You are on page 1of 8

Available online at www.sciencedirect.

com

Procedia Materials Science 1 (2012) 207 – 214

11th International Congress on Metallurgy & Materials SAM/CONAMET 2011.

Effect of temperature and chloride concentration on the anodic


behavior of nickel alloys in bicarbonate solutions
N. Zadoroznea,b, R. Rebakc, M. Giordanob, A. Aresa and R. Carranzab

a
CONICET -Prog. De Materiales, Modelización y Metrología FCEQyN-UNaM, élix de Azara 1552 - C.P. N3300LQH - Posadas -
Misiones – Argentina
b
Depto. Materiales - Comisión Nacional de Energía Atómica Instituto Sabato - Univ. Nac. de San Martín / CNEA Av. Gral. Paz 1499
San Martín, B1650KNA Buenos Aires, Argentina
c
GE Global Research Schenectady, NY 12309, USA

Abstract

It has been shown that the presence of bicarbonate ions and chloride is necessary to produce stress corrosion cracking
(SCC) in alloy 22 at anodic applied potentials. It has been suggested that the susceptibility to SCC could be related to the
occurrence of an anodic peak in the polarization curves below the transpassive range.
The aim of this work is to study the anodic behavior of alloy 22 and other nickel alloys in different media containing
bicarbonate ions and chloride in various concentrations and temperatures to determine if there is a correlation between the
presence of the anodic peak and the occurrence of SCC; and to determine which element in the alloy maybe be
responsible for the cracking. The studied alloys included alloy 22, Ni-200, alloy 800 and alloy 600. Results show that in
general the potential and the peak current increased with temperature and chloride concentration in the solution. For the
alloys with the smaller alloying content the peak potential decreased to the point of disappearing. There was no
dependence of the current peak with the amount of alloying.

th
©2012
© 2012Published
Publishedbyby Elsevier
Elsevier Ltd.
Ltd. Selection
Selection and/or
and/or peer-review
peer-review underunder responsibility
responsibility of 11
of SAM/
International Congress on Metallurgy
CONAMET 2011, Rosario, Argentina. & Materials SAM/CONAMET 2011.

Keywords: nickel alloys, N06022, corrosion, bicarbonate, anodic behavior

* Corresponding author. Tel.: ; fax: +54-0376-4425414 .


E-mail address: nataliazadorozne@gmail.com

2211-8128 © 2012 Published by Elsevier Ltd. Selection and/or peer-review under responsibility of SAM/CONAMET 2011, Rosario, Argentina.
doi:10.1016/j.mspro.2012.06.028
208 N. Zadorozne et al. / Procedia Materials Science 1 (2012) 207 – 214

1. Introduction

The ability of nickel (Ni) to retain large amounts of different alloying elements has led to the development
of several Ni alloys families. The most versatile family is the Nickel-Chromium-Molybdenum (Ni-Cr-Mo) or
C-type, which provides corrosion resistance in reducing and oxidizing hot acids. Alloy 22 (UNS N06022) is
one of the most versatile members of this family. Due to its excellent corrosion resistance in a wide variety of
environments, alloy 22 has been selected for the fabrication of the corrosion-resistant outer shell of the high-
level nuclear waste container (Rebak 2003, Gordon 2002).
Since over their life-time the containers may be exposed to multi-ionic aqueous environment, it is estimated
that this material could suffer three different types of deterioration: general corrosion, localized corrosion
(specifically crevice) and stress corrosion cracking (SCC). Generalized and localized corrosion has been
extensively studied in environments in presence of chemicals that simulate groundwater. It is generally found
that alloy 22 is highly resistant to localized and general corrosion in the ground water environments
(Cragnolino 2002, Rebak 2009, Carranza 2008, Rebak 2011). On the contrary, relatively few experimental
studies were performed to understand the processes of stress corrosion cracking of alloy 22.
Researchers at Lawrence Livermore National Laboratory studied the range of potential susceptibility to
SCC of alloy 22 in slow strain rate testing using simulated groundwater. Variables studied included effect of
potential, type of chemicals and temperature. Their results suggest that alloy 22 is susceptible to SCC in a
potential range between +0.3 and +0.4 VSSC (VSSC = VENH + 0.198V) in concentrated groundwater above 65
ºC. No SCC was found in basic or acidified saturated groundwater. Chiang et al. (2007) studied the effect of
some components of saturated groundwater on SCC for alloy 22 at 95 ºC. Their results suggest that is essential
the coexistence of bicarbonate and chloride ions to produce SCC. The susceptibility seems to increase with
increasing chloride concentration. Laboratory studies found that the probability of cracking was low at the
corrosion potential in all the media investigated.
The need for the presence of bicarbonate ions to produce cracking was confirmed by researchers at
Southwest Research Institute. It was found experimentally that SCC occurs in the pH range between 8.5 and
10.5. They determined that the susceptibility to SCC could be related to the appearance of an anodic peak in
the polarization curves in simulated concentrated water at high potential. The current density at the peak was
higher in the bicarbonate solution containing the higher chloride concentration (Rebak 2005, King 2004,
Chiang 2006) .
Due to SCC of the alloy22 could be related to the appearance of the anodic peak, the objective of this work
is focused on studying the anodic behaviour of alloy 22 and other nickel alloys (600, 800H and 201 alloys) in
different media containing bicarbonate and chloride ions in various concentrations and temperatures, to
determine the characteristics and dependencies of the anodic peak current with these parameters and to
compare the results with the SCC tests to be conducted as a follow up of this work in the near future.

2. Experimental Procedure

The chemical compositions of the tested alloys in weight percent are listed in Table 1. The specimens used
were in the form of parallelepipeds, a variation of the ASTM G 5 (AST; G5-94, 2004) specimen, with
approximate dimensions of 12 mm x 12 mm x 15 mm. Each specimen was screwed to a metal rod that was
used as electrical contact, which was introduced in a glass test tube holder and was isolated from the solution
with a PTFE gasket. The exposed area to the solution was approximately 10 cm2. The specimens had a finish
grinding of abrasive SiC paper number 600 and were degreased in acetone and washed in distilled water
within the hour prior to testing. Figure 1 shows a schematic of the specimen. All the electrochemical tests
were conducted in a one-liter, three-electrode vessel. The nitrogen (N2) was purged through the solution 1
N. Zadorozne et al. / Procedia Materials Science 1 (2012) 207 – 214 209

hour prior to testing and was continued throughout the entire test. A water-cooled condenser combined with a
water trap was used to avoid evaporation of the solution and to prevent the ingress of air (oxygen). The
temperature of the solution was controlled by immersing the cell in a water bath, which was kept at a constant
temperature. The reference electrode was a saturated calomel electrode (SCE) (VSCE = VSHE + 0.242V). The
reference electrode was connected to the solution through a water-cooled Luggin probe. The counter electrode
consisted in a flag of platinum foil (total area 50 cm2) spot-welded to a platinum wire. All the potentials in
this paper are reported in the SCE scale.

Table 1. Chemical composition of nickel based alloys (Weight %)

Alloy Ni Cr Mo W Fe Si Mn C S Other
22 56.58 22.26 13.90 3.15 3.81 0.02 0.24 0.004 0.001 Co: 0.2
-
600 73.44 16.56 - 8.83 0.27 0.24 0.06 <0.001 Al: 0.253

Al:0.49
800H 31.49 20.92 - - 44.89 0.28 1.03 0.09
Ti:0.57
201 99.50 0.11 0.10 0.23 0.02 0.001 Co: 0.03
Data from alloy certifications

Fig. 1. Schematic of the prismatic specimens.

The solution used were: 1.148 mol/L NaHCO3, 1.148 mol/L NaHCO3 + 0.1 mol/L NaCl and 1.148 mol/L
NaHCO3 + 1 mol/L NaCl. The choice of these environments was based in multi-ionic solution called SCW
(Simulated Concentrated Water), used in the works on SCC in alloy 22. The pH of the solutions was kept in a
range between 8 and 8.5 at the start of the test. The test temperatures were 25ºC, 60ºC, 75ºC and 90ºC. The
potentiodynamic polarization curves were performed using a scan rate of 0.167 mV / s.

3. Results

Figures 2-5 show the potentiodynamic polarization curves of alloys 22, 600, 800H and 201 respectively in
1.148 mol/L HCO3Na at 25ºC, 60ºC, 75ºC and 90ºC. The potentiodynamic polarization curves performed in
210 N. Zadorozne et al. / Procedia Materials Science 1 (2012) 207 – 214

1.148 mol/L HCO3Na + 0.1 mol/L NaCl and 1.148 mol/L HCO3Na + 1 mol/L NaCl showed similar behavior
to the curves made in 1.148 mol/L NaHCO3. The curves in presence of NaCl are not shown. Alloy 22
presented a wide passive zone, followed by an increase in current which declined after forming a current peak
prior to the transpassive zone.

10m
10m
90°C
90°C
1m 75°C
1m 75°C
60°C
60°C
100µ 25°C
25°C
100µ

i (A/cm )
i (A/cm )

2
10µ
2

10µ


Alloy 22 1µ
100n 1,148 mol/L HCO3Na
Alloy 600
100n
1,148 mol/L HCO3Na
10n
10n
1n -0,8 -0,6 -0,4 -0,2 0,0 0,2 0,4 0,6 0,8 1,0
-0,8 -0,6 -0,4 -0,2 0,0 0,2 0,4 0,6 0,8 1,0
E(VSCE)
E(VSCE)

Fig. 2. Potentiodynamic polarization curves of alloy 22, in 1.148 Fig. 3. Potentiodynamic polarization curves of alloy 600, in
mol/L NaHCO3 at 25ºC, 60ºC, 75ºC and 90 ºC 1.148 mol/L NaHCO3 at 25ºC, 60ºC, 75ºC and 90 ºC

10m
10m
90 °C
90°C 1m 75 °C
1m 75°C 60 °C
60°C 25 °C
100µ 25°C 100µ
i (A/cm )
2
i (A/cm )

10µ
2

10µ


100n Alloy 201


Alloy 800 H 1,148 mol/L HCO3Na
100n
1,148 mol/L HCO3Na
10n

10n
1n -1,0 -0,8 -0,6 -0,4 -0,2 0,0 0,2 0,4 0,6 0,8 1,0 1,2
-0,8 -0,6 -0,4 -0,2 0,0 0,2 0,4 0,6 0,8 1,0
E(VSCE)
E(VSCE)

Fig. 4. Potentiodynamic polarization curves of alloy 800H, in


Fig. 5. Potentiodynamic polarization curves of alloy 201, in
1.148 mol/L NaHCO3 at 25ºC, 60ºC, 75ºC and 90 ºC
1.148 mol/L NaHCO3 at 25ºC, 60ºC, 75ºC and 90 ºC

Alloy 600, like alloy 22, presented a passive zone, an anodic peak and finally a transpassive zone. For
alloy 600, it was observed that after the anodic peak, the current reached lower values than for alloy 22. This
may be an effect of Mo in the alloy 22.
For Alloy 800H the behavior was different. Alloy 800H showed a small passive zone compared with alloys
22 and 600 and then showed two increases in the current density previous to the transpassive zone. The first
increase came in the form of a well defined peak at a potential close to 0 mVSCE when the temperature was
90°C. Recalling that 800H is the material that has the highest amount of iron (44.89%), this peak can be
attributed to the transformation of Fe+2 to Fe+3 in the oxide layer. After the first peak, alloy 800H showed a
second increase of the current much more gradual, the alloy 800H reached a maximum value of current and
then decreased slowly. Approximately at 670 mVSCE and at 90ºC, the current increased abruptly leading to a
transpassive zone. During data analysis, the anodic peak potential selected corresponded to the highest current
value in the second peak.
N. Zadorozne et al. / Procedia Materials Science 1 (2012) 207 – 214 211

Alloy 201 showed an active to passive transition peak followed by a passive range. This peak of activity
after the corrosion potential is not observed in alloys 22, 600 and 800H due to the presence of Cr in the latter.
Alloy 201 showed an anodic peak prior to transpassivity only when the test temperature was 90 ºC.
All the alloys in all the environments studied showed a modification of the polarization curves when the
test temperature changed. As the temperature decreased the potential of the anodic peak that occurred prior to
the transpassive range was higher. The decrease in temperature also influenced the appearance of anodic peak,
making it less obvious as the temperature decreased. Alloys 600 and 800H did not have a well defined peak
when the temperature was 25 ºC.
At the end of the tests, the pH of the solution was measured and an increase of pH was observed compared
to the initial value. The highest increase in pH occurred when the test temperature was 90ºC.At this
temperature the pH of the solution after the test was10.5.When the test temperature was 25ºC, the change in
pH was lower reaching 9.
After testing all the specimens were cleaned by ultrasound and observed by scanning microscope. Figures
6-20 show representative micrographs of the samples of alloys 22, 600, 800H and 201, respectively, after the
polarization curves in the studied environments. Alloy 22, which presents a well-defined anodic peak, did not
show localized attack, while the alloy 201, in which no anodic peaks were observed, showed localized
corrosion.

Fig. 6: SEM image of alloy 22 after polarization curve in 1.148 Fig. 7: SEM image of alloy 600 after polarization curve in 1.148
mol/L NaHCO3 at 75ºC mol/L NaHCO3 at 75ºC

Fig. 8: SEM image of alloy 800H after polarization curve in 1.148 Fig. 9: SEM image of alloy 201 after polarization curve in 1.148
mol/L NaHCO3 at 75ºC mol/L NaHCO3 at 75ºC
212 N. Zadorozne et al. / Procedia Materials Science 1 (2012) 207 – 214

This same behavior was found in the specimens tested at all the temperatures and in all the environments.
It seems that the appearance of the anodic peak is not directly related to the presence of localized attack in the
material.
From the polarization curves the values of potential and current density at which the anodic peak occurs
before the transpassivity were obtained .It was found that the potential (EPEAK) and the current density (IPEAK)
at which the peak occurred also changed both with the test temperature, and with the addition of different
concentrations of chloride ions to the base solution of bicarbonate.

4. Discussion

4.1. Effect of temperature on the appearance of the anodic peak.

EPEAK in alloys 22 and 600 changed with the test temperature. EPEAK showed an increase as the test
temperature decreased. IPEAK in alloys 22 and 600 were not affected by the temperature variation of the tests.
This behavior can be observed in Figure 10, which shows the variation of the EPEAK and IPEAK with the
temperature for alloy 22 in 1.148mol/L HCO3Na. This behavior was observed for all the environments tested.
For alloy 800H, EPEAK and IPEAK changed with the temperature. EPEAK increased with decreasing temperature,
while IPEAK decreased with the temperature. The behaviors described above were repeated in the three
environments studied in this work.

4.2. Effect of concentration of the solutions in the appearance of the anodic peak.

As chloride ions were added to the bicarbonate solution, IPEAK and EPEAK showed increased values for
alloy22 at all the temperatures tested. Figure 11 shows the variation of the EPEAK and IPEAK with respect to the
concentration of the solution for alloy 22 at 90 ºC. Alloy 600 showed an increase in the EPEAK with the
increase of chloride ions in the solution, while IPEAK appeared approximately constant. Alloy 800H showed no
relationship of EPEAK and IPEAK for the different concentrations of the solutions tested.

4.3. Effect of chemical composition of the alloys in the appearance of the anodic peak.

The variations presented by the anodic peak, for both potential and current, for different percentages of
alloying elements were analyzed. Due to the absence of a peak in the curves of alloy 201, it is estimated that
the anodic peak may be associated primarily with one or more of the alloying elements in alloy 22. Because
the alloys that do not have Mo (600 and 800) also present the anodic peak, it is assumed that Mo is not
responsible for the appearance of the anodic peak. A correlation was anticipated between the variation of
EPEAK and the composition of the alloys. For the current results, no relationship was found with any of the
alloys. It was observed that the increase of nickel content of the alloys (800H, 22 and 600) does not modify
EPEAK significantly. Figure 12 shows the variation of the potential of the anodic peak as a function of nickel
content of the alloys studied in this work. From this relationship it is seen that high alloyed nickel present the
anodic peak at potentials near 0.4 mVSCE, and EPEAK did not show a clear variation with nickel content. As for
the current density peak, no relationship was found between the characteristics.
More recently, additional measurements with newly acquired alloys: Ni-20% Cr and Ni-28.5% Mo were
made. These new alloys are expected to give more information on how the alloying elements may influence
the anodic behavior of alloy 22. To conclude the research program, slow strain rate tests will be conducted to
determine if there is a correlation between the anodic peak and the susceptibility to SCC for the studied nickel
based alloys.
N. Zadorozne et al. / Procedia Materials Science 1 (2012) 207 – 214 213

-3
0,6 3,0x10

EPEAK Alloy 22 -3
2,5x10
1,148 mol/L NaHCO3
0,5
-3
2,0x10

EPEAK (VSCE)

IPEAK (A/cm )
0,4 -3
1,5x10

2
-3
1,0x10
0,3 IPEAK

-4
5,0x10

0,2
20 30 40 50 60 70 80 90
T (ºC)

Fig. 10. Variation of potential (EPEAK) and current density (IPEAK) of the anodic peak as a function of the temperature

0,29

-3
EPEAK 1,50x10
0,28
Alloy 22
90ºC
IPEAK -3
1,25x10
EPEAK(VSCE)

IPEAK(A/cm )
0,27

2
0,26 -3
1,00x10

0,25
-4
7,50x10

1,148 mol/L NaHCO3 1,148 mol/L NaHCO3


1,148 mol/L NaHCO3
+ 0,1 mol/L NaCl + 1 mol/L NaCl
Concentration of solutions

Fig. 11. Variation of potential (EPEAK) and current density (IPEAK) of the anodic peak as a function of the concentration of the solution

0,6

Alloy Alloy
Alloy Alloy
0,5 800H 600
22 201

0,4
EPEAK (VSCE)

0,3

0,2

75ºC
0,1 1.148 mol/L NaHCO3
1.148 mol/L NaHCO3 + 0.1 mol/L NaCl
1.148 mol/L NaHCO3 + 1 mol/L NaCl
0,0

30 40 50 60 70 80 90 100

Nickel (%)

Fig. 12. Variation of the potential of the anodic peak as a function of nickel content in the alloys
214 N. Zadorozne et al. / Procedia Materials Science 1 (2012) 207 – 214

5. Summary and Conclusions

Previous studies related the susceptibility to SCC of alloy 22 to the appearance of an anodic peak that
occurs in the pre-transpassivity region. This paper found that alloys 22, 600 and 800H had an anodic peak
when tested in media containing bicarbonate and chloride ions. The anodic peak became more evident at the
higher test temperatures. Alloy 201 showed an anodic peak prior to transpassivity only when the test
temperature was 90ºC.
The potential at which the anodic peak appeared increased when the test temperature decreased. This
behavior occurred for all the alloys and all the environments studied. The current density presented by the
anodic peak remained approximately constant by varying the temperature in the case of alloy 22 and 600.
Alloy 800H showed a decrease in the current density of the anodic peak with decreasing test temperature.
The addition of chloride to the base bicarbonate solution resulted in an increase in the value of EPEAK in
alloys 22 and 600.For alloy 22, IPEAK increased when the concentration of chloride in the solution increased.
The anodic peak was not present in the alloy with the lowest content of alloying elements (alloy 201:
99.5%Ni). It is assumed that the anodic peak may be related to the oxidation of one or more of the alloying
elements of alloy 22 (e.g. Cr. Mo).

References

ASTM G5-94, 2004, Standard Reference Test Method for Making Potentiostatic and Potentiodynamic Anodic Polarization
Measurements” in Annual Book of ASTM Standards, vol. 03.02 (West Conshohocken, PA: ASTM Intl.), pp. 53-64
Carranza, R. M., 2008, The Crevice Corrosion of Alloy 22 in the Yucca Mountain Nuclear Waste Repository, JOM, pp. 58-65
Chiang, K.T., Dunn D.S., Cragnolino G.A., 2007, Effect of simulated groundwater chemistry on stress corrosion cracking of Alloy 22,
Corrosion 63/10 940-950.
Chiang K. T., Dunn, D. S., Cragnolino, G. A., 2006, The Combined Effect of Bicarbonate and Chloride Ions on the Stress Corrosion
Cracking Susceptibility of Alloy 22, NACE Corrosion/2006, Paper 06506.
Cragnolino, G. A., Dunn, D. S., Pan, Y.-M., 2002, Localized corrosion susceptibility of alloy 22 as a waste package container material, in
Scientific Basis for Nuclear Waste Management XXV, Vol. 713, pp. 53-60
Gordon, G.M., 2002, “Corrosion Considerations Related to Permanent Disposal of High-. Level Radioactive Waste,” Corrosion, 58,
No.10, 811
King, K. J., Wong, L. L., Estill, J. C., Rebak, R. B., 2004, Slow Strain Rate Testing of Alloy 22 in Simulated Concentrated Ground
Waters, NACE Corrosion/2004, Paper 04548
Rebak, R. B., 2009 ,Corrosion testing of nickel and titanium alloys for nuclear waste disposition, Corrosion, 65, No. 4, pp. 252-271
Rebak, R. B., 2011, Environmental Degradation of Engineered Barrier Materials in Nuclear Waste Repositories, Chapter 36 in Uhlig’s
Corrosion Handbook, 3rd Edition by R. Winston Revie, John Wiley & Sons, 2011, pp. 503-516
Rebak, R. B., 2005, Factors Affecting the Crevice Corrosion Susceptibility of Alloy 22, NACE Corrosion/2005, paper 05610,
Rebak, R. B., 2003, Metallurgical effects on the corrosion behavior of nickel alloys, ASM Metals Handbook, Vol. 13A, Corrosion:
Fundamental, Testing, and Protection, pp 279-286.

You might also like