Sunteți pe pagina 1din 17

Corrosion Science 45 (2003) 1257–1273

www.elsevier.com/locate/corsci

An electrochemical impedance spectroscopy


study of the corrosion behaviour of PVD
coated steels in 0.5 N NaCl aqueous solution:
Part II. EIS interpretation of corrosion behaviour
a,b,*
C. Liu , Q. Bi a, A. Leyland a, A. Matthews a

a
Research Centre in Surface Engineering, Hull University, Hull HU6 7RX, UK
b
Wolfson School of Mechanical and Manufacturing Engineering, Loughborough University,
Loughborough LE11 3TU, UK
Received 22 April 2002; accepted 21 September 2002

Abstract

In Part I, of this work the equivalent circuits for electrochemical impedance spectroscopy
(EIS) modelling of PVD coated steels in 0.5 N NaCl solution were established. In this paper,
Part II, the EIS spectra of such coated systems are modelled using the equivalent circuits. The
circuit parameters obtained are correlated with the dielectric characteristics, and microstruc-
ture of steels and PVD hard coatings. Coating porosity and localised corrosion with exposure
time have also been determined using the corrosion potential difference (DEcorr ) between mild
steel and PVD coatings and polarisation resistance Rp , which was obtained through EIS
modelling using equivalent circuits. In addition, diffusion rates of the reactants (e.g. oxygen)
through ÔpermeableÕ defects (e.g. pores) are studied by introducing the diffusion impedances W
and O in EIS modelling. It has been found that the usage of impedances W and O is closely
related to the crystallite features of PVD coatings. Warburg impedance (W ) is most suitable
for columnar crystallites, while the co-tangent-hyperbolic diffusion impedance (O) is best for
the equiaxed crystallite structure. Finally, visual inspection, SEM examination, and the
scanning reference electrode technique were employed to observe the corrosion progress of
PVD coated steels with immersion time, in order to validate the EIS interpretation.
Ó 2002 Elsevier Science Ltd. All rights reserved.

*
Corresponding author. Address: Wolfson School of Mechanical and Manufacturing Engineering,
Loughborough University, Loughborough LE11 3TU, UK. Tel.: +44-1509-227678; fax: +44-1509-227648/
267725.
E-mail address: c.liu@lboro.ac.uk (C. Liu).

0010-938X/03/$ - see front matter Ó 2002 Elsevier Science Ltd. All rights reserved.
PII: S 0 0 1 0 - 9 3 8 X ( 0 2 ) 0 0 2 1 4 - 7
1258 C. Liu et al. / Corrosion Science 45 (2003) 1257–1273

1. Introduction

Electrochemical corrosion processes are complex, and are influenced by many


variables with regard to electrolyte, materials and interfacial geometry, etc. Cer-
tainly, mass migration also plays an important role in electrochemical corrosion.
Electrochemical impedance spectroscopy (EIS) has been increasingly employed by a
number of investigators to study corrosion behaviour of materials [1–18]. However,
modelling and interpretation of EIS spectra with an equivalent circuit are always
necessary for the extraction of useful information. For PVD ceramic-coated steels in
0.5 N NaCl solution, galvanic corrosion is usually initiated at pores, where the
difference of electric potential between the coatings and steels is established. EIS
spectra evolution with exposure time reflects corrosion progression of the coated
systems. Modelling and interpretation of EIS data is difficult and tedious, but es-
sential to the EIS study on electrochemical corrosion. Encouragingly, commercially
available instrumentation and computing software have made EIS modelling and
interpretation more efficient and reliable.
In this Part II of the paper, the equivalent circuits and the correspondent CDCs
established in an associated paper [17], were used to carry out the fitting of the ex-
perimental EIS spectra after different times of exposure to 0.5 N NaCl solution. The
parameters of Y0  Qc , Y0  Qs , Rpore , Rp , Y0  W , Y0  O, B, which were used to
define the elements, have been determined through the EIS modelling. In particular,
the variation with the time of immersion is further explored and correlated with the
corrosion behaviour of PVD TiN (or CrN) coated steels. The porosity of PVD
coatings, and the localised corrosion with exposure time were also investigated in
terms of the variation of the polarisation resistance Rp . Finally, visual inspection,
SEM, and the scanning reference electrode technique (SRET) were also employed to
physically examine corrosion behaviour to validate the interpretation of EIS spectra.

2. Experimental details

2.1. PAPVD of Ti–N and Cr–N coatings

The flat substrate samples of mild steel of £30  5 mm, and AISI 316L stainless
steel of £50  3 mm were ground with a decreasing grit size of SiC paper, and finally
polished with 14 lm liquid diamond suspension. The composition of steels used in
this study and the details of the deposition processes were given in Part I [17].

2.2. Electrochemical impedance spectroscopy

AC impedance spectra were derived using a Solartron 1260 impedance gain-phase


analyzer with a Solartron 1286 electrochemical interface, while the PVD coated steels
are exposed to 0.5 N NaCl solution. A 3-electrode potentiostatic regime mode was
utilized with a saturated calomel reference electrode (SCE) as reference electrode.
The EIS spectra were collected at various intervals within the exposure of one week.
C. Liu et al. / Corrosion Science 45 (2003) 1257–1273 1259

Finally, the acquired spectra were analyzed and modelled with the EQUIVCRT
program, using the procedures described in the Part I [17].

2.3. Grazing incident angle XRD and SEM

A Seimens D5000 X-ray diffractometer with thin film attachment was employed to
study the crystallographic characteristics of the PVD coatings. A grazing incidence
angle diffraction system using a thin film attachment is particularly useful to identify
the phase composition and structure of PVD coatings. A continuous scan mode at
scan step size of 0.02° (2h) was used to collect XRD spectra at glancing angles of 1°,
2°, 5°, 10° and 20°, using CuKa radiation under 40 kV voltage and 40 mA current.
A Cambridge Instruments ÔStereoscanÕ 200 scanning electron microscopy (SEM)
was also employed to investigate the surface morphology and microstructure of PVD
coatings. The morphology (e.g. defects), coverage and uniformity of the coatings can
be directly examined. However, to view the cross-sectional microstructure, fractured
cross-sections of the coated steels were prepared as follows: (1) Cut with a handsaw
from the uncoated side until it is near the substrate/coating interface (1 mm); (2)
Break the specimen through the pre-cut location. The samples can then be readily
examined through SEM.

2.4. Scanning reference electrode technique

SRET is a newly emerging electrochemical technique allowing the surface electric


potential to be imaged in relation to the reference electrode potential. It provides
useful information about the active pits (localised corrosion), which are usually
surrounded by a large passive area [18]. In this study, A SRET apparatus (SP100
model, of EG&G Instruments) was employed to monitor the potential on the surface
of PVD coated mild steel during exposure to a diluted (104 N) NaCl solution. Test
coupons of the single layer (2–3 lm) thin TiN (or CrN) coated mild steel mounted in
a shallow glass tray immersed in the solution are integrated with the measuring
systems controlled by the EG&G computer software. With fine lead screw threads a
positional resolution of 0.5 lm can be achieved within a scan area of 1 cm2 . At the
end of scanning images were produced to indicate potential variation between
electropositive and electronegative regions.

3. Results and discussion

3.1. Composition and microstructure of PVD hard coatings

PVD ceramic (e.g. TiN, CrN) coatings usually consist of fine and dense crystal-
lites, the composition of the coatings varies depending upon the deposition condi-
tions. The bombardment due to applying a negative bias during deposition produces
a number of intrinsic defects, such as phase and grain boundaries and dislocations.
However, the permeable defects are more detrimental to the corrosion performance,
1260 C. Liu et al. / Corrosion Science 45 (2003) 1257–1273

Fig. 1. SEM micrographs of PVD hard coating of (a) overall surface morphology; (b) an inclusion (size:
8 lm); (c) a pinhole (size: 3 lm).

as they provide direct paths to allow the corrosive electrolyte to access the steel
substrate [17,19–21]. Fig. 1a shows the SEM micrograph of a typical PVD ceramic-
coated surface, from which two types of the observed permeable defects are shown in
Fig. 1b and c. Fig. 1b shows an inclusion of 8 lm in diameter (size of inclusions
ranges from a few to tens of microns), around which the crevice is likely to be
permeable. Fig. 1c shows another common defect in PVD ceramic coatings, a pin-
hole. As Sonobe et al. [21] reported, the size of most pinholes in PVD coatings are
below 5 lm in diameter, a few are above 20 lm, which shows a good agreement with
the observation in this work. The cross-sectional SEM micrographs of TiN and CrN
coatings, shown in Fig. 2a and b, indicate that TiN coatings are composed of fine
columnar crystallites, which are susceptible to a high density of defects during
growth. In contrast, CrN coatings containing the finer, denser equiaxed crystallites
are expected to be more corrosion resistant.

Fig. 2. SEM micrographs of fractured cross-section of (a) TiN coating with columnar crystallite structure,
and (b) CrN with equiaxed crystallite structure.
C. Liu et al. / Corrosion Science 45 (2003) 1257–1273 1261

Fig. 3. Grazing incident XRD patterns for PVD TiN coated mild steel at glancing angle of 1°, 2°, 5°, 10°,
20° with CuKa radiation.

According to the grazing incident angle XRD spectra in Fig. 3, the multi-layered
TiN coatings mainly consist of the TiN phase plus a small amount of a-Ti phase,
which is understood to be the interlayer from each stage of deposition. However, the
multi-layered CrN coating is a mixture of three phases of Cr, Cr2 N and CrN from
the XRD spectra presented in Fig. 4. The e-Cr2 N phase is the hardest, and a-Cr is
softest among the phases of the Cr–N system according to Engel [22]. The combi-
nation of fine microcrystallites of e-Cr2 N and a-Cr phases was expected to give
optimised properties in terms of wear and corrosion resistance [22,23]. Furthermore,
Engel et al. also found that formation of Cr2 N phase could stimulate growth of a
glassy structure, which may be beneficial for corrosion resistance [22].

3.2. EIS spectra and modelling

Experimental impedance spectra, which were collected from PVD coated systems
after 0.5, 2 h, 1.5 and 7 d exposure to the 0.5 N NaCl solution, are presented in the
Bode plots in Figs. 5 and 6. For TiN/MS and CrN/MS systems (in Fig. 5), two time
constants are observable. The one appearing at high x represents the dielectric
characteristic of the PVD coating, while the one at low x corresponds to the mild
steel in the pores. As shown by the broken lines in the plot of a–x presented in Fig.
5b and d, the positions of two peaks for TiN remain unchanging with the immersion
time, indicating a stable property of the coatings. However, the peaks for CrN shift
toward low x, as the time increases, suggesting that the dielectric property of CrN
coatings may be effected due to the exposure. In addition, the decrease of peak height
1262 C. Liu et al. / Corrosion Science 45 (2003) 1257–1273

Fig. 4. Grazing incident XRD patterns for PVD CrN coated mild steel at glancing angle of 1°, 2°, 5°, 10°
with CuKa radiation.

(a) (b)

(c) (d)

Fig. 5. Bode plots of EIS spectra of TiN/MS (a,b) and CrN/MS (c,d) systems exposed to 0.5 N NaCl
solution for 0.5, 2 h, 1.5, and 7 d; solid lines are the fitting curves with equivalent circuits. (a) and (c) are
plots of jZj vs. x; (b) and (d) are plots of a vs. x.

with exposure time indicates that the response becomes less capacitive; indeed, as
immersion time increases, the pores permit the solution to penetrate to the steel
C. Liu et al. / Corrosion Science 45 (2003) 1257–1273 1263

Fig. 6. Bode plots of EIS spectra of TiN/SS (a,b) and CrN/SS (c,d) systems exposed to 0.5 N NaCl so-
lution for 0.5 h, 0.5, 1.5, and 7 d; solid lines are the fitting curves with equivalent circuits. (a) and (c) are
plots of jZj vs. x; (b) and (d) are plots of a vs. x.

substrate, causing localised corrosion. Thus, current flow through the defects (i.e.
localised corrosion) becomes significant.
Fig. 6 presents the Bode plots of EIS spectra from TiN, and CrN coated AISI 316L
stainless steel. From a–x plot the maximum phase angles are close to p=2, repre-
senting a pure capacitive response throughout 7 d of exposure. It is expected that the
response from the AISI 316L steel at pores is also capacitive as it is highly passivated.
The passive layer is normally an inert thin oxide film with a low conductivity [9],
which results in a capacitive response which is similar to that from the PVD ceramic
coatings, such that only single time constant is observed through 7 d of exposure.
Using the CDCs of the equivalent circuits proposed in the previous paper (Part I)
[17], the EIS spectra presented in Figs. 5 and 6 could be fitted using EQUIVCRT
software. The fitting curves are presented in Figs. 5 and 6 as the solid lines for each
spectrum. Therefore, the parameters obtained through modelling together with the
corrosion potentials (Ecorr ) as a function of immersion time are summarized in Tables
1 and 2. Clearly, the corrosion potential (Ecorr ) of PVD coated AISI 316L steel was
generally more positive than that of PVD coated mild steel. The Ecorr of CrN/MS is
50 mV more positive than TiN/MS, which suggests that the durability of the TiN
coatings is markedly reduced due to corrosion (via through-coating porosity) at the
coating/substrate interface [22].

3.3. Corrosion behaviour of PVD ceramic-coated steels

In EIS modelling, the elements Rpore and Qc represent the coating properties, so
that parameters such as Rpore and Y0  Qc (equivalent to Cc ) are often used to
1264
C. Liu et al. / Corrosion Science 45 (2003) 1257–1273
Table 1
EIS data simulations for MS and SS with PVD TiN coatings in 0.5 N NaCl solution
PVD Exposed Ecorr Re Y0  Qc n  Qc Rpore a Y0  Qs n  Qs Rs Y0  W
p
systems time (mV, SCE) (X cm2 ) (l/cm2 sn X) (kX cm2 ) (l/cm2 sn X) (kX cm2 ) (cm2 s X1 )
TiN/MS 0.5 h 31 4.6 52.6 0.91 13 16.4 0.69 331.5
2h )250 5.6 97 0.85 7.9 28.2 0.99 10.0 0.7
1.5 d )590 5.1 268 0.80 1.2 99.1 0.64 3.9 1.6
7d )650 4.6 315 0.78 0.1 485 0.78 6.0 37.5
TiN/SS 0.5 h 65 8.7 30.5 0.89 14.3 6:8  103
0.5 d 175 4.5 32.7 0.89 10.8 5:9  103
1.5 d 188 1.9 32.3 0.89 18.3 6:2  103
7d 189 3.6 44.7 0.89 74.0 3:8  103
a
This becomes polarisation resistance for TiN/SS systems.
C. Liu et al. / Corrosion Science 45 (2003) 1257–1273
Table 2
EIS data simulations for MS and SS with PVD CrN coatings in 0.5 N NaCl solution
PVD Exposed Ecorr Re Y 0  Qc n  Qc Rpore a Y0  O B Y0  Qs n  Qs Rs
p p
systems time (mV, SCE) (X cm2 ) (l/cm2 sn X) (kX cm2 ) (l/cm2 s X) ( s) (l/cm2 sn X) (kX cm2 )
CrN/MS 0.5 h 36 5.2 2.5 0.68 111 16.8 1.5 12.8 0.79 8792
2h )375 7.9 3.5 0.89 4.1 75.1 0.9 14.7 0.55 157
1.5 d )453 3.3 18.3 0.89 2.5 125 1.5 577 0.82 11
7d )585 4.4 64.1 0.96 12.4 138 2.5 881.5 0.64 5
CrN/SS 0.5 h 13 9.0 1.6 0.94 31.7 3.4 5.3
0.5 d 49 1.8 2.4 0.92 38.6 4.1 5.5
1.5 d 75 4.5 4.6 0.91 20.2 8.0 1.8
7d 141 1.7 9.7 0.93 0.3 3.6 4.5
a
This becomes polarisation resistance for TiN/SS systems.

1265
1266 C. Liu et al. / Corrosion Science 45 (2003) 1257–1273

correlate to the dielectric properties of PVD coatings. Rpore is the resistance to current
flow through the pores:
ql
Rpore ¼ ð1Þ
PA
where q is the resistivity of solution in pores, l and A are the length of pores and the
exposed test area. P is the porosity of the coatings. l is equivalent to the coating
thickness, which is approximately 10 lm, while q is dependent on the solution, the
size and shape of defects. Accordingly, P is inversely proportional to the Rpore ; the
greater Rpore , the smaller the P .
Coating capacitance Cc is determined by the composition and structure of the
coating material [9]:
Cc ¼ Y0  Qc ¼ ree0 ð1  P ÞA=d ð2Þ
Here, e0 is the permittivity of free space, e the dielectric constant of the coating
material. d is the thickness of coating. r the surface roughness factor which is 2 here
for PVD coatings [9]. d ¼ 10 lm and e0 ¼ 8:854  1012 F/m for both TiN and CrN
coatings. Cc for CrN/MS and TiN/MS systems obtained from EIS modelling (Tables
1 and 2) are 2.5 and 52.6 lF cm2 , respectively. Thus, the dielectric constants (e) of
the TiN and CrN coatings are estimated to be 26.7 and 1.3.
The admittance constant Y0  Qc , Y0  Qs and exponent factors n  Qc and n  Qs
used to define the Qc and Qs represent the capacitive characteristics of coating/
electrolyte and steel/electrolyte interfaces. According to Tsai [24]:
Cc ¼ Y0  Qc ¼ Cc0 ð1  P ÞA ð3Þ

Cs ¼ Y0  Qs ¼ Cs0 PA ð4Þ
where Cc0 , Cs0 are normalised capacitance of coating and steel, respectively, which
remain unchanged during exposure. The exponent n is usually related to the surface
texture or morphology of an electrode, n is always smaller than unity [17,25]. Ac-
cording to Tables 1 and 2, as the value of Y0  Qs increases with time, the porosity of
the exposed mild steel for TiN/MS or CrN/MS system increases. This suggests that
more and more pores are gradually revealed because of the penetration. However,
the exposed area of coating (Ac ) varies with time in two ways. On the one hand, Ac
decreases due to the exposure of more pores and defects with time. On the other
hand, Ac also increases as more walls of pores are revealed. Cc (i.e. Y0  Qc ) increases
with immersion time in Tables 1 and 2, this indicates the resultant Ac increase due to
the increase in wall area is predominant. However, the decrease of n  Qc with time
suggests a less capacitive interface of the coating/solution, for instance, the coating
with the pores performs as a ‘‘leaky capacitor’’ [4,5,26]. However, AISI 316L steel
generated capacitive response through 7 d of exposure, the surface morphology of
the TiN/SS or CrN/SS systems is unchanged with time, thus, n  Qc almost remains
constant close to unity.
The diffusion of reactant agents (e.g. oxygen) in the electrochemical corrosion of
PVD coated steel is microscopically confined within pores or grain boundaries.
C. Liu et al. / Corrosion Science 45 (2003) 1257–1273 1267

Fig. 7. A schematic illustration of oxygen diffusion mechanism for (a) columnar (e.g. TiN), and (b)
equiaxed (e.g. CrN) crystallite coating structure. Ci is the concentration of reactant (e.g. oxygen) at the
coating/steel interface, the thickness of coatings is 10 lm.

Therefore, the microstructure of PVD coatings can significantly influence the dif-
fusion behaviour. As schematically represented in Fig. 7, the oxygen diffusion, which
is of great concern in this work, can be described in terms of the diffusion path for
two typical coating microstructures. The straight columnar structure (e.g. TiN
shown in Fig. 2a) allows the oxygen concentration to gradually decreases across the
electrolyte/coating interface Fig. 7a. However, the zig-zag grain boundaries of fine
equiaxed structure (e.g. CrN shown in Fig. 2b) could result in a dramatic decrease in
oxygen concentration across the electrolyte/coating interface Fig. 7b. According to
the boundary conditions of the diffusion theory, the former is known as semi-infinite-
length diffusion, while the latter is finite-length diffusion. This explains why the use of
Warburg impedance W was suitable for EIS modelling of TiN coated systems, and
the co-tangent-hyperbolic diffusion impedance (O) was used for EIS modelling of
CrN coated systems [17]. The mathematical expressions of admittance of W and O
elements are given below:
1=2
Y  W ðxÞ ¼ Y0 ðjxÞ ð5Þ

1=2 1=2
Y  OðxÞ ¼ Y0 ðjxÞ cotanh½BðjxÞ
ð6Þ
Accordingly to Franceschetti and Macdonald [27]:
p
Y0  W ; or Y0  O ¼ le =ðRD0 DÞ ð7Þ
p
B ¼ le = D ð8Þ
where RD0 and D are the diffusion resistance (the resistance when x ! 0 limit) and
the diffusion coefficient respectively. In Tables 1 and 2, Y0  W of PVD coated mild
steel is several orders greater than that of stainless steel. A higher diffusion resistance
1268 C. Liu et al. / Corrosion Science 45 (2003) 1257–1273

(RD0 ) is expected for PVD coated AISI 316L steel due to the formation of the oxide
films. However, increase of Y0  W with immersion time for the coated MS, is at-
tributed to the decrease of the diffusion coefficient (D) by ÔpluggingÕ of corrosion
products through pores, making the diffusion more difficult according to Eq. (7).
Interestingly, Y0  W for TiN coatings was 2–3 orders of magnitude higher than
Y0  O for CrN, which may indicate that diffusion coefficient (D) of CrN coatings is
2–3 orders smaller than that of TiN. This can further confirm that the CrN coatings
with the equiaxed crystallite structure are more effective in reducing the speed of
diffusion, which is certainly desirable to enhance corrosion resistance.

3.4. Porosity and localised corrosion of PVD hard coatings

The porosity of PVD ceramic coatings has been extensively studied by a number
of research groups using electrochemical measurement techniques [1,13,21,28–30].
Fundamentally, the evaluation of porosity by electrochemical methods is based upon
the ratio of the current density through the pores and the coating [30].
isr  ðia;c þ ic;c Þ
P¼ ð9Þ
ia;u þ ic;s
The subscripts a and c represent anodic and cathodic, and s and c the substrate and
coating, respectively. In most cases, the cathodic currents are negligible for deter-
mination of porosity, and the current density is inversely proportional to the po-
larisation resistance. Therefore, for PVD ceramic-coated mild steel, an equation for
the estimation of porosity is given below [1]:
 
Rp;s jDEcorr j
P¼  10 ba ð10Þ
Rp
where Rp is the polarisation resistance of the PVD coated mild steel systems, DEcorr is
the difference in corrosion potential between the uncoated and PVD coated mild
steel, Rp;s and ba are the polarisation resistance and Tafel slope of the active disso-
lution of mild steel, respectively. Rp;s (870 X cm2 ) and ba (50 mV/decade) were de-
termined by a separate measurement with the bare mild steel substrate in the same
conditions. Ecorr (650 mV vs. SCE) of mild steel was also obtainable. Rp ¼ Rpore þ Rs ,
the values of Rpore and Rs are given by EIS modelling in Tables 1 and 2. For the
evaluation of porosity, the values of Rpore and Rs for the initial exposure (i.e. 30 min
or 2 h) are chosen to ensure that only the pre-existing pores were exposed to the
solution, and no further delamination is counted. P can then be determined in
Table 3.
The porosity of PVD TiN coatings determined at 30 min of exposure is ap-
proximately 0.0025%, which is an order lower in magnitude than the values reported
by the other groups [1,13,21,28,30]. This may indicate that the 30 min exposure is not
sufficient to produce a response in every pre-existing pore. It needs a longer time to
allow the electrolyte to penetrate the finer defects such as micropinholes. In partic-
ular, the thickness of the multi-layer coatings in this work is about 10 lm, which is
probably the reason for the need for longer exposure. The porosity determined after
C. Liu et al. / Corrosion Science 45 (2003) 1257–1273 1269

Table 3
Porosity of PVD TiN and CrN coatings determined by AC (EIS) electrochemical testinga
Immersion time TiN/MS CrN/MS
Rp (MX cm2 ) P (%) Rp (kX cm2 ) P (%)
0.5 h 0.35 0.0025 17.9 0.0001
2.0 h 8.9 0.04 161.4 0.005
1.5 d 5.1 0.17 13.6 0.06
7.0 d 6.1 0.14 17.4 0.05
a
Ecorr values in mV vs. SCE of 550 (for TiN and CrN) were used upon stable potential measurements.
Rp ¼ Rpore þ Rs , Rpore and Rs was obtained from EIS modelling (Tables 1 and 2), Rpore , Rs varies as im-
mersion time increases.

2 h was similar to other reports. By then, the electrolyte had penetrated through the
coating via pre-existing pores or defects to establish the galvanic cells, but the cells
have not propagated or extended. According to Table 3, the porosity of PVD CrN
coatings determined at 2 h exposure is 0.005%, this value is one order lower than
0.04% of TiN, which suggests that the CrN coatings are preferable in terms of the
enhancement of corrosion performance. Indeed, the cross-sectioned microstructure
for TiN and CrN coatings observed by SEM (Fig. 2) showed that CrN coatings are
generally finer and denser than TiN. Sonobe et al. [21] reported their measurements
of porosity of PVD CrN coatings which were produced by a hollow cathode dis-
charge PVD process using multi-stage coating method. According to their results,
the multi-stage coated steel exhibited much better corrosion resistance than that of
the single stage coatings. The area ratio of defects (which relates to the porosity) was
dramatically reduced due to the application of a multi-staged coating. The magni-
tude of porosity for the CrN coatings of 10 lm thickness from Sonobe et al.
reached 105 , this shows a good agreement with the porosity determined in the
present paper. Sonobe et al. also demonstrated the mechanism of reduction of po-
rosity by a multi-staged coating process, i.e. that the discontinuity of deposition
through several stages can drastically reduce the tendency to form pores. However,
PVD coatings prepared in this work contained several metallic interlayers (e.g. a-Ti
or a-Cr), that are expected to further reduce porosity. Tato and Landolt [30] found
that a thicker interlayer of Ti (1.2 lm) can significantly reduce the porosity of the
PVD TiN coatings (0.8 lm thick) on a brass substrate. This is attributed to the
discontinuity of electrochemical behaviour due to the coating microstructure. For
instance, the formation of a thin oxide film at the defects due to the passivation of
the a-Ti or a-Cr interlayers may further resist the penetration of electrolyte [9].
As immersion time increases, galvanic corrosion localised at the pores will
propagate, which dissolves the mild steel, thus causing delamination of the coating
adjacent to the pores. Delamination degree with exposure time may be estimated
using Eq. (10). Table 3 also listed P values calculated from the Rp values of PVD
TiN, and CrN coated mild steel after 1.5 and 7 d exposure, respectively. Accordingly,
the delamination was indicated to be greater for TiN coatings than CrN on mild steel
for the same time of immersion (e.g. 1.5 or 7 d). Fig. 8 shows a typical delamination
of PVD coatings on mild steel, as observed around a pit, where the PVD coating
1270 C. Liu et al. / Corrosion Science 45 (2003) 1257–1273

Fig. 8. SEM micrographs on a pinhole after corrosion test, showing the delamination occurred in adjacent
of the pinhole due to corrosion of mild steel.

cracked and spalled off due to the adhesion failure along the coating/substrate in-
terface.
The pitting corrosion of mild steel with PVD coated mild steel was also studied in
this work using the SRET. Fig. 9a and b show the SRET images obtained from
single layer (2–3 lm) thin TiN and CrN films coated mild steel after 5 d of immersion
in 104 N NaCl, respectively. The bright spots in the images were more electro-
negative in relative to the large dark surrounding areas, which represents an active
anodic pit at a pore in a TiN or CrN covered surface. SRET showed that more pits
initiated on the same scanned area on the TiN than the CrN coatings. The pits were
also more rapidly propagated on TiN coated mild steel compared to CrN. This
evidence further confirmed the above findings from the EIS modelling.

4. Conclusions

The corrosion behaviour of multi-layered PVD TiN and CrN coated steels has
been studied through the modelling of EIS spectra at different times of exposure to
0.5 N NaCl solution. The obtained parameters for circuit elements were interpreted
and correlated with the dielectric characteristics and microstructure of the coated
systems.

1. EIS spectra from TiN/MS or CrN/MS systems presented two time constants
ðsc ; ss Þ, the constant at low xðss Þ represents the electrochemical response of the
localised corrosion at the pores, which was distinguishable from the one at high
x for the coatings. The formation of a thin oxide film on the AISI 316L steel re-
sulted in the response from the pores to be hardly separated for TiN/SS and CrN/
SS systems, such that only one time constant (scþf ) was observed, representing a
highly capacitive surface.
C. Liu et al. / Corrosion Science 45 (2003) 1257–1273 1271

Fig. 9. SRET images by monitoring the pitting corrosion for (a) TiN/MS, and (b) CrN/MS system after 5
d of immersion in 104 N NaCl solution. The bright spots represent the anodic localised corrosion (i.e.
pits) surrounded by large cathodic coating area.

2. An excellent corrosion resistance from the CrN coatings was attributed to the
denser, and finer microstructure of CrN coatings that can significantly eliminate
the porosity, and thus the localised corrosion. In addition, the equiaxed crystal-
lites in the CrN coatings can reduce the corrosion rate, by further restricting
the oxygen diffusion to the coating/steel substrate interface; while the TiN coat-
ings which consist of the columnar crystallites provided the straight boundaries
for the diffusion of oxygen. Two characteristic crystallite features lead to the uses
1272 C. Liu et al. / Corrosion Science 45 (2003) 1257–1273

of two diffusional impedance elements (W and O) for EIS modelling in connection


with their boundary conditions of diffusion processes.
3. Porosity has been quantitatively calculated from the polarisation resistance
(Rp ¼ Rpore þ Rs ). An exposure time of a few hours is necessary to ensure the full
uncovering of pre-existing pores for the determination of Rp . The results obtained
at 2 h exposure has shown that the porosity of CrN coatings was an order in mag-
nitude lower than TiN, which showed a good agreement with the SEM observa-
tion of their microstructure. Delamination of the PVD coatings on mild steel due
to localised corrosion was also investigated as the exposure time increased, indi-
cating a severe and rapid localised corrosion developed on TiN/MS systems. This
was further confirmed by visual inspection, SEM and SRET examination.

Acknowledgements

The author Liu would like to acknowledge an ORS scholarship awarded by


CVCP, UK. The authors would like to thank Dr. J.M. Schneider, Dr. S.J. Dowey
and Dr. M. Bin-Sudin for valuable discussions during this work at RCSE. Thanks
are to Dr. S.B. Lyon at the CPC of UMIST for the SRET testing. The technical
support of Mr. T. Pawson and Mr. G. Robinson is greatly appreciated.

References

[1] B. Elsener, A. Rota, H. Bohni, Mater. Sci. Forum 44–45 (1989) 29.
[2] R. Brown, M.N. Alias, R. Fontana, Surf. Coat. Technol. 62 (1993) 467.
[3] R. Li, M.G.S. Ferreira, Mater. Sci. Forum 192–194 (1995) 237.
[4] C. Liu, A. Leyland, S. Lyon, A. Matthews, Surf. Coat. Technol. 76–77 (1995) 615.
[5] C. Liu, A. Leyland, S. Lyon, A. Matthews, Surf. Coat. Technol. 76–77 (1995) 631.
[6] S. Rudenja, C. Leygraf, J. Pan, P. Kulu, E. Tallimets, V. Mikli, Surf. Coat. Technol. 114 (1999) 129.
[7] D. Hanzel, A.C. Agudelo, J.R. Gancedo, M. Lakatos-Varsanyi, J.F. Marco, Hyperfine Interact. 111
(1998) 93.
[8] J. Piippo, B. Elsener, H. Bohni, Surf. Coat. Technol. 61 (1993) 43.
[9] S. Rudenja, J. Pan, I. Odnevall, C. Leygraf, P. Kulu, J. Electorchem. Soc. 146 (1999) 4082.
[10] L. Cunha, M. Andritschky, L. Rebouta, K. Pischow, Surf. Coat. Technol. 116–119 (1999) 1152.
[11] F. Mansfeld, M.W. Kendig, S. Tsai, Corrosion 38 (1982) 478.
[12] I. Thompson, D. Campbell, Corros. Sci. 36 (1994) 187.
[13] M. Lakatos-Varsanyi, D. Hanzel, Corros. Sci. 41 (1999) 1585.
[14] M.A. Pech-Canul, S. Turgoose, Corros. Sci. 35 (1993) 1445.
[15] C. Deslouis, M.M. Musiani, B. Tribollet, M.A. Vorotyntsev, J. Electrochem. Soc. 142 (1995) 1902.
[16] R.S. Lillard, J. Kruger, W.S. Tait, P.J. Moran, Corrosion 51 (1995) 251.
[17] C. Liu, Q. Bi, A. Leyland, A. Matthews, Corros. Sci. Part I 45 (2003).
[18] C. Liu, Ph.D. Thesis, Hull University, 1997.
[19] H.A. Jehn, Surf. Coat. Technol. 125 (2000) 212.
[20] C. Liu, A. Leyland, Q. Bi, A. Matthews, Surf. Coat. Technol. 141 (2001) 164.
[21] M. Sonobe, K. Shiozawa, K. Motobayashi, JSME Int. J. Ser. A––Solid Mech. Mater. Eng. 40 (1997)
436–444.
[22] P. Engel, G. Schwarz, G.K. Wolf, Surf. Coat. Technol. 98 (1998) 1002.
[23] C. Rebholz, H. Ziegele, A. Leyland, A. Matthews, Surf. Coat. Technol. 115 (1999) 222.
C. Liu et al. / Corrosion Science 45 (2003) 1257–1273 1273

[24] C.H. Tsai, F. Mansfeld, Corros. Sci. 49 (9) (1993) 726.


[25] P.R. Roberge, E. Halliop, V.S. Sastri, Corrosion 48 (1992) 447.
[26] L. van Leaven, M.N. Alias, R. Brown, Surf. Coat. Technol. 53 (1992) 25.
[27] D.R. Franceschetti, J.R. Macdonald, J. Electroanal. Chem. 82 (1977) 271.
[28] J. Piippo, B. Elsener, H. Bohni, Mater. Sci. Forum 111–112 (1992) 219.
[29] P. Ernst, A. Earnshaw, I.P. Wadsworth, G.W. Marshall, Corros. Sci. 39 (1997) 1329.
[30] W. Tato, D. Landolt, J. Electrochem. Soc. 145 (1998) 4173.

S-ar putea să vă placă și