant of the saturated paths to the interface with the intact coating (the percolation case
polymer. Experimental data which are known to involve water of the diffusion model). This is the mechanism when the inter-
intrusion effects have been obtained in this laboratory [42]. face resistance is assumed infinite, where the electric field below

Fig. 7. Bode modulus and phase angle spectra from model calculations. The equivalent circuit in Fig. 2 is applied with the interface resistance infinite, simulating
percolation paths. The equivalent circuit is then numerically fit to the Extended Life Topcoat measurements of Fig. 3. The diffusion front advance rate is used to predict
EIS spectra for various times: (a) shows the impedance as a function of frequency for various times assuming a 0.2 ␮m/week movement of the diffusion-affected
zone interface with a saturation of 5% by volume of water; (b) shows the same as (a) but with a saturation of 20% by volume water.
4512 B.R. Hinderliter et al. / Electrochimica Acta 51 (2006) 4505–4515

Fig. 8. EIS Bode modulus spectra as predicted by model calculations when the interface resistance is set to 0. Part (a) shows results predicted by “Maxwell” type
averaging of both relative dielectric constant and resistivity. Parts (b)–(d) shows the calculated resistivity of inclusions combined with a few percolation paths (Fig. 6c)
which reduce the resistance () to various levels, listed in figure as R. The small impact on dielectric constant due to a small number of percolation paths is neglected.

the advancing percolation paths cannot communicate with the the likely decreasing resistivity as water penetrates to increase its
electric field along the interface. Thus, resistance drops precip- concentration in the inclusions, Figs. 8 and 9(b)–(d). The results
itously as water infiltrates. Fig. 7 employs the equivalent circuit of this modeling (see Fig. 6) give qualitative agreement with EIS
of Fig. 2 with the interface resistance removed since it is effec- spectral measurements used for the low frequency impedance
tively infinite. It was calculated with Eq. (16), which assumes modulus data.
that the water saturating the coating can be approximated as Fig. 9(a)–(d) shows the Bode phase angle spectra from the
direct channels from the surface across the diffusion-affected same model calculations as Fig. 8. One can see that we have clear
zone. This is the limiting case for the EMT since the resistivity evidence of two time constants in parts (b)–(d) due to the two
of the diffusion-affected zone is minimized. The capacitance of coating regions, the water saturated region and pristine region
the diffusion-affected zone is only slightly affected. Fig. 7 shows (Fig. 1). The presence of two time constants in the literature
the predicted EIS Bode modulus spectra for the ‘Extended Life concerning coated metal impedance spectra has almost always
Topcoat’ assuming water saturation for the diffusion-affected been interpreted with two RC elements, with the second ele-
material occurs at 5% and 20%, respectively. The advance of ment describing a pore resistance and a double layer capacitance
the diffusion-affected zone interface, and thus its volume, is associated with a delamination area of the coating. These calcu-
assumed to be linear with time (Type II diffusion) and it is lations, however, show this behavior may also occur in an intact
assumed to move at a rate of approximately 1% of the total coating (dominated by bulk water transport, not defects) during
film thickness per week. The impact of the higher water at satu- planar infiltration of water with no need to invoke a delamination
ration is to increase the spread in the high frequency regime of of the coating from the metal surface.
the Bode plot. The matching predicted Bode phase angle spectra The Bode and phase angle spectra for the same set of data
for this situation are also given in Fig. 7. used for the low frequency impedance fit for Fig. 4, taking only
If the interface resistance is zero, then we are assuming essen- the 10th weeks data are given in Fig. 10. The data is in reasonable
tially that the coating is in two distinct layers, each of which has agreement given that the water/diffusion advance rate is taken
a finite resistance. Figs. 8 and 9(a) are constructed assuming from the low frequency data and the water affected material is
the water moves between discontinuous voids to the interface, homogenized via effective media theory approximations. The
not maintaining a continuous path. Thus, the resistivity of the Bode plot in Fig. 10 matches the trends in Fig. 8c, though the
diffusion-affected zone stays high. The Maxwell equation (Eq. low frequency plateau is much less flat in the measured data.
(20)) is used to calculate an effective dielectric, 10.0 (that of The measured phase angle spectra of Fig. 10 also show behavior
water is assumed to be 82), and resistivity, 1.5 × 1011  cm, of consistent with the model (Fig. 9b and c). The measured phase
the saturated coating. The resistivity is then decreased, to model angle has much broader peaks than the model, likely due to the
B.R. Hinderliter et al. / Electrochimica Acta 51 (2006) 4505–4515 4513

Fig. 9. The EIS Bode phase angle spectra for the same conditions as given in Fig. 8. Spectrum (a) gives results predicted by “Maxwell” type averaging of both relative
dielectric constant and resistivity. (b)–(d) shows the combined effect of inclusions with a few percolation paths (Fig. 6c) which reduce the resistance to various levels,
listed in figure as R. The small impact on dielectric constant due to a small number of percolation paths is neglected.

homogenizing of the water in the model instead of having a water saturated polymer material is assumed here to be homogeneous
concentration gradient. Another example of published EIS data with respect to water volume fraction. It is reasonable to assume
presents Bode and phase angle plots in Fig. 5 for a thick epoxy the water volume fraction varies with depth within the diffusion-
polyamide-coated steel and phase angle in Fig. 2 for a thin epoxy affected zone. This gradient in the water concentration and thus
polyamide-coated steel [34] and shows the same trends as the the resistivity and effective relative dielectric constant will be
composite system in the present paper in Figs. 8 and 9. incorporated into our model as better experimental data war-
Additional refinements will be added to the models as rants further detail.
justified by experimental data. The diffusion-affected, water- In summary, an equivalent circuit model with two physical
asymptotes is proposed to simulate a coating possessing a diffu-
sion front progressing into a coat (see Fig. 1). The first asymptote
(when the interface resistance is infinite and models percola-
tion paths that advance into the topcoat) is two serial capacitors
in parallel with two serial resistances (one is negligible). The
second asymptotic equivalent circuit model (when the interface
resistance is negligible and describes a nearly constant elec-
tric field at the water advance interface as occurs in Maxwell’s
approximation of spherical water inclusions) is a Randles cir-
cuit that describes the water-saturated coating in series with a
Randles circuit that describes the remaining pristine coating. In
both cases the volume of the water saturated coating increases
and the volume of the pristine coating decreases as the water
penetrates deeper into the coating. One can easily see that if
failure is identified with the saturated “front” reaching the metal
substrate, a lifetime prediction based on this model can easily
be made.
The transport of water with entrained ions to the substrate
Fig. 10. Bode and phase angle plots for the High Gloss (also used in Fig. 4) metal should be restricted for limiting corrosion initiation and
topcoat is plotted every 10 weeks. growth. A topcoat is the major barrier to water infiltration, and
4514 B.R. Hinderliter et al. / Electrochimica Acta 51 (2006) 4505–4515

typifies the subject of this model. Whereas topcoats are very This modeling that we describe in this paper is important to
resistive and are designed to have low water permeability, primer coatings characterization by EIS as well as to the use of EIS
layers are often designed to easily transport water and inhibitors experimental data in the prediction of real-world coating per-
to the sites of corrosion. Modeling such a primer layer contin- formance. The modeling that we have done differs significantly
ues to be a part of our research. The design and quantification of from the standard emphasis on “modeling” that uses equivalent
the barrier properties of the topcoat, separate from the primer, circuit elements such as resistors, capacitors, constant phases
remains an important aspect of corrosion prevention and coating elements, etc., and focuses on the modeling of physical property
system lifetime prediction. Additional complications associated parameters of such elements, in this case resistivity and dielec-
with the cyclic wet–dry accelerated weathering make this model tric constant that can then be used in resistor and capacitance
limited to a net advance of the water front. Finite element models elements. This allows us to bring into EIS studies the work done
are being developed to predict the water distribution through a in coatings science and polymer science on changes in resistiv-
coating. The overall system impedance response to a frequency ity and dielectric behavior shown by polymers and pigmented
dependent applied voltage applied to this water distribution organic coatings upon exposure to UV and/or water/electrolyte.
within a coating is simulated via finite element modeling as The results we have shown give illustration to the value of such
a part of our continued research to simulate EIS response. modeling in interpreting and fitting experimental EIS data. Such
quantitative predictive capabilities will be crucial to the success-
4. Summary and conclusions ful rapid development and deployment of new coatings materials
for corrosion protection.
Using a simple linear thickness loss with time from UV expo-
sure or loss of resistivity due to the advance of a water infiltration
front and inserting this into a simple Randles circuit description Acknowledgements
for EIS measurements, one can predict a linear decrease with
time of the log of the low frequency impedance (equivalent to a This work was performed with the support of the Air Force
linear loss in the log of the film resistance with time). This agrees Office of Scientific Research (AFOSR) under Grant F49620-04-
qualitatively with data on weathering and accelerated exposure 1-0398. The earlier work of this cited within the text and below
of coating systems. Further, if we model the change in dielectric had various sponsors, the primary being the Office of Naval
properties as well as resistivity of coating films due to water Research (ONR) and AFOSR.
penetration in immersion or exposure to electrolyte using exist-
ing effective medium theories, we can predict and emulate many References
of the qualitative features of the EIS spectra in the range from
low frequencies to 105 Hz that are seen in experimental systems, [1] G. Bierwagen, L. He, D. Tallman, Time–temperature effects in polymer
including both good and poor performance corrosion protective coatings for corrosion protection as analyzed by ELS, Macromolecular
coatings. This is done using a two layer model for the coating Symposia, vol. 187, Presented at FATIPEC XXVI, Dresden, Germany,
September 9–11, 2002. Wiley-VCH, Weinheim, Germany, 2002, pp.
in exposure based on a water-saturated layer moving into the
909–918.
film following Type II diffusion over a pristine coating layer [2] J.M. Hu, J.Q. Zhang, C.N. Cao, Determination of water uptake and
unaffected by water intrusion. In this case we insert the model diffusion of Cl− ion in epoxy primer on aluminum alloys in NaCl solu-
calculations into an equivalent circuit description using two Ran- tion by electrochemical impedance spectroscopy, Prog. Organic Coat. 46
dles circuits in series to describe the two layers. This leads to (2003) 273–279.
[3] S. Duval, M. Keddam, M. Sfaira, A. Srhiri, H. Takenouti, Electrochem-
two time constants, but with an interpretation different from the
ical impedance spectroscopy of epoxy-vinyl coating in aqueous medium
conventional reasoning that the only reason for a second time analyzed by dipolar relaxation of polymer, J. Electrochem. Soc. 149
constant is delamination. The physical parameter predictions (2002) BS20-9.
that are based on film thickness loss or dielectric and resistiv- [4] J.H.W. de Wit, Inorganic and organic coatings, in: P. Marcus, J. Qdar
ity changes due to water intrusion use experimental values for (Eds.), Corrosion Mechanisms in Theory and Practice, Marcel Dekker,
film loss or dielectric constant values from several coating sys- New York, 1995, pp. 581–627 (Chapter 16).
[5] W. Funke, Corrosion tests for organic coatings—a review of their use-
tems as their starting point. This leads one to believe that more fulness and limitations, J. Oil Chem. Assoc. 62 (1979) 63–67.
complete quantitative descriptions of many experiments might [6] H. Leidheiser Jr., Electrical and electrochemical measurements as pre-
be predicted better by more realistic values of physical parame- dictors of corrosion at the metal organic coating interface, Prog. Organic
ters for coatings systems estimated by other measurements. This Coat. 7 (1979) 79–104.
will allow us to develop a more fundamental description of the [7] W.S. Tait, A discussion of the reliability of electrochemical impedance
spectroscopy data from coated metals div of polymeric materials: sci.
changes undergone by coatings to both accelerated lab exposure eng. preprints, Vol. Amer. Chem. Soc. National Meeting, Denver, April
as well as field exposure, allowing, perhaps, true lifetime predic- 1993, p. 101.
tions for coatings in actual use. Further, the model can provide [8] J.R. Macdonald (Ed.), Impedance Spectroscopy, Wiley-Interscience, New
formulation guidelines by allowing one to identify key physi- York, D.D. Macdonald and M.C.H. McKubrie, 1987, p. 301 (Chapter
cal parameters of the coating which control its performance in 4.3).
[9] G. Bierwagen, D. Tallman, J. Li, L. He, EIS studies of coated metals
exposure. As one might well expect, even this simple modeling in accelerated exposure, in: Proceedings of the Fifth International Sym-
has shown that water “solubility” in the coating on exposure is posium on Electrochemical Impedance Spectroscopy, E1S 2001, June
a key parameter for controlling coating performance. 2001 Marilleva, Italy Prog. Organic Coat. 46 (2003) 148–157.
B.R. Hinderliter et al. / Electrochimica Acta 51 (2006) 4505–4515 4515

[10] J. Kittel, N. Celati, M. Keddam, H. Takenouti, Influence of the coating- [25] S.G. Croll, A.D. Skaja, Spectroscopic absorption and effective dosage
substrate interactions on the corrosion protection: characterisation by in accelerated weathering of a polyester-urethane coating, J. Mater. Sci.
impedance spectroscopy of the inner and outer parts of a coating, Prog. 37 (2002) 4889–4900.
Organic Coat. 46 (2003) 135–147. [26] G. Wypych, Handbook of Material Weathering, second ed., Chem Tec
[11] I.D. Raistrick, J.R. MacDonald, D.R. Franceschetti, The electrical Publishing, 1995.
analogs of physical and chemical process, in: J.R. MacDonald (Ed.), [27] J. Crauk, The Mathematics of Diffusion, second ed., Oxford Science
Impedance Spectroscopy, John Wiley & Sons, New York, 2003, pp. Publications, 1975 (Chapter 11).
27–132. [28] A.S. Argon, R.E. Cohen, A.C. Patel, A mechanistic model of case II
[12] F. Mansfeld, H. Shih, H. Greene, C.M. Tsai, Analysis of EIS data for diffusion of a diluent into a glassy polymer, Comput. Theor. Polymer
common corrosion process, in: J. Scully, D.C. Silverman, M. Kendig Sci. 9 (1999) 339–352.
(Eds.), Electrochemical Impedance: Analysis and interpretation, ASTM [29] A. Friedman, G. Rossi, Phenomenological continuum equations to
STP 1188, ASTM, Philadelphia, PA, 1993, p. 37. describe case II diffusion in polymeric materials, Macromolecules 30
[13] B. Normand, H. Takenouti, M. Keddam, H. Liao, G. Monteil, C. Cod- (1997) 153–154.
det, Electrochemical impedance spectroscopy and dielectric properties of [30] E. Hoffman, A. Saracz, Weathering of paint films: IV. Influence of radi-
polymer: application to PEEK thermally sprayed coating, Electrochim. ation intensity on chalking of latex paints, J.O.C.C.A. 55 (1972) 101–
Acta 49 (2004) 2981–2986. 113.
[14] J. Vogelsang, G. Strum, New interpretation of electrochemical data [31] S. Hollande, J.-L. Laurent, Degradation process of an industrial ther-
obtained from organic barrier coatings, Electrochim. Acta 46 (2001) moplastic elastomer polyurethane-coated fabric in artificial weathering
3817–3826. conditions, J. Appl. Polym. Sci. 73 (1999) 2525–2534.
[15] G.W. Walter, A review of impedance plot methods used for corrosion [32] A.S. Argon, R.E. Cohen, A.C. Patel, A mechanistic model of case
performance analysis of painted metals, Corros. Sci. 26 (1986) 681–703. II diffusion of a diluent into a glassy polymer, Polymer 40 (1999)
[16] G.P. Bierwagen, J. Li, N. Davis, D.E. Tallman, Thickness dependence 6991.
of electrochemical properties of organic coatings, in: Proceedings of the [33] J.R. Scully, Electrochemical impedance of organic-coated steel: corre-
Fifth Nuernberg Congress, vol. 1, Nuemberg, Germany, April 12–14, lation of impedance parameters with long-term coating deterioration, J.
1999. Vincentz Verlag, Hannover, Germany, p. 315. Electrochem. Soc. 136 (4) (1989) 979–990.
[17] G. Bierwagen, J. Li, L. He, D. Tallman, Fundamentals of the measure- [34] M.M. Wind, H.J.W. Lenderink, A capacitance study of pseudo-Fickian
ment of corrosion protection and the prediction of its lifetime in coating, diffusion in glassy polymer coatings, Prog. Organic Coat. 28 (1996)
in: J. Martin, D. Ram (Ed.), Proceedings of the Second International 239–250.
Symposium on Service Life Prediction Methodology and Metrologies, [35] G. Bierwagen, D. Tallman, J. Li, L. He, EIS studies of coated metals
Monterey, CA, November 14–17, 1999. ACS Symposium Series #805, in accelerated exposure, Prog. Organic Coat. 46 (2003) 148–157.
ACS Books, Washington, DC, 2001, pp. 316–350 (Chapter 14). [36] C. Schillcr, W. Strumz, The evaluation of experimental dielectric data
[18] J.H. Park, G.D. Lee, H. Ooshige, A. Nishikata, T. Tsuru, Monitoring of barrier coatings by means of different models, Electrochim. Acta 46
of water uptake in organic coatings under cyclic wet–dry condition, (2001) 3619–3625.
Corrosion Sci. 45 (2003) 1881–1884. [37] R. Meredith, C.W. Tobias, Conduction in heterogeneous systems, in:
[19] U. Rammelt, O. Reinhard, Application of electrochemical impedance C.W. Tobias, P. Delehay (Eds.), Advances in Electrochemistry and Elec-
spectroscopy (EIS) for characterizing the corrosion-protective perfor- trochemical Engineering, vol.2, Wiley-Interscience, NY, 1962, pp. 15–47
mance of organic coatings on metals, Prog. Organic Coat. 21 (1992) (Chapter II).
205–226. [38] G.K. van der Wel, O.C.G. Adan, Moisture in organic coatings—a review,
[20] U. Hammelt, G. Reinhard, Characterization of active pigments in damage Prog. Organic Coat. 37 (1999) 1–14.
of organic coatings on steel by means of electrochemical impedance [39] D.M. Brasher, A.H. Kingsbury, J. Appl. Chem. 4 (1954) 62.
spectroscopy, Prog. Organic Coat. 24 (1994) 309–322. [40] R.E. De La Rue, C.W. Tobias, On the conductivity of dispersions, J.
[21] See, for example, G.P. Bimagen, Calculation of noise resistance from Electrochem. Soc. 106 (1959).
simultaneous electrochemical voltage and current noise data, J. Elec- [41] A.J. Hill, Positron annihilation lifetime spectroscopy to probe free-
trochem. Soc. 141 (1994) L155–L157. volume effects in high temperature-polymers and composites, in: M.R.
[22] J.R. Macdonald (Ed.), Impedance Spectroscopy, Wiley-Interscience, New Tant, J.W. Connell, H.L.N. McManus (Eds.), High-Temperature Proper-
York, 1987. ties and Application of Polymeric Materials of ACS Symposium Series,
[23] G.P. Bierwagen, Calculation of noise resistance from simultaneous elec- vol. 603, 1995 (Chapter 5).
trochemical voltage and current noise data, J. Electrochem. Soc. 141 [42] G.P. Bierwagen, J. Li, L. He, L. Ellingson, D.E. Tallman, Consider-
(1994) L155–L195. ation of a new accelerated evaluation method for coating corrosion
[24] B.R. Hinderliter, S.G. Croll, Monte Carlo approach to estimating the resistance—thermal cycling testing, Prog. Organic Coat. 39 (2000)
photo-degradation of polymer coatings, J. Coat. Tech. 2 (2005) 483. 67–78.

You might also like