You are on page 1of 22

Corrosion Science 48 (2006) 26892710

www.elsevier.com/locate/corsci

Assessment of stress corrosion crack initiation


and propagation in AISI type 316 stainless
steel by electrochemical noise technique
T. Anita, M.G. Pujar, H. Shaikh *, R.K. Dayal, H.S. Khatak
Corrosion Science and Technology Division, Indira Gandhi Centre for Atomic Research,
Kalpakkam 603 102, Tamil Nadu, India
Received 26 January 2005; accepted 10 September 2005
Available online 8 November 2005

Abstract
Electrochemical noise (EN) analyses, using time domain, frequency domain and statistical parameters, were carried out on the potential and current data obtained on AISI type 316 stainless steel
during (i) pitting corrosion in deaerated 0.5 M sodium chloride (NaCl) solution, and during (ii) stress
corrosion cracking (SCC) in boiling acidied NaCl solution. Visual records and statistical analysis of
the current and potential data alongwith the spectral estimation using maximum entropy method
(MEM) gave useful informations on dierentiating between these corrosion processes and on the
eect of stress in accelerating pitting corrosion. The results correlated well with the optical microscopic observations.
 2005 Elsevier Ltd. All rights reserved.
Keywords: Stress corrosion cracking; Pitting corrosion; Visual records of current and potential; Standard deviation; Power spectral density

1. Introduction
Stress corrosion cracking (SCC) is the degradation of the material under the combined
action of a load and a corrosive medium, neither of which when acting alone would have
*

Corresponding author. Tel.: +91 44 27480121; fax: +91 44 27480301.


E-mail address: hasan@igcar.ernet.in (H. Shaikh).

0010-938X/$ - see front matter  2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.corsci.2005.09.007

2690

T. Anita et al. / Corrosion Science 48 (2006) 26892710

caused the failure. It has a few typical characteristics viz. (i) it occurs only in specic media
in which a critical balance exists between slip step generation rate, corrosion rate and
repassivation rate, (ii) it occurs only in materials that show activepassive behaviour,
(iii) it is a macroscopically brittle failure that occurs in ductile materials, and (iv) it can
occur at stress levels much below yield stress (YS). SCC has been a recurrent problem
in the nuclear industry aecting components like fuel clad, steam generators, turbines, pipings, nozzles, valves etc. In the fast reactors, austenitic stainless steels are the major material for construction of many critical components. These steels are susceptible to SCC in
many environments, more so in the naturally prevailing chloride environments in the
coastal atmosphere. Many SCC failures of nuclear components, made of austenitic SS,
have been reported during storage and during service in the seacoast of Kalpakkam [1,2].
The detection of occurrence of SCC during storage and/or during operation assumes
signicance since costly breakdowns of the operating system can occur without any warning. On-line monitoring of SCC is a gray area of research with a number of non-destructive testing techniques being studied for the purpose of early detection of degradation by
corrosion. Of these, acoustic emission technique and electrochemical noise (EN) techniques have held promise since these do not require input signals that could cause perturbations to the operating systems. In some systems, SCC could initiate from pits and this
would require clear demarcation of the signals emitting from pitting corrosion and SCC
processes. Electrochemical noise is a versatile technique to clearly distinguish between various forms of corrosion. Hence, its suitability in assessing on-line damage caused by simultaneously occurring corrosion processes is being rigorously pursued. Electrochemical noise
(EN) is referred to as the spontaneous current and potential uctuations in the process of
corrosion. Spontaneous EN can be a rich source of information from corrosion processes
[3].
Iverson rst described the measurement of electrochemical noise (EN) for corrosion
measurement in 1968 [4]. In many pioneering research papers by Tyagai [57] investigations were conducted to establish EN technique to be useful for monitoring corrosion.
EN technique measures the stochastic uctuations of the corrosion potential and the
corrosion current. Electrochemical noise can be measured in potentiostatically polarized
conditions and in freely corroding system. The rst technique is suitable for the study of
corrosion processes [8], while the second technique is more appropriate for corrosion
monitoring [9]. Analysis of the spontaneous uctuations of potential and current is used
for understanding of dierent corrosion processes, such as general corrosion, pitting corrosion, crevice corrosion, stress corrosion cracking (SCC) as well as passive lm buildup
[1015]. The qualitative predictions of the corrosion phenomena were attempted using
the statistical parameters such as standard deviations and root-mean-squared (rms) values of current and potential [10,11,16]. Most of the EN studies have been focused on
pitting corrosion and on passivation phenomenon, while a few studies are related to
SCC [1724].
SCC may propagate by discontinuous or continuous processes, and these have dierent
EN characteristics. In the rst reported work on EN during SCC of a-brass in 1 M sodium
nitrite solution by Newman and Sieradzki [17], it was found that the current noise had a
good correlation with acoustic emission events. They used random uctuations in current
during transgranular SCC of a-brass in sodium nitrite to demonstrate the discontinuous
nature of cracking in that system [17]. Loto and Cottis measured EN during SCC of abrass, aluminum and high strength steel [1820]. By using the frequency domain analysis

T. Anita et al. / Corrosion Science 48 (2006) 26892710

2691

of potential uctuations, they concluded that crack initiation and specimen fracture gave
the highest power spectral density (PSD). They concluded that standard deviation and
roll-o slopes were the indicators of cracking of the specimen [18,19]. However, Bagley
showed that dierent forms of corrosion could not be reliably distinguished based on
the roll-o slope alone [25]. Based on their work on type 304 stainless steel and a-brass,
Luo and Qiao [26] concluded that standard deviation of the noise signal was a useful
method to determine the initiation of stress corrosion cracks. They suggested that during
analysis in the frequency domain, the DC component in the potential data should be ltered before the PSD plot can be used to distinguish between random noise and SCC
noise. They found that the characteristic noise signals generated during SCC consisted
of a quick drop and slow recovery of potential and that potential drop time and potential
recovery time were approximate constants for a given SCC system. Inoue et al. [27] and
Yamakawa and Inoue [28] studied the EN of type 304 stainless steel in boiling magnesium
chloride solutions. They found EN with a characteristic potential uctuation pattern in
25% and 35% magnesium chloride solutions but not in 42% magnesium chloride solution.
In another study on EN analysis of sensitized type 304 stainless steel in high temperature oxygenated water at 461 K, Stewart et al. [22] used the shot noise analysis which correlates the current transients, charge and mean frequency of charge emission. They
reported a good one-to-one correspondence between the number of current transients
and the number of cracks. They observed smooth transients (without spikes) with short
transient rise times and very slow repassivation from peak current. They attributed the
decrease in the rate of current rise with increasing strain to mechanical separation along
the grain boundary in front of the main crack. Some reports suggested that frequency
of current and potential signals decreased with decreasing applied stress and that there
was a clear periodicity in these signals [23,24].
In the present work, the EN signals from AISI type 316 stainless steel in unstressed and
stressed (U-bend) conditions in deaerated sodium chloride solution at room temperature,
and stressed (U-bend) condition in boiling acidic chloride solution have been analysed to
understand the current and the potential transients during metastable and stable pitting,
and during the initiation and growth stages of stress corrosion cracks.
2. Experimental procedures
Electrochemical noise (EN) studies were carried out on solution annealed AISI type 316
stainless steel (chemical composition, wt.%: C = 0.055%, Cr = 18.00%, Ni = 11.86%,
Mo = 2.30%, Mn = 1.7%, Si = 0.45%, P = 0.067%, S = 0.027%, Fe balance) in both the
stressed and unstressed conditions. In the unstressed condition, EN studies were performed using two nominally identical electrodes of the same size (50 25 3 mm) that
were polished up to 600 grit, washed in soap water and degreased in acetone.
The EN studies were carried out using an electrochemical system of Solatron make. The
electrochemical system actively holds the working electrode connection at the ground
potential by a small amplier circuit. If one working electrode is directly connected to
ground and the other is connected to the working electrode cable, they are both held at
the same potential and are, in eect, shorted together. Any current which ows between
the two electrodes is measured by the current measurement circuit thus creating a Zero
Resistance Ammeter (ZRA). The potential is measured between the working electrodes
(since they are shorted together, both working electrodes are at the same potential)

2692

T. Anita et al. / Corrosion Science 48 (2006) 26892710

and a reference electrode. The frequencies above the Nyquist limit were ltered out by the
anti-aliasing lters incorporated in the electrochemical system.
Electrochemical noise measurements during pitting corrosion of type 316 stainless steels
were conducted in deaerated 0.5 M NaCl solution at open circuit potential (OCP) in both
the stressed and unstressed conditions. In the unstressed condition, rectangular coupons,
with 8 cm2 of area immersed in solution, were used to obtain the electrochemical noise
data. In the stressed condition, two U-bends with the same immersed surface area of
22 cm2 were used to obtain the electrochemical noise data during pitting corrosion. Noise
signals were collected at the sampling frequency of 4 Hz. Saturated calomel electrode
(SCE) was used as a reference electrode for the measurement of potential noise. Electrochemical noise measurements during SCC of type 316 stainless steels were conducted using
U-bend specimens in a boiling solution of 5 M NaCl + 0.15 M Na2SO4 + 2.5 ml/l of HCl
(pH = 1.3; b.p. = 381 K). In every test, two nominal electrodes in U-bend condition with
the same surface area were used as the working and the counter electrodes. For both the
pitting corrosion and SCC studies, the U-bend specimens were prepared after polishing
type 316 stainless steels strips of 100 10 3 mm up to 600 grit nish, cleaning with soap
solution followed by rinsing in acetone. The strips were bent using an 18 mm diameter
mandrel. The U-bend was then bolted using Incoloy 800 pins, which have higher stiness
than type 316 stainless steels and hence prevent the relaxation of stresses that could occur
by outward movement of the arms of the U-bend. The specimen area other than that seeing the environment was coated with a free owing silicone elastomer. The area of the Ubent working electrode immersed in the solution was 22 cm2. One end of the U-bend was
soldered to a stainless steel rod of the same chemical composition. The soldered joint was
insulated by application of silicone elastomer and was well above the solution.
The linear trend component was estimated by least squares method and then eliminated
by subtraction using commercial software. All the potential and current noise data collected in the time domain were transformed in the frequency domain through the MEM
method, by a dedicated software. It has been reported that MEM computed with a few
coecients produces a very smooth spectrum, preferable to that obtained by FFT, even
after averaging and that is why low values of MEM order are commonly used in corrosion
applications [29]. However, Bertocci et al. [30] reported that MEM computed with an
order of 10 gave better precision than FFT with one average, and hence, lower values
of M were adequate. However, in the present work, PSD data were obtained by computing
MEM with an order of 15. In order to reduce the leakage of the low as well as the high
frequencies in the calculated PSD values, Welch window was used for signal analysis.
The initiation and propagation of SCC was studied by the analysis of the 8092 data
points, which were conducted at suitable time intervals to study the current PSD and other
statistical parameters.
3. Results and discussion
The results of the potential and current variations with respect to time were analysed
for EN behaviour using the time domain, frequency domain and statistical analyses to
detect metastable and stable pitting, and SCC. The time domain analysis was studied
by a detailed analysis of potential vs. time and current vs. time data. The data were analysed for every 2500 s for convenience. The statistical parameters, namely standard deviation of current (rI), standard deviation of potential (rV), noise resistance, RN = (rV/rI)

T. Anita et al. / Corrosion Science 48 (2006) 26892710

2693

[31], and the dc limit of the spectral noise resistance at zero frequency, R0SN , were then calculated. Mathematically R0SN is represented as,
R0SN limfRSN g
f !0

Since dc limit was not usually detected in the present experiments, R0SN was calculated as
the mean value of the 10 data points at the lowest frequencies [32]. Earlier it was reported
by Pistorius [33] that a direct relationship between the increase in the number of transients
with the increase in the electrode area existed. Current noise increased with increasing corrosion rate and the standard deviation of the current noise tracked the changes in the corrosion rate. However, their quantitative relationship depended on both electrode area and
the corrosion process [33]. However, recent research indicated that there was no unique
relationship and that dierent area eects were possible [29]. Thus, it was reported that
normalisation of the electrochemical current noise data by dividing the amplitude of the
current by the area was improper, and that both current and potential noise data must
be reported as recorded alongwith the specimen area. Other parameters used in the frequency domain analysis were coecient of variation of current and roll-o slope. The nature of the corrosion process was determined using the above parameters and supported by
optical microscopic observations.
3.1. Detection of pitting corrosion
Fig. 1(a) and (b) shows the current and potential noise records for type 316 stainless
steel specimen in unstressed and stressed conditions respectively immediately after the
immersion in deaerated 0.5 M sodium chloride solution. The typical current potential
transients observed during the metastable pitting events for unstressed and stressed type
316 stainless steel are shown in Fig. 2(a) and (b). Fig. 3(a) and (b) shows the current
and potential transients during the development of stable pitting events on unstressed
and stressed type 316 stainless steel in deaerated chloride solution. Fig. 4 shows that the
current and potential transients for the stressed material during the growth of stable pits.
The plots of standard deviations of current (rI) and potential (rV), noise resistance, RN,
and spectral noise resistance at zero frequency, R0SN , for the pitting corrosion process
are shown in Fig. 5(a) and (b).
It is seen from Fig. 1(a) and (b) that in both, the unstressed and stressed, conditions,
irregular current transients occurred at the start of the experiment, the amplitude of the
current transients being higher for the stressed condition. The potential noise in case of
the unstressed condition showed a constant decrease while in the case of stressed specimen
the potential transients showed a slow rise followed by a rapid fall. Uruchurtu and Dawson attributed these irregular current transients to the occurrence of breakdown and passivation processes occurring on the surface [34]. This could also be due to the dissolution
or breakdown of the Cr2O3 layer and subsequent dissolution of the bare metal resulting in
the formation of an environmentally compatible passive lm. The current transients were
of higher magnitude in the stressed material due to increased dissolution of the material
under stress. The increased dissolution in the stressed material was attributed to the fact
that, in the initial period of up to 7500 s, the rI values, shown in Fig. 5(a), continuously
increased in the stressed specimen compared to the unstressed one which showed a continuous decrease in the rI values in the same time duration. The potential noise of the

2694

T. Anita et al. / Corrosion Science 48 (2006) 26892710


0.05

1.0x10
Potential
Current

5.0x10

0.00

0.0

Current, A

-1.0x10

-0.10

-1.5x10

-0.15

-2.0x10
-0.20
-2.5x10
-0.25

-3.0x10

-0.30

-3.5x10
0

100

200

300

500

Potential
Current

3.5x10

-0.16

3.0x10

-0.18

2.5x10

-0.20

2.0x10

-0.22

1.5x10

-0.24

1.0x10

-0.26

5.0x10

-0.28

Potential, V(SCE)

Current, A

400

Time, s

(a)

-0.30

0.0
0

(b)

Potential, V(SCE)

-0.05
-5.0x10

100

200

300

400

500

Time, s

Fig. 1. Current and potential noise in deaerated chloride solution for (a) unstressed and (b) stressed specimens, at
the start of the test.

unstressed material showed a constant decrease, indicating a rapid initial breakdown of


the naturally occurring passive lm on the specimen surface. The potential and current
transients of the stressed specimen showed trends that were opposite to each other. Every
drop in current corresponded to a rise in potential indicating passivation of the surface;
while every rise in current was accompanied by a drop in potential indicating dissolution
of the material. This indicated repetitive breakdown and recovery of the passive lm during the early stages of the test. Gonzalez-Rodriguez et al. suggested that the generation of
noise signals was due to the rupture of the oxide layer from straining and the active dissolution of the electrode due to the action of the electrolyte on the bare surface [35]. This
behaviour suggested that the passive condition is of a dynamic nature. Okamoto [36] and
Fromhold [37] explained the sudden current and potential transients in terms of a crack/
heal mechanism at the base of the aws or defects caused by the combined action of compressive and/or tensile stresses present in the air-formed oxide, which expose bare metal
transiently to the electrolyte. The relief of these stresses is expected to occur at the aws
in the oxide lm, where the lm is thinner and weakened leading to its easy dissolution [38].

T. Anita et al. / Corrosion Science 48 (2006) 26892710


-0.2270

1.0x10
Potential
Current

8.0x10

-0.2272

-0.2276

4.0x10

-0.2278

2.0x10

-0.2280
0.0
-0.2282

Potential, V(SCE)

-0.2274

6.0x10

Current, A

2695

-2.0x10
-0.2284
-4.0x10
-0.2286
2370

2372

2374

2376

2378

2380

Time, s

(a)

-0.165

Potential
Current

1.0x10

-0.170

Current, A

-0.175
6.0x10

-0.180
-0.185

4.0x10

-0.190
2.0x10
-0.195
0.0

-0.200

-2.0x10

-0.205
1150

(b)

Potential, V(SCE)

8.0x10

1160

1170

1180

1190

1200

1210

Time, s

Fig. 2. Visual records of current and potential in deaerated chloride solution for (a) unstressed and (b) stressed
specimens, during metastable pitting.

The overall trend of the potential transients was seen to be shifting in the anodic direction
due to the formation of environmentally compatible passive lm. However, the formation
of this passive lm was almost instantaneous in case of stressed material while it took a
few thousand seconds for unstressed specimens. The spontaneity of formation of this passive lm in case of stressed specimen was attributed to the higher material dissolution as
indicated by a larger current signal and the initial rI values that are almost one to two
orders of magnitude higher than the unstressed material.
Pits formed on stainless steel surfaces could either get repassivated before achieving stability or grow to become stable pits [3943]. The phenomenon of metastable pitting is
dened as the initiation of a single pit that gets repassivated immediately without growing
further [34]. Fig. 2(a) and (b) shows that metastable pitting is accompanied by short potential drops and/or small rise in current transients, with each metastable pitting event
occurring over a time range of about 10 s. This observation was also reported by Isaacs
[41] and Williams et al. [42]. They reported that metastable pitting occurs well below
the critical pitting potential. In the current transients observed during metastable pitting,
the anodic current rose in at least two or three steps. This was reported earlier in case of
corrosion of AISI type 430 stainless steel in acidic chloride medium having a pH = 3 [44].

2696

T. Anita et al. / Corrosion Science 48 (2006) 26892710


-7

4.0 x 10
3.5 x 10

-0.1974
-0.1975
-0.1976

-7

2.5 x 10

-0.1977
-7

2.0 x 10
1.5 x 10
1.0 x 10

-0.1984

-7

-0.1984

-7

-0.1984
-0.1984

-8

5.0 x 10

-0.1984

0.0
-5.0 x 10

-0.1984

-8

-0.1984
10760

10780

(a)

10800

10840

10820

Time, s
8.0 x 10
7.0 x 10

5.0 x 10
4.0 x 10
3.0 x 10
2.0 x 10
1.0 x 10

-6

-0.13

Potential
Current

-6

-0.14

-6
-6

-0.15

-6

-0.16

-6
-6

-0.17

-6

Potential, V(SCE)

6.0 x 10

Current, A

Potential, V(SCE)

-7

3.0 x 10

Current, A

-0.1973

Potential
Current

-7

-0.18
0.0
-1.0 x 10

-6

-0.19
3600

3800

(b)

4000

4200

4400

Time, s

Fig. 3. Current and potential signals in deaerated chloride solution for (a) unstressed and (b) stressed specimens,
during development of a stable pit.

-0.10
7.0 x 10

Current, A

5.0 x 10
4.0 x 10
3.0 x 10
2.0 x 10
1.0 x 10

Potential
Current

-5

-0.15
-0.20

-5

-0.25
-5

-0.30
-5

-0.35
-5

-0.40
-5

Potential, V(SCE)

6.0 x 10

-5

-0.45

0.0
-0.50
-1.0 x 10

-5

4800

4900

5000

5100

5200

Time, s
Fig. 4. Current and potential signals in deaerated chloride solution during the growth of a stable pit in stressed
specimen.

T. Anita et al. / Corrosion Science 48 (2006) 26892710

2697

-1

Without Stress (v)


With Stress (v)
With Stress (l)
Without Stress (l)

10-5

10

10

V, V

I, A

10-2
-6

10-3

10

-7

10-4
103

(a)

104

105

Time, h

10

10

10

(b)

1014

10

12

10

10

10

10

Without Stress (RN)


With Stress (RN)
Without Stress (R0SN)
With Stress (R0SN)
2

103

104

105

R0SN, Ohm.cm2

RN , Ohm.cm 2

105

Time, h

Fig. 5. The statistical parameters plotted against time for the stressed and unstressed condition, (a) standard
deviation of current and potential and (b) noise resistance and spectral noise resistance.

Hence, it appears that the metastable pitting could be attributed to the presence of chloride ions more than the pH of the medium. During metastable pitting, every potential drop
corresponds to a single lm rupture event and the recovery that follows corresponds to
repassivation.
Fig. 3(a) and (b) shows that the current and potential transients were extremely random
in nature during stable pitting due to the stochastic nature of pitting. The unstressed specimen (Fig. 3(a)) showed a larger number of transients than stressed specimens (Fig. 3(b)),
possibly due to the formation of both stable as well as metastable pits. However, Fig. 4
shows that the current and potential transients for the stressed material are well dened
and show that during the growth of stable pits, the rise of current and the fall of potential
transients are simultaneous. The current value of the transients is observed to remain at a
very high value continuously for several hundred seconds while the potential is active for
the same time duration. The associated potential drop during the pit growth event is about
325 mV. When a metastable pit can no longer repassivate, a suitable electrochemical

2698

T. Anita et al. / Corrosion Science 48 (2006) 26892710

condition is created for it to grow into a stable pit. The dissolution inside the pit now proceeds unhindered leading to a high current signal and active potential ensues due to the
creation of bare surface when pit depth and size increase. During a singular pit initiation
event in the working or counter electrode, the release of electrons associated with the localized anodic dissolution of the pit bottom momentarily lowers the potential of the pitting
electrode surface and the other coupled electrode relative to the reference electrode, creating a negative potential transient in the potential signals.
Fig. 5(a) shows that rI and rV follow similar trends for a given stress condition. The
unstressed material showed a gradual decrease in rI and rV with increasing time while
the stressed material showed an initial rise in the values of rI and rV, which was followed
by a sharp decrease. The initial rise in the values of rI for the stressed material was attributed to increased material dissolution vis-a`-vis the unstressed material. As pitting corrosion is a stochastic process and pits initiate on a passive surface, the initial interaction
of the alloy with the electrolyte is to regenerate or reform the passive lm after dissolving
the naturally existing Cr2O3 lm. It would take some time before the chloride ions can get
adsorbed on the regenerated passive lm at the defective or weak spots. The gradual
decrease in rI and rV values exhibited by the unstressed material is attributed to this phenomenon. However, in the stressed condition, current signals with high amplitude resulted
in the initial high rI values. This could be correlated with the stress-assisted dissolution of
the initially existing Cr2O3 lm. The subsequent drop in the rI and rV values could be
attributed to the regeneration of environmental compatible passive lm after dissolution.
Following this drop, the values of rI and rV oscillate between a maximum and a minimum
in the stressed specimen. These oscillations could be associated with the continuous dissolution and repassivation processes taking place during growth of the pit. Fig. 5(b) shows
the trends of RN and R0SN for the stressed and unstressed materials. It is seen that in case of
unstressed material of RN and R0SN show opposite trends, while in case of stressed material
they show trends that are very similar except in the early stages of pitting. The variations
in RN and R0SN can also be explained based on the observations for rI and rV.
A powerful analysis of EN signals can be performed by transposing the data in the
frequency domain [45]. The frequency dependence of the signal is considered to be a
potentially very important analytical technique. Also, since extraneous (electrical) interference in the measurements typically occurs at the specic frequencies (e.g. 5060 Hz),
transposing the data in the frequency domain allows the data to be distinguished from
the extraneous or system noise [33,46]. This method estimates the power present at various frequencies included in the signal, the process being known as spectral estimation.
The power spectrum obtained gives power spectral density (PSD), which is the power
per unit of frequency, as a function of frequency [29]. The usefulness of this type of analysis would be complementary due to the fact that the number of the events per unit time
is expected to increase with increase in corrosion rate [33]. Earlier it was reported that
the power level of the signal as calculated from the PSD in the frequency-independent
region seemed to indicate the corrosion rate [47]. The current PSD plots in Fig. 6(a)
and (b) shows the dependence of current PSD on frequency for dierent time durations
for the stressed and unstressed conditions respectively. It is seen that the gure can be
divided into two regions, one in which the power density does not vary with frequency
(i.e. frequency-independent region) and the other in which it decreases continuously with
frequency. It is seen that the frequency-independent region increases at higher time durations time for both the stressed and unstressed material. The current PSD in the

T. Anita et al. / Corrosion Science 48 (2006) 26892710

2699

-180
2500 s
5000 s

-200

Current, dB

7500 s
10000 s

-220

12500 s
15000 s

-240

17500 s

-260
-280
-300
10 -5

10 -4

(a)

10 -3

10 -2

10 0

Log frequency, Hz

-150

Current, dB

10 -1

2500 s

12500 s

5000 s

15000 s

7500 s

17500 s

10000 s

-200

-250

-300
10 -5

(b)

10 -4

10 -3

10 -2

10 -1

10 0

Log frequency, Hz

Fig. 6. The current PSD plots for (a) unstressed condition and (b) stressed condition in deaerated 0.5 M sodium
chloride solution.

frequency-independent region for both the unstressed and stressed materials were plotted
as a function of time (Fig. 7(a)). The MEM allows computation of the spectra below
fmin; however, it is imperative not to calculate PSD below fmin since it would generate
misleading information [30]. Thus the PSD values calculated at fmin (4.94 104 Hz in
this study) and above were used to plot Fig. 7(a). It is observed that change in the
corrosion rate indicated by the current PSD values is very similar to the one indicated
by rI values shown in Fig. 5(a). This result also shows the close correlation of the PSD
data with those obtained by statistical parameters.
The important parameters used for identifying the localized corrosion are the coecient of variation of current (CVC), and the localization parameter or localization index.
The CVC is large in case of localized corrosion [48]. Fig. 7(b) shows the dependence of
CVC on time. It is noticed that the unstressed material showed a sharp peak at 10,000 s
conrming the pitting corrosion attack on the specimen, whereas the stressed specimen
showed a broad peak with a lower CVC at 7500 s. The lower value of CVC for stressed
material could be related to higher values of current associated with a more lateral material dissolution due to lack of passivity in cold worked materials [49]. The roll-o slope

2700

T. Anita et al. / Corrosion Science 48 (2006) 26892710


-140

Without Stress
With Stress

Current, dB

-160
-180
-200
-220
-240
-260
-280
0.0

4.0 x 10

Coefficient of variation of current

8.0 x 10

1.2 x 10

1.6 x 10

2.0 x 10

Time, s

(a)

Without Stress

40

With Stress

30
20
10
0
-10
0.0

4.0 x 10

(b)

8.0 x 10

1.2 x 10

1.6 x 10

2.0 x 10

Time, s

Roll-off slope, dB/decade

50
Without Stress (l)
With Stress (l)
Without Stress (v)
With Stress (v)

45
40
35
30
25
20
15
4.0 x 10

(c)

8.0 x 10

1.2 x 10

1.6 x 10

2.0 x 10

Time, s

Fig. 7. The plots of the dierent parameters obtained from current and potential PSD as a function of time for
both the stressed and unstressed conditions in deaerated chloride solution: (a) current PSD, (b) coecient of
variation of current, and (c) roll-o slope.

of the potential and/or current PSD have been used as an indicator of the type of corrosion [34,5053]. Pitting corrosion generally results in shallow slopes of 620 dB/decade,
while the uniform corrosion or passivation exhibit spectra with slopes higher than this

T. Anita et al. / Corrosion Science 48 (2006) 26892710

2701

value. The roll-o slopes obtained from the current and potential PSD were plotted as a
function of time, as seen in Fig. 7(c). Though, many workers have reported the identication of the corrosion mechanism using the potential PSD slope values [34,54,55], in
the present work, slopes for the potential PSD were much higher than 20 dB/decade
(expected for the pitting corrosion attack) for both stressed and unstressed specimen
and remained unchanged throughout the course of the experimentation. It was noted
that, the roll-o slope from current PSD decreased continuously for unstressed specimen
and at 10,000 s and 12,500 s, the slope values were found to be <20 dB/decade, thereafter the values increased again. On the other hand, the stressed specimen showed the current PSD slope values that were always higher than 25 dB/decade. However, visual
records clearly showed pitting corrosion attack on the stressed specimen (Fig. 4). The
apparent contradiction in the slope values and the observed current potential signals
are explained as follows. Though the values of the slopes of potential and current
PSD are proposed for identication of the corrosion mechanism, such proposals are typically based on a fairly restricted dataset [56] solely by the group performing the analysis. When the roll-o slopes for a reasonably broad dataset are compared, it is
typically found that there is sucient overlap to render the roll-o slopes of little value
as a general identier of the type of corrosion. Also, these values could dier depending
on the environment and the stress state of the material. It has been reported by Bagley
[25] from her own data and from the literature that it was clear that the dierent forms
of corrosion cannot be distinguished reliably on the basis of the roll-o slope values
alone. The photomicrographs in Fig. 8 show the fully grown pits on the surface of
the unstressed specimen. The photomicrograph in Fig. 9 show the well grown pits on
the bend surface of the stressed material, a few pits of the smaller sizes too are seen adjacent to the bigger pits in the lower part of the micrograph. Thus, it is concluded that the
visual records of the current and potential signals and optical microscopic observations
are equally important in determining the type of corrosion attack apart from the results
obtained by data analysis.

Fig. 8. Photomicrograph showing pitting corrosion attack in unstressed condition exposed to deaerated chloride
solution.

2702

T. Anita et al. / Corrosion Science 48 (2006) 26892710

Fig. 9. Pitting corrosion attack in stressed material exposed to deaerated chloride solution.

3.2. Stress corrosion cracking


The visual records of the potential and current signals for stressed (U-bend) type 316
stainless steel in boiling acidied 5 M NaCl + 0.15 M Na2SO4 (pH = 1.3) solution are
shown in Fig. 10(a)(e). A rapid rise in the current signal with high amplitude at the start
of the test, superimposed with numerous transients is shown in Fig. 10(a). Potential signals, which are cyclic in nature, coming in bursts are seen in Fig. 10(b). These signals were
found to occur between 5 and 22 h after start of the test. These types of signals are also
known as noise bursts. A train of such events is shown in Fig. 10(b). Typical current signals at about the same time, which was attributed to stress corrosion cracks in the bend
region are shown in Fig. 10(c) and (d). Fig. 10(e) shows the current records, at about
32 h after start of the test, and corresponding to propagation of stress corrosion cracks.
Visual examination of the specimens indicated nucleation and growth of a number of
stress corrosion cracks in the region of the U-bend (Fig. 11(a) and (b)).
At the start of the test, the current signals with high amplitude occurred due to rapid
dissolution of the oxide lm under strain in the highly corrosive environment
(Fig. 10(a)). During this period, each potential transient ranged between about
550 mV to about 100 mV. After some time, when the environment compatible passive
lm formed, the amplitude of the current signals also decreased. Though in the initial
stages of the exposure, the amplitude of the current signal was nearly the same as in deaerated 0.5 M NaCl (Fig. 1(b)), it was consistently higher by almost two orders of magnitude
during the later stages of the test, thus indicating the higher corrosivity of the environment. After about 5 h of tests, burst type transients in potential were observed, as seen
in Fig. 10(b). This implied that the necessary condition for SCC to initiate had occurred
at about 5 h after start of test. Also during this period, the range of potential transient
reduced drastically (from about 380 mV to about 280 mV), as compared to start of
the test. Each burst type signal observed could be due to a nucleated crack. The total time
over which these burst type signals were observed in potential was about 17 h, thus implying that the total time for initiation of all the cracks was about 22 h. However, a crack

T. Anita et al. / Corrosion Science 48 (2006) 26892710

2703

Fig. 10. Visual records of the current and potential signals during stress corrosion cracking in boiling acidic
chloride solution (a) current signals at the beginning of the test, (b) potential signals at about 20 h of the test, (c)
current signals at about 20 h of the test, (d) a typical current signal during SCC showing a rapid rise and slow fall,
and (e) current signal at about 30 h of the test.

nucleated early propagated during this period. Hladky and Dawson [10] had observed that
every potential noise burst indicated growth of a single stable pit. It was noted by the same
workers that during the initiation and growth of the stable pit, symmetrically appearing

2704

T. Anita et al. / Corrosion Science 48 (2006) 26892710

Fig. 10 (continued )

potential noise signal was seen [10]. SCC initiation from pits occurs due to a highly localized dissolution resulting from high concentration of chlorides and low pH of environment
and due to a high stress concentration at the base of the pit. Hence SCC initiation is an
extension of the pitting process. It might also be possible that, numerous cracks at various
stages of initiation could be present on the surface and initiation of many cracks might
give the signal pattern as shown in Fig. 10(b). Fig. 10(c) shows that current transients with
a value as high as 3.54.0 105 A were typically observed at almost xed time intervals,
indicating anodic dissolution resulting from initiation of cracks and their subsequent
repassivation. The generation of current transients at regular time intervals is correlated
to the phenomena of material dissolution (increase in a current transient), subsequent
repassivation (fall in a current transient) and a lull period, which coincides with the period
of breakage/dissolution of repassivated lm for exposure of the bare metal. Comparison of
Figs. 3(b) and 10(c) show that during pitting the current transients were random and of
lower amplitude and much more closely spaced as compared to the more widely spaced
larger current transients for SCC. It is seen from Fig. 10(d) that the current transients
showed the typical pattern of rapid rise and a slow fall in a stepwise fashion. After about

T. Anita et al. / Corrosion Science 48 (2006) 26892710

2705

Fig. 11. Stereomicrograph showing SCC cracks initiating through pits on the edges (a) showing broadening of
cracks, and (b) ne branched cracks emanating from pits towards the end of the bend region.

22 h of test, the burst type potential signals disappeared or were of very low amplitude,
thus indicating end of nucleation of stress corrosion cracks. However, crack propagation
continued as observed by the current transients at 30 h of test, seen in Fig. 10(e), which
remained the same and were similarly spaced as in Fig. 10(c). Beyond 30 h of test to completion of test, continuous potential transients ranging between 600 and +100 mV were
observed. Besides a number of stress corrosion cracks initiating from pits, Fig. 11(a) and
(b) shows extensive pitting with which no cracks are associated. The time of occurrence of
these pits is unknown. Also, pits were found to develop on unstressed regions. The bidirectional current signals, seen in Fig. 10(c) and (e), indicated that a similar event of current transients occurred on the counter electrode too.
The plots of rI and rV in Fig. 12(a) showed that rI decreased sharply initially up to
about 5 h after start of the test, then decreased slowly up to about 22 h, increased slowly
thereafter up to about completion of the tests when a sharp increase was observed. Meanwhile, rV decreased slightly and then increased up to about 5 h after start of the test. Then

2706

T. Anita et al. / Corrosion Science 48 (2006) 26892710


10-5

Current
Potential

100

10

-1

V, V

I, A

10
-6

10-2

10-7
0

10

(a)

20

30

40

10-3

Time, h

106

10

R , Ohm.cm 2

108

105

10

10

(b)

107

10

10

0
RSN
, Ohm.cm 2

RN , Ohm.cm 2

RSN, Ohm.cm2

104
0

10

20

30

40

Time, h
Fig. 12. The plots of (a) rI and rV, and (b) RN and R0SN during SCC of type 316 stainless steel in boiling acidic
chloride solution.

it decreased continuously up to 22 h and rose again sharply before evening o after 30 h.


The plots of RN and R0SN vs. time in Fig. 12(b) showed an identical pattern of two peaks,
one at 5 h and the other at about 30 h and a trough at about 22 h. These trends matched
well with the trend shown in rV vs. time plot in Fig. 12(a). The plots of current PSD values
obtained at dierent time intervals are shown in Fig. 13(a)(c). Fig. 13(a) and (b) shows
that the roll-o slope values did not vary much with exposure time. It is seen from
Fig. 13(c) that the current PSD values from the frequency-independent region, plotted
against time, decreased sharply initially up to about 5 h after start of the test, then
decreased slowly up to about 22 h and then increased rapidly thereafter till completion
of the test. This compared exactly with the trend of rI in Fig. 12(a). From Figs. 12(a),
(b) and 13(c) three signicant occurrences during the SCC process are clearly indicated.
The rst one at about 5 h corresponded to the completion of the pitting process and beginning of the crack initiation process. The trough at about 22 h corresponded to the completion of the crack initiation process and the peak at about 30 h corresponded to completion
of the crack propagation stage.

T. Anita et al. / Corrosion Science 48 (2006) 26892710

2707

-140
0h
1.4 h

-160

Current, dB

2.8 h
4.0 h

-180

6.8 h

-200

20 h

-220
-240
-260
-4

-3

10

10

10

-2

10

-1

10

10

Log frequency, Hz

(a)
-120

24 h

-140

26.8 h

Current, dB

-160

30.9 h
35.1 h

-180

37.9 h

-200

39.3 h

-220
-240
-260
-280
-4

10

-3

10

10

-2

10

-1

Log frequency, Hz

(b)
-160

Current, dB

Current, dB

-170

-180

-190

-200
0

(c)

10

20

30

40

Time, h

Fig. 13. Current PSD plots of SCC in boiling acidic chloride solution (a) and (b) vs. frequency, and (c) vs. time in
the frequency-independent region shown in (a) and (b).

3.3. Comparison of detection of pitting corrosion and stress corrosion cracking by ECN
The present study has dealt with two situations, one in which pitting was observed in
the neutral deaerated NaCl and acidied NaCl solutions and the other in which SCC

2708

T. Anita et al. / Corrosion Science 48 (2006) 26892710

initiated from pits in boiling acidied NaCl solution. It was observed that the visual
records of potential and current very clearly demarcated between metastable and stable
pitting, and between the various stages of pitting, crack initiation and propagation during
SCC. Large current transients spread over longer time durations concurrent with active
potential transients typied stable pitting vis-a`-vis metastable pitting. During pitting the
current transients were of lower amplitude and much more closely spaced as compared
to the more widely spaced larger current transients for SCC. Alongwith visual records,
coecient of variation of current (CVC) clearly indicated early pit initiation and growth
in stressed material compared to unstressed material. SCC could be clearly demarcated
from pitting attack in tests in boiling acidic chloride solution, by visual records of potential and current, variation of rI, rV, RN and R0SN with respect to time.
4. Conclusions
Pitting corrosion studies on unstressed and stressed type 316 stainless steel in neutral
deaerated chloride and acidic chloride solution at room temperature were carried out
alongwith stress corrosion cracking studies on U-bend specimens in boiling acidic chloride
solution. The electrochemical noise was monitored during the course of the experiments.
The following were the important conclusions from the studies:
1. Pitting corrosion attack occurred much earlier on stressed specimens than on unstressed
specimens as observed from the visual records. Stable pitting could be dierentiated
from metastable pitting by sudden high current transients spread over larger time duration in case of the former.
2. Analysis of the EN signals during SCC tests clearly indicated the usefulness of the
potential and current visual records, alongwith rI, rV, RN and R0SN and PSD current
vs. time plots, in distinguishing between pitting and stress corrosion cracking. However,
the visual records of the potential alongwith plots of variation of rV, RN and R0SN with
respect to time were useful in clearly demarcating SCC initiation stage from SCC propagation stages. Current signals sharply rose and gradually decreased in stepwise fashion
during SCC.
3. The coecient of variation of current can detect the pitting corrosion attack. However,
the roll-o slope values were not found to be the true indicators of both pitting corrosion and stress corrosion and could not be used to dierentiate between the two processes. The visual records and the statistical parameters alongwith the current PSD
values were better indicators of the corrosion process than the roll-o slope.
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]

H.S. Khatak, V. Seetharaman, J.B. Gnanamoorthy, Practical Metallography 20 (1983) 570.


T.V. Vinoy, H. Shaikh, H.S. Khatak, J.B. Gnanamoorthy, B. Raj, Practical Metallography 34 (1997) 527.
R. Roberge, Journal of Applied Electrochemistry 23 (1993) 1223.
W.P. Iverson, Journal of Electrochemical Society 115 (1968) 617.
V.A. Tyagai, N.B. Lukyanchikova, Surface Science 12 (1968) 331.
V.A. Tyagai, Electrochimica Acta 18 (1973) 229.
V.A. Tyagai, Electrochimica Acta 16 (1971) 1647.
U. Bertocci, Journal of Electrochemical Society 128 (1981) 520.
A. Legat, V. Dolecek, Corrosion 51 (1995) 295.

T. Anita et al. / Corrosion Science 48 (2006) 26892710


[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]

[37]

[38]
[39]
[40]
[41]
[42]
[43]
[44]
[45]
[46]
[47]
[48]
[49]
[50]

2709

K. Hladky, J.L. Dawson, Corrosion Science 21 (1981) 317.


K. Hladky, J.L. Dawson, Corrosion Science 22 (1982) 231.
A.M.P. Simoes, M.G.S. Ferreira, British Corrosion Journal 22 (1987) 21.
J. Flis, J.L. Dawson, J. Gill, G.C. Wood, Corrosion Science 32 (1991) 877.
C. Monticelli, G. Brunoro, A. Firgnani, G. Trabanelli, Journal of Electrochemical Society 139 (1992) 706.
G. Gusmano, G. Montesperelli, E. Traversa, CORROSION/93, paper no. 355, National Association of
Corrosion Engineers, Houston, TX, 1993.
G. Gusmano, G. Montesperelli, A. DAmico, C. Di Natale, CORROSION/Asia, Singapore, National
Association of Corrosion Engineers, Houston, TX, 1994.
R.C. Newman, K. Sieradzki, Scripta Metallurgica 17 (1983) 621.
C.A. Loto, R.A. Cottis, Corrosion 43 (1987) 499.
C.A. Loto, R.A. Cottis, Corrosion 45 (1989) 136.
R.A. Cottis, C.A. Loto, Corrosion 46 (1990) 12.
D.B. Wells, J. Stewart, R. Davison, P.M. Scott, D.E. Williams, Corrosion Science 33 (1992) 39.
J. Stewart, D.B. Wells, P.M. Scott, D.E. Williams, Corrosion Science 33 (1992) 73.
D.A. Eden, A.N. Rothwell, J.L. Dawson, CORROSION/91, paper no. 444, National Association of
Corrosion Engineers, Houston, TX, 1991.
D.A. Eden, A.N. Rothwell, CORROSION/92, paper no. 292, National Association of Corrosion Engineers,
Houston, TX, 1992.
G. Bagley, Ph.D. Thesis, University of Manchester Institute of Science and Technology, Manchester,
England, 1998.
J.L. Luo, L.J. Qiao, Corrosion 55 (1999) 870.
H. Inoue, H. Iwawaki, K. Yamakawa, Materials Science A 198 (1995) 225.
K. Yamakawa, H. Inoue, Corrosion Science 31 (1990) 503.
R.A. Cottis, Corrosion 57 (2001) 265.
U. Bertossi, J. Frydman, C. Gabrieelli, F. Huet, M. Keddam, Journal of Electrochemical Society 145 (1998)
2780.
P. Bierwagen, Journal of Electrochemical Society 132 (1994) L155.
F. Mansfeld, C.C. Lee, G. Zhang, Electrochemica Acta 43 (1998) 435.
P.C. Pistorius, Corrosion 53 (1997) 273.
J.C. Uruchurtu, J.L. Dawson, Corrosion 43 (1987) 19.
G. Gonzalez-Rodriguez, V.M. Salinas-Bravo, E. Garcia-Ochoa, A. Diaz-Sanchez, Corrosion 53 (1997) 693.
G. Okamoto, in: R.W. Staehle, H. Okada (Eds.), Proceedings of the Conference on Passivity and Its
Breakdown on Iron and Iron Base Alloys, National Association of Corrosion Engineers, Houston, TX, 1976,
p. 106.
A.T. Fromhold, in: R.W. Staehle, H. Okada (Eds.), Proceedings of the Conference on Passivity and Its
Breakdown on Iron and Iron Base Alloys, National Association of Corrosion Engineers, Houston, TX, 1976,
p. 161.
J. Bessone, Ph.D. Thesis, University of Manchester Institute of Science and Technology, Manchester,
England, 1983.
P.C. Pistorius, G.T. Burstein, Philosophical Transactions of the Royal Society (London) A 41 (1992) 531.
G.S. Frankel, L. Stockert, F. Hunkler, H. Bohni, Corrosion 43 (1987) 429.
H.S. Isaacs, Corrosion Science 29 (1989) 313.
D.E. Williams, J. Stewart, P.H. Balkwill, Corrosion Science 36 (1994) 1213.
T. Schreiber, A. Schmitz, Physical Review E 55 (1997) 5443.
C. Boulleret, D. Gorse, B. Baroux, Corrosion Science 38 (1996) 999.
G. Gusmano, G. Montesperelli, S. Pacetti, A. Pettiti, A. DAmico, Corrosion 53 (1997) 860.
U. Bertocci, CORROSION/89, paper no. 24, National Association of Corrosion Engineers, Houston, TX,
1989.
I.A. Al-Zanki, J.S. Gill, J.L. Dawson, Materials Science Forum 8 (1986) 463.
N. Celati, C. Bataillon, Durability of Aluminium and Its Alloys in the Electrical Industries, Societe des
Electriciens et des Electroniciens, Paris, France, 1986, p. 117.
H.S. Khatak, J.B. Gnanamoorthy, P. Rodriguez, Metallurgical and Materials Transactions A 27A (1996)
1313.
C.N. Cao, H. Lin, X. Chang, Key Engineering Materials 2028 (1988) 257.

2710

T. Anita et al. / Corrosion Science 48 (2006) 26892710

[51] P.C. Searson, Ph.D. Thesis, University of Manchester Institute of Science and Technology, Manchester,
England, 1983.
[52] K. Hladky, L.M. Callow, J.L. Dawson, British Corrosion Journal 15 (1980) 20.
[53] K. Hladky, J.L. Dawson, Conference on Methods of Corrosion Testing and Research Manchester, England,
1982.
[54] E. Proverbio, S. DAmbrosio, V. Matres, Materials Science Forum 289292 (1998) 997.
[55] J. Uruchurtu Chavarin, Corrosion 47 (1991) 472.
[56] U. Bertocci, J.L. Mullen, Y.-X. Ye, in: Passivity of Metals and Semiconductors, Bombannes, France, 1983,
Elsevier Science Publishers, Amsterdam, 1983, p. 229.

You might also like