You are on page 1of 23

acom

1 - 2011

A corrosion management and applications engineering magazine from Outokumpu

Introduction
Stress corrosion cracking occurs as
a result of the conjoint action of
mechanical loading and a corrosive
environment. Typically conditions
in which this can occur for stainless
steels include chloride-containing
environments, caustic solutions
and cases where H2S is present, for
example in the oil and gas industry.
The stresses usually originate from
design load, but can also be a result
of local deformation or even residual
stresses around welds.
The first paper discusses use of
LDX 2101 within the oil and gas
industry. Although lean duplex
grades are not currently included
in the main industry standard
NACE MR0175 / ISO 15156 the
property profile of LDX 2101
indicates that it has good potential
for applications in both milder well
conditions and topside components.
The second paper in this issue
of Acom is a comprehensive
review of the performance of
Outokumpus duplex stainless
steel grades in chloride-containing
environments. It contains a summary of the standard tests which
have been performed on the range
of grades from the lean LDX 2101
to the superduplex 2507, and also a
discussion of the reasons for the
superior stress corrosion cracking
resistance of duplex steels compared
to the basic austenitic grades.

Lean Duplex Stainless Steel


within the Oil and Gas Industry
Stress Corrosion Resistance
of Duplex Grades

www.outokumpu.com

page 2

page 10

acom 1 - 2011

Lean Duplex Stainless


Steel within the Oil and
Gas Industry
Elisabeth Johansson, Rachel Pettersson, Outokumpu Stainless AB, Sweden

Summary
Lean duplex stainless steels, such as LDX 2101, have an attractive combination of high
mechanical strength and good corrosion properties, and can be candidates for use in oil
and gas applications. In this paper data are presented for lean duplex steels in various
types of tests and environments of relevance to such applications. They include not only
sour service conditions, but also environments which can give rise to chloride-induced
stress corrosion cracking or pitting. The results show that LDX 2101 has a resistance to
localised corrosion above austenitic grade 304L and superior resistance to stress corrosion
cracking compared to standard austenitic grades. Testing in H2S environments shows no
cracking of LDX 2101 at 90C and 0.15 bar H2S. However, as other types of corrosion can
occur, such as selective corrosion, it is important to investigate the limits for lean duplex
grades (PREN < 30) in regard to H2S, pH, chloride content and temperature so this
important group of alloys can be included in NACE MR0175 / ISO15156-3 in the future.
Keywords:

stainless, duplex, stress corrosion cracking, chloride, hydrogen sulphide,

oil and gas

1. Introduction
Duplex stainless steels are attractive materials in oil and gas application because of the
combination of high strength, good corrosion resistance, excellent resistance to chloride
induced stress corrosion cracking, and good weldability. They have successfully been used
as downhole tubing, topside equipment, subsea components, pipelines and umbilicals.
LDX 2101 is a lean duplex stainless steel with low nickel content. Manganese and nitrogen
are used to balance the duplex microstructure and give approximately equal amounts of
austenite and ferrite [1]. As a result of the duplex microstructure and the high nitrogen
content LDX 2101 has a high mechanical strength. This is combined with corrosion
resistance equal or better than 304 type standard austenitic grade for uniform, intergranular
and localised (pitting and crevice) corrosion [2]. It has also been shown that LDX 2101
has a resistance to stress corrosion cracking that is superior to that of standard austenitic
grades [3]. This type of steel can be a candidate for subsea oil and gas applications, such as
flowlines and umbilicals in milder well conditions, as well as for topside applications where
its combination of corrosion resistance and high mechanical strength can be utilized.
They can be an alternative to standard austenitic grades such as 304L, supermartensitic
stainless steels, 2205 duplex stainless steel and coated carbon steels.
In such applications one concern is the risk of cracking in environments containing H2S
(sour environments) and NACE MR0175 / ISO15156-3 provides application limits for
corrosion resistant alloys (CRA) intended for sour service in oil and gas production.
However, while standard austenitic grades are covered by this standard, duplex grades are
only included if their pitting resistant equivalent, PREN, value is 30 and the molybdenum
content 1.5 [4]. This means that duplex grades such as LDX 2101 and 2304 are not
covered by NACE MR0175 / ISO15156-3.
In this study corrosion properties, pitting corrosion and stress corrosion cracking,
of LDX 2101 will be discussed. Also results from testing in H2S environments will be
presented and compared with other reported work on LDX 2101 and 2304.

acom 1 - 2011

2. EXPERIMENTAL
2.1 Materials

In Table 1 duplex and standard austenitic grades are listed with their typical chemical
composition. Also shown is the Pitting Resistance Equivalent (PRE) value, which can
be used to rank stainless steel alloys according to their resistance to localized corrosion.
Several expressions exist for calculating the PRE value based on the influence of different
alloying elements. In the NACE MR0175 / ISO15156-3 standard the PRE-value is
defined as:
PREN = Cr + 3.3 (Mo + 0.5 W) + 16 N
It is important to remember that the calculated PRE value only gives an indication of
the resistance to pitting corrosion.
Table 1

Typical chemical composition and calculated PREN values of duplex and standard austenitic grades.

Typical chemical composition [wt%]


EN

ASTM/UNS

Outokumpu

Cr

Ni

Mo

Mn

PREN*

1.4162

S32101

LDX 2101

21.5

1.5

0.3

5.0

0.22

26

1.4362

S32304

2304

23.0

4.8

0.3

0.10

26

1.4462

S32205

2205

22.0

5.0

3.1

0.17

35

1.4307

304L

4307

18.1

8.1

18

1.4404

316L

4404

17.2

10.1

2.1

24

*PREN = Cr + 3.3 x (Mo + 0.5 x W) + 16 x N


Table 2

Typical mechanical properties for hot rolled plate.

Outokumpu

Rp0.2
[MPa]

Rm
[MPa]

A5
[%]

LDX 2101

480

700

38

2304

450

670

40

2205

510

750

35

4307

280

580

55

4404

280

570

55

2.2 Test procedures

Pitting corrosion

The resistance to pitting corrosion was tested using the ASTM G150 method to determine the critical pitting temperature (CPT) in 1 M NaCl. Measurements were carried
out on 60 x 60 mm specimens wet ground to 320 grit surface finish and mounted in a
crevice free flush port cell. A potential of 700 mVSCE is applied to the specimen and starting
at 0C the temperature is ramped with 1C/s while the current is monitored. The CPT is
determined where the current density exceeds 100 A/cm2 for 60 seconds. After testing,
visual examination using optical light microscopy with a 20x magnification is used to
confirm the existence of pits and absence of crevice corrosion.
The critical pitting temperature was also determined using the ASTM G48 Method E.
Specimens were dry ground 120 grit and exposed in an acidified ferric chloride solution
(6% FeCl3 + 1% HCl) for 24 hours. The temperature where the pitting occurs was determined
in 2.5C steps, which is more rigorous than the 5C step specified in the standard. Pitting
corrosion is considered to be present if the local attack is 0.025 mm or greater in depth.

acom 1 - 2011

Stress corrosion cracking (SCC)

Both evaporative and immersion test methods have been used for stress corrosion testing.
The ASTM C692 method uses insulation to conduct a chloride-containing solution to
the outside of a hot stressed specimen. Chlorides are carried along with the solution and
concentrated at the specimen surface by evaporation. The wick test is based on the
ASTM C692 Dana Test Procedure. From 50 x 200 mm coupons U-bend specimens
were bent to a 25.4 mm radius and with the legs parallel to within 1.6 mm. Holes were
drilled for tightening bolts and heating connectors. The specimen edges and the exposed
surface were dry ground to 120 grit. Using the yield strength for the specific material, the
leg deflection was adjusted to achieve the desired elastic stress of 100% of the actual yield
strength. The specimens were stressed parallel to the rolling direction. A bolt and nut (with
Teflon washers) were used to deflect the legs the required amount. A thermo-electric couple
was spot welded to the specimen in the centre of the concave surface for temperature control.
The specimen was fitted into the U-shaped groove of the Kaowool insulation material, which
was placed in a NaCl-solution with a concentration of 1500 ppm Cl-. By applying an AC
current the specimen was heated to 1006C. The level of the solution was controlled
and stabilised with an external solution container and a float switch. The test duration was
672 hours (28 days). After exposure, the specimens were first flattened and then re-bent to
roughly the original shape. This procedure loosens accumulated solids on the specimen
and aids the detection of any cracks.
For all the immersion test methods U-bend specimens were used. From 127 x 13 mm
coupons U-bend specimens were prepared using a two stage method. Holes with diameter
6.4 mm were drilled 13 mm from the short edges of the specimen to allow insertion of
the tightening bolt. The surface and edges were dry ground to 120 grit. A mandrel with
12 mm radius was used for the first stage of forming to an open U-shape. The final stage
was used to attain the correct U-shape with parallel legs, and a tightening bolt (Ni-base
or Ti alloy) with Teflon washers was used to maintain the load.
Immersion test was performed in 40% (by weight) calcium chloride solution at 100C.
The test solution consisted of 40% CaCl2 plus 1% calcium acetate and pH was adjusted
to 6 using calcium oxide. The test continues for 500 hours or until cracking is observed.
Throughout the test period air is bubbled through the test solution.
Two different tests were carried out in boiling 25% (by weight) sodium chloride solution.
The first one used a solution with unadjusted pH, while the second test was carried out
according to the ASTM G123 practice with pH adjusted to 1.5 using phosphoric acid.
In the second test a fresh test-solution was used for each one-week period and the final
pH for the solution after testing is determined at room temperature. A final pH over
about 2.5 suggests that uniform corrosion or pitting has occurred. The duration for both
tests was 1000 hours (~ 6 weeks) and the test temperature was the boiling point of the
test solution, measured to be 107C.
All specimens were examined for cracking using a binocular microscope at up to
20x magnification. If a crack is identified, this constitutes a failure.
Sulphide stress cracking (SSC)

4-point bend specimens were strain gauged in accordance with EFC 17 [5]. Both longitudinal
and transverse strains were monitored in order to establish the applied strain. The specimens
were deflected in the jigs until a strain equivalent to 90% of the proof strength at room
temperature was achieved. The gauges were left in position and the strain monitored for
up to one hour to detect any relaxation. If relaxation occurred the specimens were further
stressed to obtain the desired value.
The solution used was NACE TM0177-1 which consisted of 5% NaCl plus 0.5% acetic
acid with the pH adjusted to 3.5. The pH of the test solution was recorded before the
introduction of any gases. The test temperature was 90C. Following the 720 hours test
duration the pH of the test solution was measured again. Three different levels of H2S
were used, these were; 0.5 bar H2S, 0.1 bar H2S and 0.15 bar H2S in nitrogen.
After stressing duplicate 4-point bend samples they were placed in a glass culture
vessel. De-aerated test solution was introduced into the vessel to a level that completely

acom 1 - 2011

covered the bent beam jigs. The vessel was sealed and de-aeration with nitrogen was
continued, the solution being stirred to ensure even dispersion.
After 2 hours the nitrogen was replaced by the test gas mixture and the temperature of
the solution was slowly raised to 90C. The test time commenced when the test solution
was fully saturated with the test gas and the test temperature was stable at 90C. A water
cooler condenser was inserted in the lid of the culture vessel to prevent solution loss
due to evaporation. Solution temperature and level and gas flow were monitored over
the 30 day test period.
The specimens were examined for cracking using a binocular microscope at up to
40x magnification in the stressed condition (in the stressing jigs) and again in the unstressed
condition. The specimens were further examined by taking sections longitudinally through
the high stress portion of the beam. These were mounted in a cold set resin and ground
and polished and further examined.

3. RESULTS
In Figure 1 the results from pitting corrosion test according to ASTM G150 in 1 M NaCl
and ASTM G48 Method E in 6% FeCl3 + 1% HCl are shown. The standard austenitic
grade 4307 gives a CPT value below 10C using ASTM G150, which is the minimum
value that should be reported according to the standard. As shown in the diagram the
lean duplex grade LDX 2101 has a CPT value in between standard austenitic grades
4307 and 4404 for both test methods.
In Table 3 the results from the SCC tests are summarised as the number of cracked
specimens compared to the number of tested specimens. LDX 2101 shows superior
Fig. 1 CPT values determined according to ASTM G150 in 1 M NaCl and ASTM G48

Method E in 6% FeCl3 + 1% HCl.


60

ASTM G150
ASTM G48E

50

CPT (C)

40

30

20

10

<10

0
4307

4404

LDX 2110

2304

2205

Summary of stress corrosion cracking testing (number of cracked

Table 3

samples with number of tested samples in parenthesis).

Test method

4301

Wick

5 (5)

40% CaCl2

6 (6)

25% NaCl

ASTM G123

4 (4 )

4404
1)

LDX 2101

2205

4 (4)

0 (6)

0 (6)

1 (6)

0 (6)

0 (6)

0 (6)

0 (4)

0 (4)

0 (4)

0 (4)

2 (4)
2)

2304

4 (4)

2)

1)

Test terminated after 96 hours (4 days)

2)

Test terminated after 165 hours (1 week)

acom 1 - 2011

resistance to stress corrosion cracking compared to the standard austenitic grades 4301
and 4404. In the immersion tests the standard austenitic grades often exhibited cracking
in a short time.
After removal of salt deposits and corrosion products on the wick specimens some
degree of pitting was found on all tested materials. Stress corrosion cracking was detected
on the austenitic grade and on one single specimen of 2205, which may be anomalous.
In CaCl2 none of the duplex specimens showed any evidence of stress corrosion
cracks. However, a few cases of pitting were found on LDX 2101 and 2205 specimens.
The 4301 specimens cracked after only a few days exposure.
In 25% NaCl without adjusted pH LDX 2101 and 2205 did not show any signs
of cracking while two out of four 4404 specimens cracked. Those 4404 specimens that
showed no signs of stress corrosion cracks were thinner than the other tested materials
and were thus less severely stressed. Crevice corrosion was found on some of the 4404
and LDX 2101 specimens between washers and sample.
The acidified NaCl solution is more severe and all austenitic specimens cracked within
a week while no cracks were detected on the duplex grades. Evidence of crevice corrosion
between washers and sample were found on all tested grades with the most severe attack
in 4301. Also, attack on the specimen surface was found on 4301 and LDX 2101
specimens. It is difficult to compare the degree of surface attack between austenitic and
duplex grades as the exposure time was much longer for the duplex grades.
In Table 4 the results from the tests of LDX 2101 in H2S environments are summarised.
In none of the tested environments was any evidence of cracking found. However, examination of polished cross section revealed slight selective corrosion on both the stressed and

Results for 4-PB testing of LDX 2101 at 90% of proof strength,

Table 4

5% NaCl + 0.5% H3COOH, pH 3.5, 90C, 720 hours.

Environment
bar H2S

Start

0.05

3.50

3.76

No evidence of surface or subsurface cracking.

0.1

3.50

3.88

No evidence of surface or subsurface cracking.


Evidence of surface selective corrosion on both
the stressed and unstressed surface.

0.15

3.45

3.83

No evidence of surface or subsurface cracking.


Evidence of surface selective corrosion on both
the stressed and unstressed surface, slightly
deeper and more extensive than in 0.1 bar H2S.

Fig. 2

Solution pH
Finish

Results

Selective corrosion of the ferritic phase in LDX 2101 specimens tested in 0.15 bar H2S (left)
and 0.1 bar H2S (right).

acom 1 - 2011

unstressed surface of the specimens tested at 0.1 and 0.15 bar H2S. As Figure 2 illustrates
it is the continuous ferrite phase that has corroded in the duplex microstructure. In some
locations the selective attack had advanced along the phase boundary but this did not
appear to have developed into stress corrosion cracks.
Results from testing of welded LDX 2101 and a 13Cr super-

Table 5

martensitic steel, 4-PB at 100% of proof strength [6].

H2S
[bar]

Cl[%]

CO2

pH

Temp.

Results*

SMSS

0.01

16.5

Bal.

4.7

R.T.

6/6

LDSS

0.1

16.5

Bal.

4.7

R.T.

0/3

Grade
13Cr
LDX 2101

*Number of cracked specimens/total number of specimens.

4. DISCUSSION

Test conditions of carcass

Table 6

specimens [7].

Temperature

130C

pH2S

0.1 bar

pCO2

3.9 bar

Water pressure

2.7 bar

Total pressure

6.7 bar

Chloride content

50 g/l

pH

4.0

The presented results show that LDX 2101 has a resistance to stress corrosion cracking
that far exceeds that of standard austenitic grades. It may even be at the same level as
2304 and 2205 duplex stainless steels but this is difficult to judge because of the lack of
cracking in the duplex grades. However, LDX 2101 has a resistance to pitting which is
between grades 4301 and 4404, so there is some risk that other forms of corrosion can
occur. This is shown in the SSC test where levels of 0.1 bar or above can lead to selective
corrosion of the ferritic phase.
Some testing on LDX 2101 in H2S environments has been reported in the literature.
Welded LDX 2101 was included in a test where it was compared to 13Cr supermartensitic
stainless steel [6]. The result is shown in Table 5 and while 13Cr cracks at 0.01 bar H2S
none of the LDX 2101 samples cracked at 0.1 bar H2S. Compared to the current study
this one was performed at a lower temperature of R.T. and at a higher pH 4.7, while the
chloride content was higher at 16.5%.
A qualification test of LDX 2101 for use as carcass material in flexible pipes has been
reported by Gudme and Nielsen [7]. Carcass specimens of LDX 2101 and 316L were
tested for 360 and 1440 h. The test environment was designed to be close to the limit for
cracking/no-cracking and is detailed in Table 6.
The results showed that there were no signs of cracking on any of the specimens.
However, discoloured areas were found on both 316L and LDX 2101. No signs of
preferential attack were found on the LDX 2101 samples. It was claimed that in a prior
test in a very similar environment but at a H2S content of 0.3 bar, selective corrosion
of the austenitic phase was found in discoloured areas.
Like LDX 2101, the duplex grade 2304 has a PRE value below 30 and thus falls
outside the current specifications in NACE MR0175 / ISO15156-3. Coudreuse et al
[8] have reported tests on 2304 specimens at a constant load of 90100% of the yield
strength for 720 hours and under test conditions according to Table 7.
Table 7

Test conditions for 2304 [8].

NACE
TM0177

EFC 17
formation
water

Field
formation
water

Formation
water

Temperature [C]

25

80

25 and 80

25 and 80

pH2S [bar]

0.05, 0.5 and 1

0.04

0.01

pCO2 [bar]

20

NaCl [%]

16.5

16.5

25

2.7

4.5

4.5 - 5

pH

acom 1 - 2011

No failure or cracking was observed in any of the test conditions. However some selective
corrosion of the ferritic phase were found, with more severe attack found on specimens
exposed to the EFC 17 formation water with 0.5 and 1 bar H2S. The results for 2304
were compared to supermartensitic grades developed for flowlines. The threshold stress
for SSC was compared and the highest threshold stress was found for 2304 (450 MPa)
even though 2304 has lower yield strength. This was explained by the observation that
no cracking occurred up to 100% of the yield strength for 2304. It was suggested that
2304 could be used up to at least 0.05 bar H2S and probably also for higher H2S levels.
In Figure 3 a summary is given of SSC results for LDX 2101 and 2304 from the current
study and the studies cited above. It shows that no SSC has been observed at H2S levels
up to 0.15 bar for LDX 2101 and 1 bar for 2304. It is also shown that selective corrosion
can occur over a wide range of test conditions as other factors such as temperature, pH and
chloride content will influence whether selective corrosion will occur or not.
Fig. 3

Summary of SSC tests performed on LDX 2101 and 2304 [6, 7, 8].
SSC = sulphide stress cracking, SC = selective corrosion.

10

LDX 2101 No SSC, No SC


LDX 2101 No SSC, SC

1
pH2S (bar)

2304, No SSC, SC

0,1

0,01

0,001
1000

10000

100000

1000000

Chloride content (ppm)

5. CONCLUSIONS
The lean duplex grade LDX 2101 has a property profile which makes it appropriate for
application within the oil and gas industry, in both milder well conditions and topside
components. The results show that LDX 2101 has a resistance to localised corrosion
above austenitic grade 4301 and superior resistance to stress corrosion cracking compared
to standard austenitic grades. Testing in H2S environment shows no cracking of LDX
2101 90C and 0.15 bar H2S. However, as other types of corrosion can occur, such as
selective corrosion, it is important to investigate the limits for lean duplex grades (PREN
< 30) in regard to H2S, pH, chloride content and temperature so this important group
of alloys can be included in NACE MR0175 / ISO15156-3 in the future.

acom 1 - 2011

6. REFERENCES
[1] P. Johansson and M. Liljas, A New Lean Duplex Stainless Steel for
Construction Purposes, 4th European Stainless Steel Science and Market
Congress, p. 153157. (Paris, France: ATS, 2002).
[2] A. Bergquist, A. Iversen and R. Qvarfort, Corrosion Properties of UNS S32101
A New Duplex Stainless Steel with Low Nickel Content Tested For Use As
Reinforcement in Concrete. Corrosion 2005, paper no. 05260, (San Diego,
CA: NACE International, 2005).
[3] E. Johansson and T. Prosek, Stress Corrosion Cracking properties of UNS S32101
A New Duplex Stainless Steel with Low Nickel Content. Corrosion 2007,
paper no. 07475, (Nashville, TN: NACE International, 2007).
[4] NACE MR0175/ISO 15156-3:2003, Petroleum and natural gas industries
Materials for use in H2S containing environments in oil and gas production
Part 3: Cracking-resistant CRAs (corrosion-resistant alloys) and other alloys.
[5] Corrosion Resistant Alloys for Oil and Gas Production: Guidence on General
Requirements and Test Methods for H2S Service, European Federation of
Corrosion Publications Number 17, 2002.
[6] Stainless Steels in Oil and Gas Production, Outokumpu Corrosion Handbook,
Tenth edition, 2009.
[7] J. Gudme and T. S. Nielsen, Qualification of lean duplex grade LDX 2101
(UNS S32101) for carcass material in flexible pipes. Corrosion 2009,
Paper no. 09075, (Atlanta, GA: NACE International, 2009).
[8] L. Coudreuse, V. Ligier, J. P. Audourd and P. Soulignac, Lean duplex stainless
steel for oil and gas applications, Corrosion 2003, Paper no. 03529, (Houston,
TX: NACE International, 2003).

Originally published at EUROCORR 2010


and reproduced with permission.

acom 1 - 2011

10

Stress Corrosion
Resistance
of Duplex Grades
Rachel Pettersson and Elisabeth Johansson, Outokumpu Stainless AB

Abstract
One of the major reasons for the success of duplex stainless steels is that they show
appreciably better resistance to chloride-induced stress corrosion cracking than their
austenitic counterparts. This offers significant design advantages, particularly when
combined with the higher strength attainable with duplex grades. In this paper results
from a variety of sources are surveyed to explore the limiting conditions for stress corrosion
cracking in duplex and austenitic stainless steels. The question of the different response
of the two steel groups to static or dynamic loading is discussed and illustrated using
examples. Finally, the role of the two phase duplex microstructure in increasing the
resistance to stress corrosion cracking is examined.
Keywords:

stainless, duplex, stress corrosion cracking, chloride, mechanisms.

Introduction
Stress Corrosion Cracking (SCC) occurs under the conjoint action of three critical factors
stress, a susceptible material and a corrosive environment. For stainless steels the unfortunate
situation is that the ubiquitous chloride ion, found in seawater, cooling systems, make-up
water, chemical process environments, and even the kitchen sink and the human body, is
one on the main causes of stress corrosion cracking. The stresses involved can be design
loads, but also the result of thermal expansion, residual stresses or localised deformation.
One 1983 review of almost a thousand failure cases of 304 stainless steel in the chemical
process industry attributed 37% of these failures to stress corrosion cracking so the problem
is one which must be taken most seriously.
To the non-corrosionist stress corrosion cracking is the most bewildering of all the types
of corrosion encountered. Uniform corrosion can be dealt with using corrosion tables, work
that began in the 1930s and has been extended by recent publication [1]. Pitting corrosion
is often related to the composition using the empirical tool of the Pitting Resistance Equivalent
(PRE) and has well-accepted tests in the form of critical pitting temperatures, in NaCl (as
specified in ASTM G150 or ISO 17864), or in FeCl3 (ASTM G48). Crevice corrosion is
more complicated, since it is critically dependent on the geometry of the crevice and can
require an appreciable length of time for the stable crevice chemistry to develop but again
ASTM G48 has proved a good starting point for evaluation of materials. However, this is
simplicity itself compared to the plethora of test environments and loading methods
which can be used to evaluate stress corrosion cracking.
Firstly, there are a number of standardised environments for testing chloride-induced
stress corrosion cracking. The classic environment of ~45% MgCl2, boiling at 155C and
standardised in ASTM G36, has the considerable advantage that it causes cracking, but
has received much criticism over the years for being harsher towards molybdenum-alloyed
grades than is indicated by practical experience. At the other extreme, ASTM G123 specifies
an environment of 25% NaCl at pH 1.5 and a boiling point of ~108C, which is suitable
only for the lower-alloyed steel grades. Wick testing, described in ASTM C692 as a method
for testing insulation material, has creatively been adapted as a test for stress corrosion cracking
under evaporative conditions.
In addition, there is a range of loading methods which can be used. These are well

11

acom 1 - 2011

Fig. 1 Schematic illustration of the load


on the remaining section during a stress
corrosion cracking test, modified from [2].
Rm

Load on remaining section

Slow strain rate

Constant
load

Load

Bend
testing
Time

Load
Residual compressive stress

described, both in the numerous parts of ISO 7539 and in ASTM standards G30 (U-bend)
G38 (C-ring), G39 (bent beam), G49 (constant load), G129 (slow stain rate), G186
(precracked double beam) and E1681 (threshold stress intensity factor determination).
The loading method is critical to the outcome of the test and, as will be discussed in the
following sections, gives rise to large differences in relative behaviour. The conceptually
simplest type of test is constant load, using a hanging weight which means that the load
on the remaining section increases as soon as a crack begins to decrease the effective cross
section area, as illustrated in Figure 1. Constant deflection testing, on the other hand, gives
rise to a decreasing load due to stress relaxation even before cracking initiates. Slow strain rate
testing in its simplest form involves monotonically increasing stress under dynamic loading,
which means that there is a greater degree of dislocation motion involved. Compressive
residual stresses are a particular pitfall because they may mean that the actual applied load is
far below the intended level, and mean that surface preparation can be critical for SCC testing.
The total number of permutations of environment and loading method is vast, and
further extended by the inclusion of other chlorides, particularly the widely-used CaCl2,
specific environments to simulate service conditions and field testing. A comprehensive
review of all available data on chloride-induced stress corrosion cracking of duplex stainless
steels would require many hundreds of pages, if not a separate conference, so the scope of
this paper must by necessity be limited to the more common types of testing and works in
which a wide range of steel grades have been evaluated.

Grades and properties


Crack initiation

Fig. 2

The appearance of duplex stainless steels on the industrial scene in the 1970s was hailed as
a solution to the stress corrosion cracking problem and there are numerous examples of
successful replacement of austenitics by duplex grades [ 35 ] Extensive testing programmes have demonstrated the superiority of the two phase microstructure and a number of examples are given in the following section. In order to put results in perspective, a
fruitful approach is to compare the duplex spectrum of alloys, ranging from the lowmolybdenum 2304 to the superduplex 2507, with the corresponding range of austenitic
grades from the all-purpose 304(L) to the 6% Mo grade 254 SMO. Also included is a
newer range of LDX alloys, characterised by a higher strength and more economical use
of nickel, which shows such strong price volatility. Figure 2 shows the typical compositions and properties of the grades and illustrates the point that for each austenitic steel
there is a close duplex equivalent with a similar corrosion resistance but higher strength.

Spectrum of duplex and austenitic grades referred to in this paper. The critical pitting temperatures are in 1M NaCl at
+700mVSCE according to ASTM G150.
Cr

Mo

Ni

LDX 2101

21.5

0.3

1.5

0.22

5 Mn

LDX 2404

24

1.6

3.6

0.27

3 Mn

2304

23

0.3

4.8

0.10

2205

22

3.1

5.7

0.17

2507

27

0.27

304/304L

18.1

316/316L

17.2

2.1

10.1

317LMN

17.3

4.1

13.7

904L

20

4.5

25

254 SMO

20

6.1

18

Yield strength (ASTM min, sheet) [MPa]

750
Duplex
Austenitic
LDX
2101

500

LDX
2404

2507
2205

2304

317
LMN

250
304

316

20

254
SMO
904L

40
60
Typical CPT (C)

80

100

8.1
0.14
0.20

12

acom 1 - 2011

Duplex performance in standardised tests


Sodium chloride environments

As indicated in the Introduction, sodium chloride has its limitations as a medium for stress
corrosion cracking testing, because the limited solubility gives a maximum test temperature
of 108C at ambient pressure. Results of U-bend testing are shown in Table 1 and demonstrate that the standard austenitic grades 304 and 316 are susceptible in this test, while
none of the duplex grades tested show cracking.
Performance of U-bend specimens in boiling chloride environments.

Table 1

Number of cracked samples with number of tested samples in parentheses [6].

Test environment

304

316

904L

254
SMO

LDX
2101

LDX
2404

2304

2205

25% NaCl.

2 (4)

0 (4)

0 (12)

0 (4)

(ASTM G123).

4(4)

4 (4)

0 (4)

0 (12)

0 (4)

~45% MgCl2 (ASTM G36)

3 (3)

3 (3)

0 (3)

3(3)

3 (3)

6 (6)

3 (3)

3 (3)

Acidified 25% NaCl

There is a possibility to extend the temperature range for testing in NaCl by autoclave
testing, as illustrated in Figure 3. The lines in this diagram are the result of extensive constant
load testing by Sandvik Materials Technology [7] and show the limiting conditions causing
cracking, with the line shifting to higher temperatures and chloride concentrations with a
higher alloying level in the steel. Testing with U-bend specimens [8] in the same laboratory,
Figure 4, gave reasonably consistent data for the duplex grades but indicated that the
cracking limit for the austenitic grades was appreciably higher than in the constant load
tests. It also exceeded the limit from practical experience, which indicates a maximum
service temperature of 40 60C. This dependence on loading method is a recurrent theme
in stress corrosion cracking. It has been explored in detail in [9], where the surprisingly
good performance of austenitic grades in U-bend tests was attributed to their higher degree
of relaxation, ~30% compared to ~10% for the duplex grades.

Fig. 3

Compilation of cracking limits for various steel grades


based on constant load testing at the proof stress and
practical experience [7].



300

200
2205
150

2304

100
304/316

50

-5

-4

-3

-2

-1

Log chloride concentration (%)

0.10%
1%

250
Temperature (C)

Temperature (C)

250

Lowest temperature at which cracking was


observed in autoclave testing of U-bend
specimens in NaCl [8].

300

2507
No cracking

904L

Fig. 4

200
150
100
50

30

4L

90

25

M
4S

04

23

05

22

07

25

acom 1 - 2011

13

Magnesium chloride environments


Magnesium chloride has a much broader range of usefulness, since the boiling point extends
to above 160C at ambient pressure. U-bend testing, included in Table 1, gives failure of
all tested grades apart from the high-Ni 904L within 24 hours. Molybdenum-alloying
appears to decrease the resistance to stress corrosion cracking in this environment for
example, the constant load testing in Figure 5 indicated a lower threshold stress for 316
than for 304, also lower for 2205 compared to 2304 in aerated 44% MgCl2 at 150C [10].
One possible reason for unusual behaviour in saturated magnesium chloride solution,
discussed in [1112], is the pH shows a rapid drop when the concentration approaches
saturated and this environment gives active corrosion of stainless steels, sometimes even
accompanied by hydrogen evolution, Figure 6. This offers the explanation that a lower
alloyed steel may show a reduced propensity to cracking because potential initiation sites
are cathodically protected or blunted by the active corrosion of the surface.
An interesting and relevant modification of the MgCl2 environment was proposed in
[13] and involves creating an environment to simulate that formed when seawater is concentrated. Precipitation of sodium chloride means that the solution becomes predominantly
magnesium chloride. Such an environment with 487g/l MgCl2, a total chloride content of
10.7M and a boiling point of 125C was used for U-bend testing by Usinor Industeel
(now Arcelor Mittal) within the framework of [8]. The results in Figure 7 show a logical
dependence on alloying level for the austenitic grades, but interestingly a higher threshold for 2304 than either 2205 or 2507. In this work it was noted that 2304 exhibited a
high level of uniform corrosion, which ties in with the previous line of argument that
uniform or selective corrosion may immunize the steel against developing cracks.

Fig. 5

Fig. 6 Polarisation curve for 2507 in saturated


MgCl2 showing active corrosion peak [11].

Results of constant load testing in MgCl2 [10].

160

600

140

400

304/
304L

300
200

2205
2304

316/
316L

100
0

Current (mA/cm2)

Time to fracture (h)

500

20C
47C

120
100
80
60
40
20

0.2

0.4

0.6

0.8

0.0

0.2

Fig. 7

Fig. 8

Lowest temperature at which cracking was


observed for U-bend specimens tested in
simulated evaporated seawater [8].

0.8

1.0

100

100
Stress rupture (%Rp0.2)

Critical temperature (C)

0.6

Lowest stress at which stress corrosion cracking


was observed in the drop evaporation test, related
to the yield stress at 200C [6,18].

120

80
60
40
20
0

0.4

Potential (VSCE)

Stress/Tensile strength at 150C

304

316

904

230

220

250

80
60
40
20
0

304

316

L
904 SMO 2101 2404
4
X
X
25
LD
LD

230

220

250

acom 1 - 2011

14

Evaporative conditions
Three types of methods have been used fairly extensively to characterise the stress corrosion
cracking of duplex stainless steels in conditions in which evaporation causes local concentration
of chloride. The wick test is based on ASTM C692 and uses insulation to conduct a NaCl
solution with 1500 ppm Cl- to the outside of a resistance-heated U-bend specimen with
an outer fibre stress of 100% of the actual yield strength. The chloride droplet test, based
on the method developed in [14] and reported in [15 , 16], involves depositing drops of
saturated MgCl2 or CaCl2 solution on the top of U-bend specimens which are placed in
desiccators with respective saturated salt solutions to keep the relative humidity at the
deliquescence point. The Drop Evaporation test, developed in [17], described in detail in
[18] and standardised as MTI-5 [19] simulates the situation which can occur on heat
transfer surfaces, where material is repeatedly wetted and dried, leading to a build-up
of chloride. Dilute (0.1M) sodium chloride solution is dripped onto uniaxially loaded,
electrically heated specimens such that one drop is evaporated just before the next one
falls on the specimen. This gives a specimen temperature which fluctuates in the range
80 120C.
The wick and droplet test results in Table 2 both show good stress corrosion cracking
for the entire spectrum of duplex steels, and thus do not differentiate between the duplex
alloying levels. However, the fact that load is used as a variable in DET gives greater
possibilities of differentiation, as is seen in Figure 8. Both the standard austenitic grades
304 and 316 are highly susceptible to cracking, even at loads of only 10% Rp0.2 while all duplex
grades show higher limiting stresses of 40% Rp0.2. The DET results echo to some extent
those described for MgCl2 in the previous section in that the lower alloyed LDX 2101
and 2304 show higher cracking thresholds but some degree of selective corrosion. It should
be noted in the context of such constant load testing that, as illustrated in Figure 1, the
yield stress of the duplex grades are a factor or two or more higher than that of standard
austenitic grades. This opens up the possibility to two types of comparison. Firstly the
materials can be compared at the same relative stress level, usually normalised to the yield
stress, as shown in Figure 8. This is the most usual form of comparison and means that
the actual load applied to the duplex steel is higher. This may seem unfair but actually
reflects the true loading situation if the higher strength of a duplex steel is utilised to give
a thinner gauge in a construction. Alternatively the absolute stress levels can be compared,
in which case the actual stress corrosion cracking resistance of the duplex grades at a given
load is emphasised.
Results of two different types of tests under evaporative conditions [15, 16, 18].

Table 2

For the wick test the duration was 28 days while for the droplet tests it was 22 weeks.

Test method

304

316

904L

254
SMO

LDX
2101

LDX
2404

2304

2205

2507

Wick, 100C

5 (5)

4 (4)

1*(4)

0 (3)

0 (6)

0 (7)

0 (6)

1*(6)

MgCl2 droplet, 50C

2 (2)

2 (2)

0 (2)

0 (2)

0 (2)

0 (2)

0 (2)

0 (2)

CaCl2 droplet, 50C

2 (2)

2 (2)

0 (2)

0 (2)

0 (2)

0 (2)

0 (2)

0 (2)

* Minor cracking, interpretation complicated by the occurrence of pitting and /or selective corrosion.

Other concentrated chlorides


Both CaCl2 and LiCl have been used as test media for stress corrosion cracking, with the
aim of attaining higher temperatures than possible in NaCl while avoiding the serious
disadvantages of the anomalous ranking and active corrosion behaviour in MgCl2. Results
for the three duplex grades 2304, 2205 and 2507, plus a series of tie line alloys, designed
to have varying phase ratios but the same composition in each phase as the austenite and
ferrite phases of 2205, were first presented at this conference 19 years ago [20]. The slow
strain rate curves in Figure 9 underline the point made in the previous section of the

acom 1 - 2011

Fig. 9 Slow strain rate curves for 2205, also model


alloys with the same composition as the austenite
and ferrite phase [20].

Fig. 10 Normalised SSRT results in deaerated 50% LiCl

and aerated 40% CaCl2 at 100C [20].

700
Normalised failure elongation (%)

100

600

Stress (MPa)

500
400
300
Austenite
Duplex
Ferrite

200
100
0

15

10

20

30

90
80
70
60
50
40
30
10
0

40

LiCl
CaCl2

20

20

40

Strain (%)

60

80

100

120

Austenite content (%)

usefulness of normalisation in order to compare alloys with rather different mechanical


properties. The yield strength of the ferrite phase alloy is more than twice that of the austenite
phase alloy, and the elongation less than half. However, even if such normalisation is carried
out, the SCC results give a confusing picture. Both the normalised failure elongation and
the time to failure of U-bend specimens showed a shifting dependence on both alloying
level and austenite content in the different media tested. This is illustrated in Figure 10,
which shows a switch in the relative behaviour of the austenite and ferrite depending on
the test environment: ferrite was more seriously affected than austenite in CaCl2 but the
reverse was seen in LiCl. With hindsight the explanation and conclusion is that evaluation
of the extent of cracking in this manner is not helpful. It contains components of both
initiation and propagation which may balance in slightly different ways depending on the
number of cracks and the local chemistry.
A more fruitful approach is that subsequently used in [8] in which the speed or extent
of cracking is ignored, and limiting conditions instead used to define the border between
cracking and no cracking. The resultant stress corrosion maps in Figure 11 show much
more consistency. There are a few cases in which cracking occurs in LiCl but not in CaCl2
Fig. 11 Figure 11. Stress corrosion cracking maps showing the areas in which cracking occurs in U-bend and slow strain
rate tests as a function of temperature and chloride concentration [8].
CaCl2
316L

2304

120

100

100

80

80

120

904L

Temperature (C)

Temperature (C)

120

LiCl

2205

100
80

120

254 SMO

2507

120

120

80

80
10

12

14

10

12

14

Chloride concentration (M)


U-bend: SCC/no SCC

904L

2205

254 SMO

2507

80

100

2304

100

100

316L

10

12

14

Chloride concentration (M)


SSRT: SCC/no SCC

10

12

14

acom 1 - 2011

16

but the general results are in agreement. The poor resistance of 316L to cracking is seen
in the low limit lines, with cracking occurring at 80C and even down to the relatively low
concentration of 6M chloride. The higher alloyed austenitic grades 904L and 254 SMO
showed no cracking at 80C or 100C respectively. The picture for the three duplex grades
was not as clearly dependent on alloying level. All three grades, 2304, 2205 and 2507
showed some instances of cracking at 80C if the chloride level was sufficiently high.
The clear differentiation seen in dilute chlorides (e.g. Figure 3) was not apparent.
This raises the interesting question of why such high chloride levels are required to
cause cracking. Why does 8M chloride at 80C give no SCC for 2205, while cracking
occurs if the concentration is increased to 12M? How many chloride ions are needed to
cause a breakdown in the passive film and maintain a sharp propagating crack? There has
been much discussion over the years on the mechanisms for stress corrosion cracking,
including the concepts of anodically-dominated slip dissolution, surface-dominated cleavage
or hydrogen effects. However, to the best knowledge of the authors, there is no satisfactory
explanation for necessity of such very high chloride concentrations to cause cracking.
A second point clearly seen in the stress corrosion maps is the different response of the
duplex and austenitic grades to the two types of loading. Slow strain rate testing gives
markedly lower limits than the U-bend (constant deflection) testing, particularly for 2304.
The opposite is seen for 316L. The reason why the duplex grades are particularly sensitive
to slow strain rate testing is still open to debate, but part of the explanation seems to be that
there is no scope for either stress relaxation or creep, which would be to the advantage of
duplex materials.

Limiting conditions design diagrams


The concept of the limiting conditions for pitting, and a comparison with the same type
of limits for pitting and crevice corrosion has been the subject of work within a recentlyconcluded European project [12]. The idea of the design diagram is based on the concept
of taking the lowest limit- whether it be from constant load, slow strain rate, or constant
deflection testing, and using this to define the risk area for stress corrosion cracking. In view
of the different response of austenitic and duplex grades to the latter two types of loading,
both need to be included. In a construction both decreasing load due to stress relaxation
and dynamic loading may occur. An example of such a diagram for stress corrosion cracking
in 2507 is seen in Figure 12. Interestingly, constant load appears to be the most conservative
method in this case, in contrast to the situation seen for the same grade in dilute environments
in Figure 3 and Figure 4.
The limits seen in Figure 12 agree well with those from Figure 11 and demonstrate the
same abrupt shift from immunity to cracking when the chloride level becomes sufficiently
high. No cracking was seen in any of tests in 5M chloride, but many instances in 7.5M.
Fig. 12 Compilation of stress corrosion cracking data for 2507 from [12].
200

2507

Temperature (C)

150
U-bend
100

SSRT
CL

50

0
0

10
Chloride content (M)

No cracking Ubend NaCl


No cracking Ubend MgCl2
No cracking SSRT MgCl2
No cracking CL MgCl2

No cracking CL NaCl
Cracking Ubend MgCl2
Cracking SSRT MgCl2
Cracking CL MgCl2

15

acom 1 - 2011

17

Role of microstructure and phase balance


The critical question to be answered here is Why are duplex stainless steels so resistant to
stress corrosion cracking? Much of the answer lies in the microstructure and the way in
which a crack can propagate in a complex two-phase microstructure. Figure 13 shows what
may be regarded as a typical stress corrosion crack in a duplex steel with transgranular,
sometimes branched, propagation in ferrite, which is the continuous phase. In places the
crack seems to avoid the austenite, passing instead through the phase boundary and in places
there is transgranular cracking in the austenite too.
One of the simplest explanations is that the two component phases of duplex stainless
steels show completely different slip behaviour by virtue of their crystal structure. Austenite, as
an fcc material, slips on {111}<110>, while the bcc ferrite shows slip primarily on {110}<111>
(also on {123}<111> and {112}<112>) and requires thermal activation of some systems.
Regardless of whether an anodic or cathodic (hydrogen-related) mechanism at the crack
tip is invoked to explain propagation of a stress corrosion crack, dislocation motion is a
vital component. Cracking in a single phase material can progress rapidly across a single
grain, then requires a build-up of stress at the grain boundary to initiate a crack in the
next grain. In duplex steels the situation is even more difficult because of the different slip
characteristics as soon as the crack crosses the phase boundary. This is basically the keying
effect first invoked by Flowers et al [21] in 1963 to explain the process whereby a crack
in the dominant austenite phase was mechanically stopped or diverted by the presence of
a ferrite island in cast duplex stainless steel. As cited in [22], it was later elaborated by Cottis
and Newman [23] that for a crack in the ferrite to cross to the austenite, it is also necessary
that the prevailing environmental conditions (potential, pH) are favourable to the promotion
of cracking in both phases.
Consequences of such a mechanical blocking effect are that cracking is dependent on
both the phase ratio and the grain size/phase separation. An illustrative example of an
unfavourable phase ratio is shown in Figure 14, where a duplex stainless steel has been
deliberately welded with a Ni-overalloyed filler to give a predominantly austenitic weld.
Testing in dilute chloride at elevated pressure gave rise to extensive cracking in the austenitic
weld, which halted abruptly at the fusion line where the ferrite level was higher. The NORSOK
specification for duplex piping [24] place strict requirements on the phase ratio, 3555%
ferrite in the base material and 3565% in welds, and these types of limits are also echoed
in user specification for many applications. The phase separation, commonly given as the
austenite spacing, is also critical because it is directly related to the mean free path for a crack
to propagate within a single phase. This has been demonstrated for H2S environments,
where cracking is related to hydrogen embrittlement of the ferrite phase [25, 26] and in
MgCl2 [27]. In the latter case it was found that the stress corrosion cracking resistance
of a 25Cr-6Ni-2Mo duplex steel could decrease to the level of 304 if heat treated to give
microstructural coarsening.

Fig. 13 Typical stress corrosion crack in 2205, showing

Fig. 14 Stress corrosion cracks in a weld in 2205 deliberately

propagation in the ferrite phase (etched dark)


and in the austenite-ferrite phase boundary [20].

10 m

overalloyed with 14.8% Ni to give approximately


75% austenite in the weld metal [28].

200 m

acom 1 - 2011

18

The idea that other types of interaction between the austenite and ferrite phases could
also contribute to the impressive resistance of duplex stainless steels to stress corrosion
cracking was first put forward by Hochmann et al [29] and Desestret et al [30] in 1973.
Figure 15 shows results n which constant load testing was carried out two concentrated
chloride environments on both the duplex stainless steel Uranus 50 and single phase
alloys corresponding to the separate phases. The duplex material performed better than
either of the single phase materials at low stresses but tended towards that of the austenite
at higher stress levels. Similar results were seen for Kowaka et al [31] and Nagano et al
[32] for a 25Cr 6Ni duplex steel, illustrated in Figure 16. In these works, fracture
mechanics investigation also showed the duplex steel to have a higher threshold stress
intensity factor and lower propagation rate than the single phase alloys.
These data show an important point about which there is often a misapprehension: the
ferritic steels are by no means resistant to stress corrosion cracking. The data in Figure 16
show that failure of the ferritic and austenitic steels occurs at approximately the same
absolute stress level. If the stress is instead normalised to the yield stress, which is higher
for ferritic grades as illustrated in Figure 9, then the ferritic steel is actually more susceptible.
Only for the duplex grades is it possible to utilise the higher yield strength to decrease
the gauge thickness and nevertheless retain good resistance to stress corrosion cracking.
The concept that there may be an electrochemical interaction between the two component
phases of duplex stainless steels was suggested by both Hochmann et al [29] and Shimodaira
et al [33] at the same conference in 1973. Electrochemical investigations of different types
in both these works indicated that the austenite was the more noble phase, with a corrosion
Fig. 15 Failure times as a function of relative stress level for Uranus 50 (21Cr 8Ni 2Mo 1.5Cu) plus austenite phase

and ferrite phase alloys in 44% MgCl2 at 153C [29] and in LiCl at 150C [30].
1000
MgCl2
Ferrite
Austenite
Duplex

100

Time to failure (h)

Time to failure (h)

1000

10

LiCl
Ferrite
Austenite
Duplex

100

10

0
0.0

0.5

1.0

1.5

0.0

0.5

Relative stress (R/R p0.2)

1.0

1.5

Relative stress (R/R p0.2)

Fig. 16 Failure times as a function of the total applied stress or relative stress for a 25Cr-6Ni stainless steel

and single phase alloys corresponding to the two separate phases [31].
100000

100000

Time to failure (min)

Time to failure (min)

28Cr-4Ni(f)
21Cr-9Ni(a)
25Cr-6Ni(d)

10000
1000
100
10
1

28Cr-4Ni(f)
21Cr-9Ni(a)
25Cr-6Ni(d)

1000

10

1
0

100

200

300

Applied stress (MPa)

400

500

0.0

0.5

1.0
Relative stress (R/R p0.2)

1.5

2.0

acom 1 - 2011

19

potential at least 10 mV and in cases as much as 50mV higher than that of the ferrite.
The line of reasoning put forward in [29] is illustrated schematically in Figure 17. The
basic concept is that at low stresses deformation only occurs in the austenite, but this is
cathodically protected by the ferrite. Only at higher stresses does the ferritic steel deform
and suffer stress corrosion cracking. In later work Magnin et al [34, 35] proposed that
cracking of the ferrite was related to the occurrence of mechanical twinning, which could
even occur considerably below the yield stress. The latter of these two papers elegantly
demonstrated how fatigue cracks initiated in the austenite phase at low plastic strain
amplitudes but in twins in the ferrite at higher strain amplitudes.
A further concept, and in hindsight a very obvious one, is that of the residual stress
state in duplex stainless steels. The higher thermal expansion coefficient of austenitic
stainless steels compared to ferritic grades means that when duplex alloys are quenched
from the solution annealing temperature, the austenite will be in tension and the ferrite
in compression. This point was made by Kowaka et al [31] as illustrated in Figure 18.
Some very precise x-ray diffraction work to investigate the residual stress state in 2304
duplex stainless steel has been done by Moverare et al [36] and is illustrated in Figure 19.
The compressive stress in the ferrite was more than 100 MPa, with the consequence that
even a plastic strain of 1.5% the tensile stress in the ferrite was still below 500 MPa.
The conclusion is that duplex stainless steels have a build-in resistance to cracking because
the matrix phase is in compression and higher stresses will be required to put it in tension
and cause cracking.
Fig. 17 Cathodic protection of austenite

at low load means that cracking


occurs only when the ferrite
deforms at higher loads [29].

~10A/cm

~5A/cm

Cathode

resistance of duplex stainless steels [31].

~Rp

<<Rp

Fig. 18 Illustration of various factors contributing to the high stress corrosion

+ tension
- compression

Cathode

cathode

Cracks

anode

(a) cathodic
protection



low
stress

tension

 transformed


SCC


SCC

high
stress

(b) deformation
behaviour

(c) residual
stress

(d) structure
change in 

Fig. 19 Stress response of the individual phases in 2304 uniaxially strained

parallel to the rolling direction [36].


800

Measured stress (MPa)

Ferrite
Austenite
Macro

600
400
200
0
-200
0.0

0.5

1.0
Applied strain (%)

1.5

acom 1 - 2011

20

Role of composition

Breaking time (h)

The role of the composition of the duplex stainless steels, or the individual phases, in
governing the resistance to stress corrosion cracking is, to put it mildly, a complex issue.
In an early work on duplex steels Ishizawa and Inazumi [37] concluded that stress corrosion
cracking of the series of laboratory melts investigated was governed more by the composition
of ferrite and austenite than by the ferrite-austenite ratio. Understanding of alloying effects is
difficult because the composition affects the mechanical properties, deformation behaviour
and electrochemical response of the phases, in addition to the influence on the phase balance.
One common misconception about alloying effects is that of the role of nickel. The Copson
curve [38] illustrated in Figure 20, has on numerous occasions created confusion that
duplex stainless steels, with their nickel contents typically in the range of 57% as shown
in Figure 2, would be expected to be particularly susceptible to stress corrosion cracking.
However, the important point to note here is that the curves were based only on austenitic
alloys, and the author himself clearly stated that the curve is extended towards 0% nickel
even though no compositions with less than 8% nickel were tested. The data is therefore
emphatically irrelevant to duplex stainless steels.
The roles of chromium and molybdenum are relatively uncontroversial, since both improve
the passivity of stainless steel and thus contribute to both resistance to depassivation and
to repassivation. The roles of nickel and nitrogen are complicated by the fact that both
these elements are strong austenite formers, so that in the absence of other changes they
will shift the phase ratio to the unfortunate situation seen in Figure 14. In single phase
austenite early work [39] indicated that nickel promoted cross-slip, by increasing the
stacking fault energy, and thus increased the resistance to stress corrosion cracking, while
nitrogen promoted planar slip which was attributed [40] to short range ordering effects
and increased the susceptibility to stress corrosion cracking in boiling MgCl2. However, as
has been discussed earlier, the performance of alloys in this environment is not representative
of service experience in dilute chlorides but instead strongly related to the tendency to
uniform corrosion, which is suppressed by nitrogen.
An attempt to clarify the role of nickel and nitrogen on the stress corrosion cracking has
been made within a European project [41] concluded in 2004. In addition to austenitic
alloys with different alloying levels, two series of duplex stainless steels based on 2205 and
2507 were prepared by Thyssen Krupp Nirosta and Sandvik Materials Technology respectively.
These included five materials so it was possible to see both the effect of nitrogen additions,
which also had the effect of increasing the austenite fraction, and that of compensating
for increased nitrogen by decreasing the nickel content to retain a roughly equal phase
ratio. The results very clearly demonstrated that there is no universal effect of nitrogen.
For example constant load testing in 6M CaCl2 indicated the high-nitrogen variant of
2205 but the low nitrogen variant of 2507 to be the more susceptible to stress corrosion
cracking than their counterparts at the other end of the nitrogen-alloying spectrum.
This would seem to tie in with work of Lillbacka et al [42] which has shown that the load
sharing between phases shifts between
Fig. 20 Copson curve showing how increased nickel improves
duplex stainless steel grades and affects the

the time to failure in boiling 42% MgCl2 [38].
transition point above which work hardening
causes the austenite to become the harder
1000
phase. In contrast, U-bend testing in 1% NaCl
in an autoclave at 225C indicated a higher
susceptibility of the high-nitrogen 2507
100
alloy than its low-nitrogen equivalent, i.e.
the opposite trend to that seen in constant
load in CaCl2. This once again makes the
point that results are method-dependent
10
and that the behaviour in concentrated and
dilute chloride environments can be very
different. There is ample room for extensive
1
future work to elucidate the differences in
0
20
40
60
80
100
operative
mechanisms.
Nickel (wt %)

acom 1 - 2011

21

Conclusions
This paper has addressed the performance of duplex stainless steels in the common tests
for chloride-induced stress corrosion cracking. An important point is that ranking is very
different in concentrated and dilute chlorides. The low pH and tendency to active corrosion
means that lower alloyed steels show less cracking than their higher-alloyed counterparts in
some tests, but can instead suffer from selective corrosion. In typical process environment with
low chloride concentrations at elevated temperatures higher alloying levels are more clearly
advantageous. The effect of test method has been elucidated, particularly observations that
slow strain rate is relatively more conservative for duplex grades than for their austentic
counterparts. The vitally important role of phase balance is discussed, together the factors
of deformation behaviour, mechanical and electrochemical interaction and residual stress
which all contribute to the superior stress corrosion cracking resistance of duplex stainless
steels.

References
[1] Outokumpu Corrosion Handbook, 10th Edition, Sandvikens Tryckeri. (2009)
[2] G. Vogt, Werkstoffe und Korrosion 29 721 (1978)
[3] P. Combrade, A. Desestret, P. Jolly, R. Mayoud: Proc. Predictive methods
for assessing corrosion damage, 1978, Japan, NACE (1982)
[4] P. Kangas, J. M. Nicholls Proc Stainless Steel 1993, Wiley (1995)
[5] P. Kangas, J. M. Nicholls, Materials and Corrosion 46, 354 365 (1995)
[6] Outokumpu Stainless, internal report
[7] S. Bernhardsson, R. Mellstrm, J. Oredsson. Proc. Duplex Stainless steels 1982,
ASM (1983)
[8] R.F.A. Pettersson et al: Stress corrosion cracking in newer stainless steel grades,
ECSC 7210-PR/046,. report EUR 20328 EN (2002)
[9] U. Kiviskk, G. Chai, Corrosion 59, 828 (2003)
[10] P. Gustafsson, H. Eriksson. Proc Duplex stainless steel 86, 381 (1986)
[11] C. Borghesan. M.Sc. thesis, KTH, Stockholm (2008)
[12] R. F. A. Pettersson et al. Avoiding catastrophic corrosion failure of stainless steels.
European commission report in press from RFCS-CT-2006-00022.
[13] J. M. Drugli, U. Steinsmo. Proc. NACE Corrosion 97 Paper 194 (1997)
[14 ] S. Shoji, N. Ohnaka: Boshoku Gijutsu, 38, 9297 (1989)
[15] E. Johansson, T. Prosek: NACE Corrosion 2007, paper 07475 (2007)
[16] T. Prosek, A. Iversen, C. Taxn. Corrosion 65 105 (2009)
[17] S. Henrikson, S, M. sberg. Corrosion 35 429,(1979)
[18] H. Andersen, P-E. Arnvig, W. Wasielewska, L. Wegrelius, C.Wolfe NACE
Corrosion Paper 251 (1998)
[19] R. S. Treseder: Guideline information on new wrought iron and nickel-base
corrosion resistant alloys Phase 1 Corrosion Test Methods, MTI, (1980)
[20] R. F.A Jargelius, R. Blom, S. Hertzman, J. Linder Proc Duplex Stainless Steels
91, (1991)
[21] J. W. Flowers, F. H. Beck. M. G. Fontana Corrosion 19, 186 (1963)
[22] R. N. Gunn Duplex stainless steels: Microstructure, properties and applications,
Abington (1997)
[23] R. A. Cottis, R. C. Newman Offshore Technology report OTH 94 440 (1995)
[24] NORSOK standard M-630 Material data sheets for piping (2004)
[25] P. Woolin, M. Maligas NACE Corrosion 2003, Paper 03132 (2003)
[26] DNV-RP-F112 Recommended practice for design of duplex stainless steel
subsea equipment exposed to cathodic protection, Det Norske Veritas (2008)
[27] K. Takizawa: Trans Iron & Steel Inst. Japan 24, (1984)
[28] R. F. A. Jargelius-Pettersson, J. Linder, S. Hertzman, H. strm SIMR report
IM-2974 (1993)
[29] J. Hochmann, A. Desestret, P. Jolly, R. Mayoud Proc. NACE-5: Stress corrosion
cracking and hydrogen embrittlement of iron-base alloys, 1973, 956, (1977)

acom 1 - 2011

22

[30] A. Desestret, F. Gauthey, J. L. Ranvier, J. C. Colson: Mm. Sci. Revue de


Mtallurgie 393 (1977)
[31] M. Kowaka, H. Nagano, T. Kudo, K. Yamanaka. Boshoku Gijutsu 30 218 (1981)
[32] H. Nagano, T. Kudo, Y Inaba, M Harada, M. Metaux-Corrosion-Industrie. 56,
8188 (1981)
[33] Shimodaira, M. Takano, Y. Takizawa, H. Kamide. Stress corrosion cracking
and hydrogen embrittlemnt of iron base alloys 1973, 1003 (1977)
[34] T. Magnin, J. Le Coze, A. Desestret: Proc Duplex stainless steels 1982, 535,
ASM (1983)
[35] T. Magnin, J. M. Lardon. Materials Scienc and Engineering A104, 21 (1988)
[36] J. J. Moverare, M. Oden, Met. and Materials Transactions 33A, 57 (2002)
[37] Y. Ishizawa, T. Inazumi Proc Duplex stainless steels 86, 392 (1986)
[38] H. R. Copson, Proc. Physical metallurgy of stress corrosion fracture, 247,
AIME (1959)
[39] P. R. Swann Corrosion 19, 102, (1963)
[40] D. L. Douglass, G. Thomas, W. R. Roser Corrosion 20 15, (1964)
[41] R. F. A. Pettersson et al: Critical evaluation of the effect of alloyed nitrogen
on the susceptibility of stainless steels to environmentally-induced cracking
Report EUR 21802 EN (2004)
[42] R. Lillbacka et al. Acta Materialia 55 5359 (2007)

Originally published at Duplex World 2010


and reproduced with permission.

1478EN-GB Art 58. March 2011

Comments on acom and its articles or suggestions on future articles are appreciated and should be
sent to the editor Andreas Persson at acom@outokumpu.com
This document is for information only and seeks to provide professionals with the best possible
information to enable them to make appropriate choices. Although every effort has been made to ensure
the accuracy of the information provided in this document, Outokumpu can not accept any responsibility
for any loss, damage or other consequence resulting from the use of this publication.
The information provided herein may be subject to alterations without notice.

Activating Your Ideas


Outokumpu is a global leader in stainless steel with the vision to be the undisputed number one.
Customers in a wide range of industries use our stainless steel and services worldwide. Being fully
recyclable, maintenance-free, as well as very strong and durable material, stainless steel is one of
the key building blocks for a sustainable future.
What makes Outokumpu special is total customer focus all the way, from R&D to delivery.
You have the idea.We offer world-class stainless steel, technical know-how and support.
We activate your ideas

Outokumpu Stainless AB, Avesta Research Centre


Box 74, SE-774 22 Avesta, Sweden
Tel. +46 (0) 226 - 810 00, Fax +46 (0) 226 - 810 77

www.outokumpu.com

You might also like