Download as pdf or txt
Download as pdf or txt
You are on page 1of 5

Materials Chemistry and Physics 126 (2011) 602–606

Contents lists available at ScienceDirect

Materials Chemistry and Physics


journal homepage: www.elsevier.com/locate/matchemphys

Anomalous corrosion resistance behavior of Mo-containing SS in alkaline media:


The role of microstructure
Thiago J. Mesquita a,b , Eric Chauveau b , Marc Mantel b , Nicole Kinsman c , Ricardo P. Nogueira a,∗
a
UMR 5631 INP Grenoble-CNRS-UJF, LEPMI, BP 75, 38402 St. Martin d’Hères, France
b
CRU Ugitech, Av Paul Girod 73400 Ugine, France
c
International Molybdenum Asociation, IMOA W4 4JE London, United Kingdom

a r t i c l e i n f o a b s t r a c t

Article history: Electrochemical and microscopy techniques were used to investigate the role of the microstructure in
Received 21 May 2010 the corrosion behavior of Mo-containing stainless steel grades in alkaline chloride rich media. It is shown
Received in revised form 25 October 2010 that under certain conditions, Mo containing grades can be even less corrosion resistant than Mo free
Accepted 8 January 2011
ones. This anomalous behavior is straightforwardly related to the austenite phase. On the other hand,
results show that Mo confirms its clearly beneficial effect in ferritic and duplex steels. Current transient
Keywords:
analysis indicated that the protection mechanism in those cases appears to be related to an inhibited pit
Alloy
initiation rather than faster repassivation kinetics.
Corrosion
SEM © 2011 Elsevier B.V. All rights reserved.
Metals

1. Introduction Mo alloys for alkaline service conditions. It is worth noticing how-


ever that those results were limited to austenitic steels in some
In the last few years, strong fluctuations and a net increasing ten- specific conditions [12] or were based upon preliminary laboratory
dency of the price of several important alloying elements yielded observations [11].
much effort from the major stainless steel suppliers to propose new The aim of the present paper is to characterize the behav-
technological solutions that minimize the use or even eliminate ior of different Mo-containing and Mo-free austenitic, ferritic and
those alloying elements (mainly nickel and molybdenum) in the duplex stainless steels exposed to chloride rich alkaline media.
composition of special grades. This at least partially explains the These conditions are particularly relevant to coastal or off shore
appearance of new less expensive alloys like lean duplex stainless rebar applications or alkali industries.
steels on the market [1,2]. In general the use of Mo-free or low
2. Experimental
Mo content alloys is now sometimes extended into service con-
ditions for which formerly Mo-containing alloys would have been In a first set of experiments, several industrial grades of stainless steel were
used. This development is mostly due to economic reasons inas- investigated in the present study: austenitic 304 and 316, ferritic 430 and 434 and
duplex 2304 and 2205, the nominal compositions of which are given in Table 1.
much as the role of Mo is still not fully understood. This lack of
It can be seen that for each microstructure, the only major difference was the Mo
scientific knowledge concerning the role of Mo can be illustrated content (0.21 and 2.17 wt.% for the austenitic grades, and 0.05 and 0.92 wt.% for
by a fundamental disagreement in the literature as to whether Mo the ferritic 0.28 and 2.78 wt.%for the duplex). The others elements for each grade
is a component of passive films [3,4] or not [5–7]. at most slightly varied so that the corrosion resistance difference for steels with
The consequences of a more widespread use of these alloys are the same microstructure could only be reasonably attributed to this presence or
“absence” of Mo. Nevertheless, to reinforce this interpretation based on Mo content
difficult to predict. Indeed, it is true that the positive role of Mo in
differences, specially prepared laboratory duplex grades for which strictly only Mo
acidic media can be considered as generally accepted [8–10]. In the concentration varied (0.02 and 3.00 wt.%, see Table 2) were also investigated.
case of alkaline media, however, the role of Mo is much less studied. Samples (2.5 cm2 ) were polished till 1200 grid emery paper, degreased with
Some previous studies have pointed out that in certain conditions alcohol, washed with distilled water and dried with hot air. Electrochemical mea-
surements were performed in a classical three-electrode electrochemical cell with
Mo has no beneficial effect on the pitting corrosion resistance of
a large surface Pt grid as the counter-electrode and a saturated calomel (SCE) as the
stainless steels [11,12], which could thus justify abandoning high reference one. The anodic polarization curves were performed after 1 h of immersion
in the electrolyte at 1 mV s−1 scan rate from 30 mV below open circuit potential up
to 100 ␮A cm−2 current density. Metastable pit transients were acquired at 500 Hz
sample frequency after proper anti-aliasing filtering [13].
∗ Corresponding author at: LEPMI, 1130, rue de la Piscine, Domaine Universitaire, The base electrolyte was a 0.025 M NaHCO3 + 0.025 M Na2 CO3 ·10H2 O + 0.6 M
BP 75, 38402 St. Martin dı̌Hères, France. Tel.: +33 04 76 82 65 93. NaCl solution at 25, 50 and 75 ◦ C and at several pHs in the 0.6–12 range adjusted
E-mail addresses: ricardo.nogueira@lepmi.grenoble-inp.fr, with HCl or NaOH droplets. In what follows results will be presented at 25 ◦ C, only.
ricardo.nogueira@enseeg.inpg.fr (R.P. Nogueira). The other temperatures gave qualitatively equivalent results.

0254-0584/$ – see front matter © 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.matchemphys.2011.01.013
T.J. Mesquita et al. / Materials Chemistry and Physics 126 (2011) 602–606 603

Table 1
Nominal composition of the industrial grades investigated.

Industrial material Elements (wt.%)

C Si Mn Ni Cr Mo Cu N Co Sa P

Austenitic 304 0.02 0.49 0.60 11.12 18.29 0.21 0.31 0.03 0.00 9 0.00
316 0.01 0.49 0.73 11.08 16.89 2.17 0.48 0.03 0.25 10 0.03
Ferritic 430 0.01 0.31 0.30 0.29 16.16 0.05 0.10 0.03 0.02 5 0.02
434 0.03 0.39 0.39 0.45 16.17 0.92 0.12 0.05 0.03 23 0.02
Duplex 2304 0.02 0.41 1.09 4.02 22.30 0.28 0.30 0.15 0.13 4 0.02
2205 0.02 0.4 1.61 5.45 22.91 2.78 0.22 0.15 0.07 3 0.02
a
S ppm.

Table 2
Nominal composition of the laboratory prepared duplex grades investigated.

Laboratory material Elements (wt.%)

C Si Mn Ni Cr Mo Cu N V Sa P

Duplex 2304 0.03 0.58 1.03 4.57 22.97 0.02 0.01 0.18 0.01 12 0.01
2205 0.03 0.56 0.98 4.56 22.97 3.00 0.01 0.19 0.01 13 0.01
a
S ppm.

3. Results and discussion tial (Epit ) of about 200 mVsce (Fig. 1a). As the pH was made less
acidic both the 304 and the 316 showed the expected increase
Fig. 1a and b shows typical polarization curves obtained for 304 of Epit values. In the alkaline domain, however, the Mo-free 304
and 316 SS at extreme pH conditions (0.6 and 12, respectively). As becomes surprisingly more resistant than the 316 as seen in Fig. 1b
largely accepted, the figure illustrates the huge positive effect of Mo (Epit = 460 mVsce and 620 mVsce for 316 and 304, respectively). It
on the corrosion resistance of austenitic grades in acidic environ- seems that, for these austenitic grades, there is a monotonic loss
ments. Indeed, in those harsh conditions (pH 0.6), 304 steel shows of Mo effect as the pH increases as already pointed out in some
no passive behavior and corrodes already at open circuit potential previous studies [11,12]. Indeed, the Epit difference between the
whilst 316 shows a clear passivation plateau and a pitting poten- 316 and 304, monotonically decreased with the pH increase from

Fig. 1. Polarization curves obtained for the different grades. (A) 304 () and 316 () at pH 0.6; (B) 304 () and 316 () at pH 12; (C) 430 () and 434 () at pH 10; (D) 2304
() and 2205 () at pH 10. The current increase up from ca. 840 mVsce for the 2205 grade is due to oxygen evolution and not pitting.
604 T.J. Mesquita et al. / Materials Chemistry and Physics 126 (2011) 602–606

Fig. 2. Polarization curves obtained for Mo-free 2304 () and Mo-containing 2205 () laboratory prepared samples at pH 4.0, 7.0 and 10.0 at 25 ◦ C. 2205 samples only showed
oxygen evolution and no pitting.

+450 (pH 0.6) to −160 mV (pH 12), negative values indicating that These results are indeed self-consistent in the sense that if ferrite
the Mo-free alloy became more resistant than the Mo-containing had also shown the anomalous behavior in alkaline media shown
316 as seen in Fig. 1b. This anomalous behavior could be tenta- by austenite, it would not be possible to understand why duplex
tively related to the reaction mechanism associated to the effect of could still behave better in the presence of Mo if its two constituting
Mo, for instance a pH dependent state of charge that could allow microstructures behaved in the opposite way. It is worth noticing
Cl− trapping by Mo cations in acidic solutions [14] but not in alka- that in this sense, the present study brings a new insight that can
line media. However, these mechanistic insights are still far from help understanding the beneficial role of Mo in the corrosion resis-
being fully researched and understood, as mentioned before and tance of different grades. Indeed, unlikely to Refs. [11,12] (limited
are beyond the aim of the present paper. to austenitic [12] or austenitic and duplex [11] grades at different
In a first glance, these first results seem to validate the idea pH values), the present paper dealt with different microstructures
that Mo is not functional in alkaline media, at least not with and showed that Mo effects remain positive at alkaline media for
respect to localized corrosion resistance. Moreover, since austenite ferritic and duplex grades.
is intrinsically more corrosion resistant than ferrite in most ser- As mentioned before, the previous results have been obtained
vice conditions, it could be assumed, that the effect discussed from from industrial grades, that, although having equivalent composi-
Fig. 1a and b equally applies to ferritic and duplex SS. This hasty con- tions as shown in Table 1, could convey differences that could be
clusion is however incorrect. Indeed, at very acidic pH, the ferritic also at the origin of these distinct corrosion resistance behaviors
steels corroded intensively, as did the 304. The main difference is (there were indeed some small differences as for instance the Ni
that, for neutral and alkaline media, the Epit difference between the content, even if not more than 1.5%). In this sense, we have also
Mo-containing and the Mo-free ferritic grades does not decrease proceeded to complementary tests on laboratory prepared duplex
with increasing pH but oscillated between +20 and +90 mV in the alloys for which the effective compositions were fully controlled so
pH range studied, whilst as mentioned before, the austenitic grade that absolutely only the Mo content varied, the others remaining
difference monotonically decreased. Fig. 1c and d illustrates this strictly constant as depicted in Table 2. Fig. 2 illustrates the typical
issue with the help of polarization curves obtained at pH 10 for the corrosion resistance behavior of the Mo containing and free grades
ferritic and duplex alloys (at pH 12 duplex grades have shown no of the laboratory prepared SS, confirming the positive Mo role on
pit at all). It clearly appears that, unlike in austenitic grades, Mo the pitting corrosion resistance of these alloys. Indeed, for the three
has a positive effect on their corrosion resistance in alkaline envi- pH values studied, the Mo containing grades showed no pitting
ronments (all graphs in Fig. 1 are just typical examples of quite whilst the Mo-free samples always presented localized attacks up
reproducible behaviors: on average the presence of Mo increased from potential values depending on the pH solution.
the pitting potential about 50 mV for the ferritic and more than Fig. 3 shows SEM pictures of pits in industrial duplex SS under
150 mV for the duplex grades). This seems to indicate that Mo does different conditions. The left picture shows that in acidic media,
have a protective role in alkaline media, but that it is mostly related in the absence of Mo, pits are initiated and propagated mostly at
to the improvement of the corrosion resistance of the ferrite phase. the ferrite phase from which grains appeared to be completely dis-

Fig. 3. Back scattered SEM images from corrosive attacks on different samples. Left: 2304 at pH 0.6 showing austenite (lighter) grains left after ferrite (darker) grains have
been dissolved. Middle: 2304 at pH 10; pit initiation at the ferrite (darker) phase. Right: 2205 at pH 10; pit initiation at the austenite (lighter) phase.
T.J. Mesquita et al. / Materials Chemistry and Physics 126 (2011) 602–606 605

Fig. 4. Back scattered electrons SEM images from 2205 at pH 10. Left: ferrite (darker) phase pins down the pit growth that attacks preferentially the austenite (lighter) phase.
Right: zoom of pit propagation inside the austenite phase.

solved whilst austenite grains remained much less attacked. In the These results indicate that the Mo role is more likely to be
presence of Mo, however, pit propagation is strongly hindered in ascribed to some prior-to-the-attack metallurgical effects rather
the ferrite phase and the attack appeared to split between the two than to post-attack interfacial kinetic processes, like those related
phases. Results are however more pronounced in alkaline media. In to Mo cations interactions with electrolyte anions. Indeed, if the
the absence of Mo, pits are, as expected, initiated in the less resistant Mo effect was related to chemical or electrochemical interactions
ferrite phase (Fig. 3, middle). On the other hand, in the presence of of Mo cations with the electrolyte, pH changes should have more
Mo, the scenario is reversed. The ferrite becomes much more resis- or less the same impact regardless of the allotropic form, which
tant and pits were now almost exclusively located at the austenite is clearly not the case as presented before. Those later interfacial
grains, as seen in Fig. 3, right. After continuing corrosion attack, the effects have been proposed to be at the origin of enhanced repas-
ferrite phase definitely shows better corrosion resistance as shown sivation kinetics [14], which can be expected to yield smaller time
in Fig. 4. Indeed, the periphery of the big pit reveals an attack pro- constants of the repassivation transient after a metastable pit ini-
file characterized by the preferential hammering of the austenite tiation, for instance. Fig. 5 shows two typical current transients
phase whilst the ferrite grains seem to locally pin down the pit – shifted to have the same arbitrary time origin – associated to
propagation (Fig. 4, left). In the right side a more zoomed picture metastable pits taking place at 430 and 434 surfaces at pH 12 and
clearly shows that the pit is not only initiated at the austenite but polarized to +500 mV with respect to their open circuit potential.
also propagates inside it before eventually eating into the adjacent The repassivation (represented by the current decay after reach-
ferrite. The same behavior has been observed in the case of the ing a maximum) appears quite steep in both cases, thus illustrating
laboratory prepared grades. equivalent repassivation kinetic in the absence and in the presence
This indirect evidence of the role of Mo is confirmed by a met- of Mo. On the other hand, the pit initiation (related to the current
allurgical analysis. A calculation of each phase composition using increase) seems to be slowed down in the presence of Mo, showing
Thermocalc® software confirmed that for the 2205 steel, Mo is pref- effectively a much less pronounced current increase profile (higher
erentially dissolved into the ferrite phase (3.2%) that hence has its time constant) represented by a rise time up to the maximum of
corrosion resistance improved compared to the austenite phase about 0.7 s for the 430 and 2.3 s for the 434 as well as a lower tran-
which contains less Mo (2.0%). More importantly, this better fer- sient amplitude. This would effectively mean that the role of Mo
rite behavior cannot be ascribed to different Cr contents in 2304 should not be related to the way it accelerates repassivation but
and 2205 steel since their Cr contents were roughly the same (pre- mostly to how it inhibits pit initiation. Another point deserving
cisely 23.7 and 20.1 and 24.0 and 20.5 for ferrite and austenite and specific mention is related to the wide use of the so-called pitting
2304 and 2205, respectively). Results obtained from Thermocalc® resistance equivalent number, PREN, calculation to forecast local-
have been complemented by both energy dispersive spectrome- ized corrosion resistance. This strictly empirical formula does not
ter (EDS) and wavelength-dispersive spectrometer (WDS) as seen factor in microstructural or electrolyte variables, which, as seen in
in Table 3.Values are in full agreement with those obtained from
theoretical estimations and confirm the preferential Mo dissolu-
tion in ferrite. It is also worth noticing that the enhanced corrosion
resistance of the ferrite phase can of course drastically change in
the presence of sigma phase precipitates. The austenite phase is in
these conditions left at the rear of the corrosion front whilst the
ferrite phase is preferentially attacked [15].

Table 3
Thermocalc® calculations and EDS and WDS measurements of the Mo and Cr com-
position in austenite and ferrite for the industrial duplex grades.

Method Sample Phase Elements (wt.%)

Cr Mo

Thermocalc 2304 ␣ 25.19 0.36


␥ 20.41 0.23
2205 ␣ 25.32 3.40
␥ 20.34 2.16
EDS 2304 ␣ 24.35 0.27
␥ 21.09 0.00
2205 ␣ 24.57 3.32
␥ 20.25 1.95
WDS 2304 ␣ 24.55 0.36
Fig. 5. Time records of metastable pit transients taking place at 430 (, gray) and
␥ 20.75 0.24
434 (, black) grades at pH 10. Both samples polarized at +500 mV with respect to
2205 ␣ 25.06 3.46
their open circuit potential. Transients were shifted so as to coincide their initiation
␥ 21.29 2.16
at an arbitrary time origin. Sample frequency 500 Hz.
606 T.J. Mesquita et al. / Materials Chemistry and Physics 126 (2011) 602–606

the present paper, are of major concern in the corrosion resistance [3] M.F. Montemor, A. Simões, M.G.S. Ferreira, M. Da Cunha Belo, Corros. Sci. 41
of steels exposed to different media. (1999) 17.
[4] K. Sugimoto, Y. Sawada, Corros. Sci. 17 (1977) 425.
[5] D. Addari, B. Elsener, A. Rossi, Electrochim. Acta 53 (2008) 8078.
4. Conclusions [6] A.E. Yaniv, J.B. Lumsden, R.W. Staehle, J. Electrochem. Soc. 124 (1977) 490.
[7] M.W. Tan, E. Akiyama, H. Habazaki, A. Kawashima, K. Asami, K. Hashimoto,
Corros. Sci. 39 (1997) 589.
In summary it is shown that Mo has a strong beneficial effect [8] A. Pardo, M.C. Merino, A.E. Coy, F. Viejo, R. Arrabal, E. Matykina, Corros. Sci. 50
on the corrosion resistance of ferritic and duplex stainless steels (2008) 1796.
in chloride-rich alkaline environments. This is more likely due to [9] A. Pardo, M.C. Merino, A.E. Coy, F. Viejo, R. Arrabal, E. Matykina, Corros. Sci. 50
(2008) 780.
a hindered pit initiation than due to kinetic aspects related to the [10] C. Lemaitre, A. Abdel Moneim, R. Djoudjou, B. Baroux, G. Beranger, Corros. Sci.
modification of electrolyte composition as proposed in the case of 34 (1993) 1913.
acidic solutions. [11] E. Chauveau, T. Sourisseau, B. Dermelin, M. Mantel, MEDACHS 08 – Int. Conf.
Coast. Mar. Environ., 2008.
[12] S.A.M. Refaey, F. Taha, A.M.A. El-Malak, Appl. Surf. Sci. 242 (2005) 114.
References [13] I.N. Bastos, F. Huet, R.P. Nogueira, P. Rousseau, J. Electrochem. Soc. 147 (2000)
671.
[1] G. Gedge, J. Construc, Steel Res. 64 (2008) 1194. [14] V. Vignal, J.M. Olive, D. Desjardins, Corros. Sci. 41 (1999) 869.
[2] L. Zhang, W. Zhang, Y. Jiang, B. Deng, D. Sun, J. Li, Electrochim. Acta 54 (2009) [15] I.N. Bastos, S.S.M. Tavares, F. Dalard, R.P. Nogueira, Scripta Mater. 57 (2007)
5387. 913.

You might also like