Top Banner
The University of Manchester Research Environmentally-assisted cracking of type 316L austenitic stainless steel in low pressure hydrogen steam environments DOI: 10.1016/j.prostr.2019.08.058 Document Version Final published version Link to publication record in Manchester Research Explorer Citation for published version (APA): Janoušek, J., Scenini, F., Volpe, L., Hojná, A., & Grace Burke, M. (2019). Environmentally-assisted cracking of type 316L austenitic stainless steel in low pressure hydrogen steam environments. Procedia Structural Integrity, 17, 440-447. https://doi.org/10.1016/j.prostr.2019.08.058 Published in: Procedia Structural Integrity Citing this paper Please note that where the full-text provided on Manchester Research Explorer is the Author Accepted Manuscript or Proof version this may differ from the final Published version. If citing, it is advised that you check and use the publisher's definitive version. General rights Copyright and moral rights for the publications made accessible in the Research Explorer are retained by the authors and/or other copyright owners and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights. Takedown policy If you believe that this document breaches copyright please refer to the University of Manchester’s Takedown Procedures [http://man.ac.uk/04Y6Bo] or contact [email protected] providing relevant details, so we can investigate your claim. Download date:16. Feb. 2022
9

Environmentally-Assisted Cracking of Type 316L Austenitic ...

Feb 16, 2022

Download

Documents

dariahiddleston
Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: Environmentally-Assisted Cracking of Type 316L Austenitic ...

The University of Manchester Research

Environmentally-assisted cracking of type 316L austeniticstainless steel in low pressure hydrogen steamenvironmentsDOI:10.1016/j.prostr.2019.08.058

Document VersionFinal published version

Link to publication record in Manchester Research Explorer

Citation for published version (APA):Janoušek, J., Scenini, F., Volpe, L., Hojná, A., & Grace Burke, M. (2019). Environmentally-assisted cracking oftype 316L austenitic stainless steel in low pressure hydrogen steam environments. Procedia Structural Integrity,17, 440-447. https://doi.org/10.1016/j.prostr.2019.08.058

Published in:Procedia Structural Integrity

Citing this paperPlease note that where the full-text provided on Manchester Research Explorer is the Author Accepted Manuscriptor Proof version this may differ from the final Published version. If citing, it is advised that you check and use thepublisher's definitive version.

General rightsCopyright and moral rights for the publications made accessible in the Research Explorer are retained by theauthors and/or other copyright owners and it is a condition of accessing publications that users recognise andabide by the legal requirements associated with these rights.

Takedown policyIf you believe that this document breaches copyright please refer to the University of Manchester’s TakedownProcedures [http://man.ac.uk/04Y6Bo] or contact [email protected] providingrelevant details, so we can investigate your claim.

Download date:16. Feb. 2022

Page 2: Environmentally-Assisted Cracking of Type 316L Austenitic ...

ScienceDirect

Available online at www.sciencedirect.com

Procedia Structural Integrity 17 (2019) 440–447

2452-3216 2019 The Authors. Published by Elsevier B.V.Peer-review under responsibility of the ICSI 2019 organizers.10.1016/j.prostr.2019.08.058

10.1016/j.prostr.2019.08.058 2452-3216

© 2019 The Authors. Published by Elsevier B.V.Peer-review under responsibility of the ICSI 2019 organizers.

Available online at www.sciencedirect.com

ScienceDirect

Structural Integrity Procedia 00 (2019) 000–000 www.elsevier.com/locate/procedia

2452-3216 © 2019 The Authors. Published by Elsevier B.V. Peer-review under responsibility of the ICSI 2019 organizers.

ICSI 2019 The 3rd International Conference on Structural Integrity

Environmentally-Assisted Cracking of Type 316L Austenitic Stainless Steel in Low Pressure Hydrogen Steam Environments

Jaromír Janoušeka, b *, Fabio Sceninic, Liberato Volpec, Anna Hojnáb, M. Grace Burkec aResearch Centre Rez, Hlavní 130, 250 68 Husinec-Řež, Czech Republic

bUniversity of West Bohemia, Department of Material Science and Technology, Univerzitní 22, 306 14 Plzeň, Czech Republic cThe University of Manchester, Material Performance Centre, Manchester, United Kingdom

Abstract

A low pressure, superheated hydrogen-steam system has been used to accelerate the oxidation kinetics while keeping the electrochemical conditions similar to those of the primary water in a pressurized water reactor. The initiation has been investigated using a Constant Extension Rate Tensile (CERT) test. Tests were performed on flat tapered specimens made from Type 316L austenitic stainless steel with strain rates of 2×10-6 and 2×10-8 ms-1 at room temperature and at an elevated temperature of 350 °C. R = 1/6 was chosen as a more oxidizing environment and R = 6 was selected as a more reducing environment, where the parameter R represents the ratio between the oxygen partial pressure at the Ni/NiO transition and the oxygen partial pressure. Different exposures (1 day and 5 days) prior to loading were investigated post-test evaluation by scanning electron microscopy. © 2019 The Authors. Published by Elsevier B.V. Peer-review under responsibility of the ICSI 2019 organizers.

Keywords: Environmentally Assisted Cracking; hydrogenated steam; oxidation; austenitic steel

1. Introduction

Environmentally-assisted cracking (EAC) is, according to Raja and Shoji (2011), the dominant issue in determining the reliability of most commercial equipment and applications that are controlled by the interactions between structural materials, cooling environments and operating stresses. Primary water stress corrosion cracking (SCC) is a form of EAC that can occur in essentially pure hydrogenated water at elevated temperatures. Hydrogen gas promotes EAC in

* Corresponding author. Tel.: +420720737082.

E-mail address: [email protected]

Available online at www.sciencedirect.com

ScienceDirect

Structural Integrity Procedia 00 (2019) 000–000 www.elsevier.com/locate/procedia

2452-3216 © 2019 The Authors. Published by Elsevier B.V. Peer-review under responsibility of the ICSI 2019 organizers.

ICSI 2019 The 3rd International Conference on Structural Integrity

Environmentally-Assisted Cracking of Type 316L Austenitic Stainless Steel in Low Pressure Hydrogen Steam Environments

Jaromír Janoušeka, b *, Fabio Sceninic, Liberato Volpec, Anna Hojnáb, M. Grace Burkec aResearch Centre Rez, Hlavní 130, 250 68 Husinec-Řež, Czech Republic

bUniversity of West Bohemia, Department of Material Science and Technology, Univerzitní 22, 306 14 Plzeň, Czech Republic cThe University of Manchester, Material Performance Centre, Manchester, United Kingdom

Abstract

A low pressure, superheated hydrogen-steam system has been used to accelerate the oxidation kinetics while keeping the electrochemical conditions similar to those of the primary water in a pressurized water reactor. The initiation has been investigated using a Constant Extension Rate Tensile (CERT) test. Tests were performed on flat tapered specimens made from Type 316L austenitic stainless steel with strain rates of 2×10-6 and 2×10-8 ms-1 at room temperature and at an elevated temperature of 350 °C. R = 1/6 was chosen as a more oxidizing environment and R = 6 was selected as a more reducing environment, where the parameter R represents the ratio between the oxygen partial pressure at the Ni/NiO transition and the oxygen partial pressure. Different exposures (1 day and 5 days) prior to loading were investigated post-test evaluation by scanning electron microscopy. © 2019 The Authors. Published by Elsevier B.V. Peer-review under responsibility of the ICSI 2019 organizers.

Keywords: Environmentally Assisted Cracking; hydrogenated steam; oxidation; austenitic steel

1. Introduction

Environmentally-assisted cracking (EAC) is, according to Raja and Shoji (2011), the dominant issue in determining the reliability of most commercial equipment and applications that are controlled by the interactions between structural materials, cooling environments and operating stresses. Primary water stress corrosion cracking (SCC) is a form of EAC that can occur in essentially pure hydrogenated water at elevated temperatures. Hydrogen gas promotes EAC in

* Corresponding author. Tel.: +420720737082.

E-mail address: [email protected]

2 Jaromír Janoušek / Structural Integrity Procedia 00 (2019) 000–000

high strength steels, which can initiate on smooth surfaces and requires no pre-existing defects such as pits, intergranular penetrations or mechanical defects. The 300 series Austenitic Stainless Steels (ASS) are widely used in the nuclear industry due to their reliable long-term performance in high-temperature water. On the other hand, according to Shoji (2003) and Couvant et al. (2006), some cases of EAC have occurred in the components of boiling water reactors (BWRs) and also in pressurized water reactors (PWRs) due to a hardened surface/subsurface layer induced during the fabrication process. The EAC degradation can develop during 20-30 years of operation conditions and it can be studied using accelerated testing in the laboratory.

The sensitivity to EAC of various surface treatments applied to ASS has been studied by Turnbull et al. (2011). It was shown that it was related to the high residual stress as well as the ultrafine-grained and deformed layer that extended several microns under the surface. The actual stress needed to initiate an EAC crack is likely a sum of the applied and residual stresses.

According to Couvant et al. (2009) it had been recognized that EAC initiation of 304L/316L SS in water cooling systems operating at approximately 300 °C was accelerated by increasing the water temperature up to 360 °C. Accelerated oxidation led to the formation of an outer Cr-enriched oxide and an inner Ni-rich oxide layer as well as grain boundary oxidation. Couvant et al. (2006) and Persaud et al. (2014) showed that an intergranular crack could initiate after fracture of the oxide or metal-oxide interface. With increasing temperature, higher acceleration of EAC and shorter times to initiation are expected. Under laboratory conditions, one can use temperatures up to 480 °C with steam. In fact, Economy et al. (1987) showed a monotonic dependence of SCC initiation time in both pressurized water and superheated steam at 368 °C, suggesting that the SCC initiation mechanism is similar for both environments.

In the paper, the EAC tests have been accelerated by utilizing three factors: slow strain rate, a hydrogenated steam environment and by increasing the temperature. Dominant acceleration via a constant extension rate test has been employed. Moreover, tapered-shape tensile specimens have been used, which permits us to examine a range of stresses and strains simultaneously on one specimen. This type of accelerated EAC test was developed by Yu et al. (1989) and recently updated by Berger et al. (2016) as a part of the “Mitigation of Crack Initiation” (MICRIN) project.

Nomenclature

EAC Environmentally Assisted Cracking SCC Stress corrosion cracking ASS Austenitic Stainless Steel Ra Arithmetical mean roughness CERT Constant Extension Rate Tensile PWR Pressurized Water Reactor SEM Scanning electron microscope BWR Boiled water reactors YS Yield strength, MPa UTS Ultimate tensile strength, MPa p Partial pressure, MPa R Ratio of partial pressures

2. Experiment

2.1. Material

This study was performed using 316L Austenitic Stainless Steel (ASS) produced by Industeel, Alcelor group for the IP EUROTRANS project (Table 1, Table 2). The steel was provided as 15 mm thick hot-rolled and heat-treated plates. A solution anneal was performed at 1050-1100 °C in air. The as-received microstructure consisted of equiaxed austenite grains and about 5% of δ-ferrite stringers oriented in the rolling direction.

Table 1. Chemical composition of 316L ASS (wt. %).

Fe Cr Mo Ni C Si Mn P S Al Cu Ti V N Bal. 16.69 2.08 9.97 0.018 0.64 1.84 0.027 0.004 0.018 0.23 0.006 0.07 0.029

Page 3: Environmentally-Assisted Cracking of Type 316L Austenitic ...

Jaromír Janoušek et al. / Procedia Structural Integrity 17 (2019) 440–447 441

Available online at www.sciencedirect.com

ScienceDirect

Structural Integrity Procedia 00 (2019) 000–000 www.elsevier.com/locate/procedia

2452-3216 © 2019 The Authors. Published by Elsevier B.V. Peer-review under responsibility of the ICSI 2019 organizers.

ICSI 2019 The 3rd International Conference on Structural Integrity

Environmentally-Assisted Cracking of Type 316L Austenitic Stainless Steel in Low Pressure Hydrogen Steam Environments

Jaromír Janoušeka, b *, Fabio Sceninic, Liberato Volpec, Anna Hojnáb, M. Grace Burkec aResearch Centre Rez, Hlavní 130, 250 68 Husinec-Řež, Czech Republic

bUniversity of West Bohemia, Department of Material Science and Technology, Univerzitní 22, 306 14 Plzeň, Czech Republic cThe University of Manchester, Material Performance Centre, Manchester, United Kingdom

Abstract

A low pressure, superheated hydrogen-steam system has been used to accelerate the oxidation kinetics while keeping the electrochemical conditions similar to those of the primary water in a pressurized water reactor. The initiation has been investigated using a Constant Extension Rate Tensile (CERT) test. Tests were performed on flat tapered specimens made from Type 316L austenitic stainless steel with strain rates of 2×10-6 and 2×10-8 ms-1 at room temperature and at an elevated temperature of 350 °C. R = 1/6 was chosen as a more oxidizing environment and R = 6 was selected as a more reducing environment, where the parameter R represents the ratio between the oxygen partial pressure at the Ni/NiO transition and the oxygen partial pressure. Different exposures (1 day and 5 days) prior to loading were investigated post-test evaluation by scanning electron microscopy. © 2019 The Authors. Published by Elsevier B.V. Peer-review under responsibility of the ICSI 2019 organizers.

Keywords: Environmentally Assisted Cracking; hydrogenated steam; oxidation; austenitic steel

1. Introduction

Environmentally-assisted cracking (EAC) is, according to Raja and Shoji (2011), the dominant issue in determining the reliability of most commercial equipment and applications that are controlled by the interactions between structural materials, cooling environments and operating stresses. Primary water stress corrosion cracking (SCC) is a form of EAC that can occur in essentially pure hydrogenated water at elevated temperatures. Hydrogen gas promotes EAC in

* Corresponding author. Tel.: +420720737082.

E-mail address: [email protected]

Available online at www.sciencedirect.com

ScienceDirect

Structural Integrity Procedia 00 (2019) 000–000 www.elsevier.com/locate/procedia

2452-3216 © 2019 The Authors. Published by Elsevier B.V. Peer-review under responsibility of the ICSI 2019 organizers.

ICSI 2019 The 3rd International Conference on Structural Integrity

Environmentally-Assisted Cracking of Type 316L Austenitic Stainless Steel in Low Pressure Hydrogen Steam Environments

Jaromír Janoušeka, b *, Fabio Sceninic, Liberato Volpec, Anna Hojnáb, M. Grace Burkec aResearch Centre Rez, Hlavní 130, 250 68 Husinec-Řež, Czech Republic

bUniversity of West Bohemia, Department of Material Science and Technology, Univerzitní 22, 306 14 Plzeň, Czech Republic cThe University of Manchester, Material Performance Centre, Manchester, United Kingdom

Abstract

A low pressure, superheated hydrogen-steam system has been used to accelerate the oxidation kinetics while keeping the electrochemical conditions similar to those of the primary water in a pressurized water reactor. The initiation has been investigated using a Constant Extension Rate Tensile (CERT) test. Tests were performed on flat tapered specimens made from Type 316L austenitic stainless steel with strain rates of 2×10-6 and 2×10-8 ms-1 at room temperature and at an elevated temperature of 350 °C. R = 1/6 was chosen as a more oxidizing environment and R = 6 was selected as a more reducing environment, where the parameter R represents the ratio between the oxygen partial pressure at the Ni/NiO transition and the oxygen partial pressure. Different exposures (1 day and 5 days) prior to loading were investigated post-test evaluation by scanning electron microscopy. © 2019 The Authors. Published by Elsevier B.V. Peer-review under responsibility of the ICSI 2019 organizers.

Keywords: Environmentally Assisted Cracking; hydrogenated steam; oxidation; austenitic steel

1. Introduction

Environmentally-assisted cracking (EAC) is, according to Raja and Shoji (2011), the dominant issue in determining the reliability of most commercial equipment and applications that are controlled by the interactions between structural materials, cooling environments and operating stresses. Primary water stress corrosion cracking (SCC) is a form of EAC that can occur in essentially pure hydrogenated water at elevated temperatures. Hydrogen gas promotes EAC in

* Corresponding author. Tel.: +420720737082.

E-mail address: [email protected]

2 Jaromír Janoušek / Structural Integrity Procedia 00 (2019) 000–000

high strength steels, which can initiate on smooth surfaces and requires no pre-existing defects such as pits, intergranular penetrations or mechanical defects. The 300 series Austenitic Stainless Steels (ASS) are widely used in the nuclear industry due to their reliable long-term performance in high-temperature water. On the other hand, according to Shoji (2003) and Couvant et al. (2006), some cases of EAC have occurred in the components of boiling water reactors (BWRs) and also in pressurized water reactors (PWRs) due to a hardened surface/subsurface layer induced during the fabrication process. The EAC degradation can develop during 20-30 years of operation conditions and it can be studied using accelerated testing in the laboratory.

The sensitivity to EAC of various surface treatments applied to ASS has been studied by Turnbull et al. (2011). It was shown that it was related to the high residual stress as well as the ultrafine-grained and deformed layer that extended several microns under the surface. The actual stress needed to initiate an EAC crack is likely a sum of the applied and residual stresses.

According to Couvant et al. (2009) it had been recognized that EAC initiation of 304L/316L SS in water cooling systems operating at approximately 300 °C was accelerated by increasing the water temperature up to 360 °C. Accelerated oxidation led to the formation of an outer Cr-enriched oxide and an inner Ni-rich oxide layer as well as grain boundary oxidation. Couvant et al. (2006) and Persaud et al. (2014) showed that an intergranular crack could initiate after fracture of the oxide or metal-oxide interface. With increasing temperature, higher acceleration of EAC and shorter times to initiation are expected. Under laboratory conditions, one can use temperatures up to 480 °C with steam. In fact, Economy et al. (1987) showed a monotonic dependence of SCC initiation time in both pressurized water and superheated steam at 368 °C, suggesting that the SCC initiation mechanism is similar for both environments.

In the paper, the EAC tests have been accelerated by utilizing three factors: slow strain rate, a hydrogenated steam environment and by increasing the temperature. Dominant acceleration via a constant extension rate test has been employed. Moreover, tapered-shape tensile specimens have been used, which permits us to examine a range of stresses and strains simultaneously on one specimen. This type of accelerated EAC test was developed by Yu et al. (1989) and recently updated by Berger et al. (2016) as a part of the “Mitigation of Crack Initiation” (MICRIN) project.

Nomenclature

EAC Environmentally Assisted Cracking SCC Stress corrosion cracking ASS Austenitic Stainless Steel Ra Arithmetical mean roughness CERT Constant Extension Rate Tensile PWR Pressurized Water Reactor SEM Scanning electron microscope BWR Boiled water reactors YS Yield strength, MPa UTS Ultimate tensile strength, MPa p Partial pressure, MPa R Ratio of partial pressures

2. Experiment

2.1. Material

This study was performed using 316L Austenitic Stainless Steel (ASS) produced by Industeel, Alcelor group for the IP EUROTRANS project (Table 1, Table 2). The steel was provided as 15 mm thick hot-rolled and heat-treated plates. A solution anneal was performed at 1050-1100 °C in air. The as-received microstructure consisted of equiaxed austenite grains and about 5% of δ-ferrite stringers oriented in the rolling direction.

Table 1. Chemical composition of 316L ASS (wt. %).

Fe Cr Mo Ni C Si Mn P S Al Cu Ti V N Bal. 16.69 2.08 9.97 0.018 0.64 1.84 0.027 0.004 0.018 0.23 0.006 0.07 0.029

Page 4: Environmentally-Assisted Cracking of Type 316L Austenitic ...

442 Jaromír Janoušek et al. / Procedia Structural Integrity 17 (2019) 440–447 Jaromír Janoušek/ Structural Integrity Procedia 00 (2019) 000–000 3

Table 2. Mechanical properties of 316L ASS.

T [°C]

E [GPa]

YS [MPa]

UTS [MPa]

Ag [%]

A30 [%]

Z [%]

25 202 251 567 44 60 80 350 143 168 421 26 38 66

2.2. Specimen

Flat tapered specimens (see Figure 1) with a thickness of 3 mm, width from 4 mm in the narrowest area to 6.4 mm in the widest part, and a gauge length of 26 mm were cut using electrical discharge machining (EDM) with the longer side parallel to the rolling direction of the plate. The parallel flat surfaces of the specimens were subjected to different surface finishes: one was manually ground to 500-grit in the direction parallel to the load axis, and the other one was polished using 1 µm diamond paste (see Figure 2).

Fig. 1. Drawing of tapered test specimen with polished and ground surface.

The arithmetical mean roughness (Ra) of the polished and ground surfaces were 0.005 µm and 0.032 µm respectively. The HV0.01 micro-hardness measured at about 20 µm below the polished and ground surface was 1.94 GPa (198HV0.01), and 2.05 GPa (209HV0.01) respectively. The 3° taper creates a variation of stress and strain along the gauge length during mechanical testing. Maximum stress is always achieved in the minimum of the cross section; the stress level in the area close to the end of the widest part stays elastic and does not overcome the yield strength. This allows the identification of threshold stress and strain conditions for the crack initiation within a single test specimen.

Fig. 2. Two different surface finishes of tapered specimen: polished (Ra = 0.005 μm) and ground (Ra = 0.032 μm) before testing.

4 Jaromír Janoušek / Structural Integrity Procedia 00 (2019) 000–000

2.3. Equipment

The cell for the low pressure superheated H2-steam environment has already been described by Janoušek et al. (2018). The cell was installed on an electromechanical creep testing machine Kappa SS-CF (see Figure 3) with a load capacity of up to 100 kN and with a speed range from 1 µm/h to 100 mm/min. The test chamber cover was tightened to the vessel using strength bolts and packed with a torus seal. The tested sample was placed in the test chamber cover with prismatic reductions. The filling mixture system consisted of a storage tank, a dosing chromatographic pump, a steam generator, Ar+H2 gas dosing system and a blender. The gas mixture Ar+H2 was mixed with steam in the blender and heated to 200 °C at which point it was admitted into the test chamber. Using a coiled heat exchanger the mixture could be heated up to 480 °C. The cooler and condenser formed the piping system wherein the gaseous mixture was cooled, and as the temperature dropped below the boiling point it condensed and proceeded to the air-leak chamber. The air-leak chamber had a free level at a height of approximately 1 metre so that it hydrostatically maintained an internal overpressure of approximately 0.1 bar. This oxidation system was originally developed by Scenini et al. (2005) and subsequently used by several laboratories.

Fig. 3. Modification of electromechanical creep testing machine Kappa SS-CF with the corrosion cell for low pressure superheated H2-steam environment and inserted broken specimen after CERT loading.

2.4. Test technique

The CERT corrosion-mechanical test technique was employed in this study. This technique uses uniaxial tensile testing performed with a very slow constant extension rate. Here, the loading was performed with two constant stroke rates of 2×10-6 m·s-1 (S1) and 2×10-8 m·s-1 (S2) corresponding to strain rates of 1×10-4 and 1×10-6 s-1 at the minimum cross section of the tapered specimens.

The oxygen partial pressure was controlled by manipulating the steam-to-H2 ratio (Rsteam/H2). The relationships for this method are described in Janoušek et al. (2018). The mixture 6% H2 + 94% Ar was used. To simulate different oxidizing and reducing environments with respect to the Ni/NiO transition: the oxygen partial pressure (pO2) was varied by changing the H2 partial pressure (e.g. by increasing the H2 flow and hence the H2 partial pressure, the redox potential decreases). The complete thermodynamics of the H2-steam environment can be described using the parameter R, which represents the ratio between the oxygen partial pressure at the Ni/NiO transition (pO2 Ni/NiO) and the oxygen partial pressure pO2 according to equation (1). For values lower than 1, NiO is stable (oxidizing environment), while for values higher than 1, Ni is stable (reducing environment).

Page 5: Environmentally-Assisted Cracking of Type 316L Austenitic ...

Jaromír Janoušek et al. / Procedia Structural Integrity 17 (2019) 440–447 443 Jaromír Janoušek/ Structural Integrity Procedia 00 (2019) 000–000 3

Table 2. Mechanical properties of 316L ASS.

T [°C]

E [GPa]

YS [MPa]

UTS [MPa]

Ag [%]

A30 [%]

Z [%]

25 202 251 567 44 60 80 350 143 168 421 26 38 66

2.2. Specimen

Flat tapered specimens (see Figure 1) with a thickness of 3 mm, width from 4 mm in the narrowest area to 6.4 mm in the widest part, and a gauge length of 26 mm were cut using electrical discharge machining (EDM) with the longer side parallel to the rolling direction of the plate. The parallel flat surfaces of the specimens were subjected to different surface finishes: one was manually ground to 500-grit in the direction parallel to the load axis, and the other one was polished using 1 µm diamond paste (see Figure 2).

Fig. 1. Drawing of tapered test specimen with polished and ground surface.

The arithmetical mean roughness (Ra) of the polished and ground surfaces were 0.005 µm and 0.032 µm respectively. The HV0.01 micro-hardness measured at about 20 µm below the polished and ground surface was 1.94 GPa (198HV0.01), and 2.05 GPa (209HV0.01) respectively. The 3° taper creates a variation of stress and strain along the gauge length during mechanical testing. Maximum stress is always achieved in the minimum of the cross section; the stress level in the area close to the end of the widest part stays elastic and does not overcome the yield strength. This allows the identification of threshold stress and strain conditions for the crack initiation within a single test specimen.

Fig. 2. Two different surface finishes of tapered specimen: polished (Ra = 0.005 μm) and ground (Ra = 0.032 μm) before testing.

4 Jaromír Janoušek / Structural Integrity Procedia 00 (2019) 000–000

2.3. Equipment

The cell for the low pressure superheated H2-steam environment has already been described by Janoušek et al. (2018). The cell was installed on an electromechanical creep testing machine Kappa SS-CF (see Figure 3) with a load capacity of up to 100 kN and with a speed range from 1 µm/h to 100 mm/min. The test chamber cover was tightened to the vessel using strength bolts and packed with a torus seal. The tested sample was placed in the test chamber cover with prismatic reductions. The filling mixture system consisted of a storage tank, a dosing chromatographic pump, a steam generator, Ar+H2 gas dosing system and a blender. The gas mixture Ar+H2 was mixed with steam in the blender and heated to 200 °C at which point it was admitted into the test chamber. Using a coiled heat exchanger the mixture could be heated up to 480 °C. The cooler and condenser formed the piping system wherein the gaseous mixture was cooled, and as the temperature dropped below the boiling point it condensed and proceeded to the air-leak chamber. The air-leak chamber had a free level at a height of approximately 1 metre so that it hydrostatically maintained an internal overpressure of approximately 0.1 bar. This oxidation system was originally developed by Scenini et al. (2005) and subsequently used by several laboratories.

Fig. 3. Modification of electromechanical creep testing machine Kappa SS-CF with the corrosion cell for low pressure superheated H2-steam environment and inserted broken specimen after CERT loading.

2.4. Test technique

The CERT corrosion-mechanical test technique was employed in this study. This technique uses uniaxial tensile testing performed with a very slow constant extension rate. Here, the loading was performed with two constant stroke rates of 2×10-6 m·s-1 (S1) and 2×10-8 m·s-1 (S2) corresponding to strain rates of 1×10-4 and 1×10-6 s-1 at the minimum cross section of the tapered specimens.

The oxygen partial pressure was controlled by manipulating the steam-to-H2 ratio (Rsteam/H2). The relationships for this method are described in Janoušek et al. (2018). The mixture 6% H2 + 94% Ar was used. To simulate different oxidizing and reducing environments with respect to the Ni/NiO transition: the oxygen partial pressure (pO2) was varied by changing the H2 partial pressure (e.g. by increasing the H2 flow and hence the H2 partial pressure, the redox potential decreases). The complete thermodynamics of the H2-steam environment can be described using the parameter R, which represents the ratio between the oxygen partial pressure at the Ni/NiO transition (pO2 Ni/NiO) and the oxygen partial pressure pO2 according to equation (1). For values lower than 1, NiO is stable (oxidizing environment), while for values higher than 1, Ni is stable (reducing environment).

Page 6: Environmentally-Assisted Cracking of Type 316L Austenitic ...

444 Jaromír Janoušek et al. / Procedia Structural Integrity 17 (2019) 440–447 Jaromír Janoušek/ Structural Integrity Procedia 00 (2019) 000–000 5

𝑅𝑅 = 𝑝𝑝𝑂𝑂2 𝑁𝑁𝑁𝑁/𝑁𝑁𝑁𝑁𝑂𝑂𝑝𝑝𝑂𝑂2

(1)

The specimens were tested in a low-pressure H2-steam environment at a temperature of 350 °C to failure. R = 1/6 was chosen as a more oxidizing environment, which corresponds to a water flow rate of 2.22 mL/min, a gas mixture (6% H2 + 94% Ar) of 50 cc/min, a steam-to-H2 ratio of 983 and an oxygen partial pressure of 2.48×10−30 atm. A more reducing environment was R = 6 which corresponds to a water flow rate of 0.37 mL/min, a gas mixture (6% H2 + 94% Ar) of 50 cc/min, a steam-to-H2 ratio of 164 and an oxygen partial pressure of 6.90×10−32 atm. Ultrahigh purity water with a conductivity of 0.055 μmS/cm was used. The tapered specimens were exposed to the steam for one day prior to CERT loading for the more reducing environment with R = 6. The time of exposure was five days prior to CERT loading for the more oxidizing environment with R = 1/6.

2.5. Characterization

The microstructure of the steel was evaluated using a TESCAN MIRA3 field emission gun scanning electron microscope (FEGSEM) and a TESCAN LYRA3 focused ion beam (FIB) - FEGSEM system. After completion of the SCC tests, specimens were examined to assess the fracture morphology. Site-specific cross-section specimens were prepared using the LYRA3 FIB-FEGSEM system.

3. Results

3.1. CERT curves

Figure 4 shows the complete results of the tests where each curve corresponds to one sample exposed to a specific environment. The first tensile test with a rate of 2×10-6 ms-1 was performed on the same rig at room temperature and 350 °C. To correctly interpret the results, it was necessary to perform tests without a sample at both temperatures. These correction curves have been subtracted from the curves with the specimens because the correction curves represent the resistance of the spring bellows. Both correction curves are the same, which means that the temperature has almost no effect on the spring bellows, because this part is located outside the chamber.

Fig. 4. Measured curves of the CERT loading for samples exposed to a specific environment.

6 Jaromír Janoušek / Structural Integrity Procedia 00 (2019) 000–000

For the tapered specimen tested at room temperature, the maximum stress at the minimum cross section was evaluated to be equal to 612 MPa (note Table 2 UTS value is 567 MPa). For the specimen tested at 350 °C the maximum stress was 458 MPa (Table 2 UTS value is 421 MPa). Both curves in the H2-steam environments showed greater elongation. The effect is likely higher for the curve in the reducing environment R = 6, than for the one oxidizing with R = 1/6. The corrected CERT curves for the specimens tested at 350 °C are plotted in Figure 5. All curves at 350 °C show oscillations of load, which are typical for dynamic strain ageing. Tests at the slower extension rate of 2×10-8 m·s-1 (S2) resulted in higher strengths.

Fig. 5. The CERT test curves of the tapered specimens in air and H2-steam at 350 °C after subtraction of the correction curve.

Small coupons of pure nickel (99.5 wt%) were used in each experiment to control the electrochemical potential with respect to the Ni/NiO transition of the high temperature environment. It was observed that an exposure of one day appeared to be insufficient for the material to be affected by the environments.

3.2. Post-test FEGSEM characterization

After testing, the specimens were evaluated to assess the extent of cracking. All specimens failed by ductile fracture with extensive plastic deformation. Post-test evaluation showed that the flat surfaces were clean, with no oxide particles. The polished and ground surfaces behaved similarly. The SEM observation of the fracture shows typical ductile fracture dimples. This confirmed that a one-day exposure is insufficient for the material to be affected by the environments.

Sporadic oxide particles of various sizes were observed at the surface for 5-day exposures. Plastic deformation of varying extents appeared in slip bands emerging from the surface. Several ductile cracks initiating from the slip bands are shown in Figure 6. EAC cracks with the characteristic fracture mode and the orientation perpendicular to the loading were not observed. Secondary electron micrographs revealed that appearance of the ground and polished surfaces was similar.

Page 7: Environmentally-Assisted Cracking of Type 316L Austenitic ...

Jaromír Janoušek et al. / Procedia Structural Integrity 17 (2019) 440–447 445 Jaromír Janoušek/ Structural Integrity Procedia 00 (2019) 000–000 5

𝑅𝑅 = 𝑝𝑝𝑂𝑂2 𝑁𝑁𝑁𝑁/𝑁𝑁𝑁𝑁𝑂𝑂𝑝𝑝𝑂𝑂2

(1)

The specimens were tested in a low-pressure H2-steam environment at a temperature of 350 °C to failure. R = 1/6 was chosen as a more oxidizing environment, which corresponds to a water flow rate of 2.22 mL/min, a gas mixture (6% H2 + 94% Ar) of 50 cc/min, a steam-to-H2 ratio of 983 and an oxygen partial pressure of 2.48×10−30 atm. A more reducing environment was R = 6 which corresponds to a water flow rate of 0.37 mL/min, a gas mixture (6% H2 + 94% Ar) of 50 cc/min, a steam-to-H2 ratio of 164 and an oxygen partial pressure of 6.90×10−32 atm. Ultrahigh purity water with a conductivity of 0.055 μmS/cm was used. The tapered specimens were exposed to the steam for one day prior to CERT loading for the more reducing environment with R = 6. The time of exposure was five days prior to CERT loading for the more oxidizing environment with R = 1/6.

2.5. Characterization

The microstructure of the steel was evaluated using a TESCAN MIRA3 field emission gun scanning electron microscope (FEGSEM) and a TESCAN LYRA3 focused ion beam (FIB) - FEGSEM system. After completion of the SCC tests, specimens were examined to assess the fracture morphology. Site-specific cross-section specimens were prepared using the LYRA3 FIB-FEGSEM system.

3. Results

3.1. CERT curves

Figure 4 shows the complete results of the tests where each curve corresponds to one sample exposed to a specific environment. The first tensile test with a rate of 2×10-6 ms-1 was performed on the same rig at room temperature and 350 °C. To correctly interpret the results, it was necessary to perform tests without a sample at both temperatures. These correction curves have been subtracted from the curves with the specimens because the correction curves represent the resistance of the spring bellows. Both correction curves are the same, which means that the temperature has almost no effect on the spring bellows, because this part is located outside the chamber.

Fig. 4. Measured curves of the CERT loading for samples exposed to a specific environment.

6 Jaromír Janoušek / Structural Integrity Procedia 00 (2019) 000–000

For the tapered specimen tested at room temperature, the maximum stress at the minimum cross section was evaluated to be equal to 612 MPa (note Table 2 UTS value is 567 MPa). For the specimen tested at 350 °C the maximum stress was 458 MPa (Table 2 UTS value is 421 MPa). Both curves in the H2-steam environments showed greater elongation. The effect is likely higher for the curve in the reducing environment R = 6, than for the one oxidizing with R = 1/6. The corrected CERT curves for the specimens tested at 350 °C are plotted in Figure 5. All curves at 350 °C show oscillations of load, which are typical for dynamic strain ageing. Tests at the slower extension rate of 2×10-8 m·s-1 (S2) resulted in higher strengths.

Fig. 5. The CERT test curves of the tapered specimens in air and H2-steam at 350 °C after subtraction of the correction curve.

Small coupons of pure nickel (99.5 wt%) were used in each experiment to control the electrochemical potential with respect to the Ni/NiO transition of the high temperature environment. It was observed that an exposure of one day appeared to be insufficient for the material to be affected by the environments.

3.2. Post-test FEGSEM characterization

After testing, the specimens were evaluated to assess the extent of cracking. All specimens failed by ductile fracture with extensive plastic deformation. Post-test evaluation showed that the flat surfaces were clean, with no oxide particles. The polished and ground surfaces behaved similarly. The SEM observation of the fracture shows typical ductile fracture dimples. This confirmed that a one-day exposure is insufficient for the material to be affected by the environments.

Sporadic oxide particles of various sizes were observed at the surface for 5-day exposures. Plastic deformation of varying extents appeared in slip bands emerging from the surface. Several ductile cracks initiating from the slip bands are shown in Figure 6. EAC cracks with the characteristic fracture mode and the orientation perpendicular to the loading were not observed. Secondary electron micrographs revealed that appearance of the ground and polished surfaces was similar.

Page 8: Environmentally-Assisted Cracking of Type 316L Austenitic ...

446 Jaromír Janoušek et al. / Procedia Structural Integrity 17 (2019) 440–447 Jaromír Janoušek/ Structural Integrity Procedia 00 (2019) 000–000 7

Fig. 6. Secondary electron micrographs of: (a) polished surface at the minimum cross-section after testing at a rate of 2×10-6 ms-1 in the H2-steam environment of R = 1/6 at 350 °C; and (b) Ground surface - cracks initiating from slip bands.

The appearance of the Nickel coupon surface exposed for 5 days at 350°C to the 6 times more oxidizing environment with respect to the Ni/NiO transition is shown in Figure 7. These secondary electron micrographs showed oxide particles localized on grain boundaries.

Fig. 7. Secondary electron micrographs of the surface of the pure nickel (99.5 wt%) coupon in the H2-steam environment of R = 1/6 at 350 °C for 5 days exposure.

4. Conclusion

A new low pressure superheated H2-steam system was used to perform Constant Extension Rate tests at room temperature and an elevated temperature of 350 °C with rates of 2×10-6 ms-1 and 2×10-8 ms-1. The special corrosion apparatus was built in cooperation with the University of Manchester and Škoda JS for the MEACTOS EU project. The following conclusions can be drawn from the test results: One day and five day exposures prior to loading the 316L steel specimen in both H2-steam environments at 350 °C did not lead to the initiation of EAC cracks. However, FEGSEM evaluation revealed some oxide particles localized

8 Jaromír Janoušek / Structural Integrity Procedia 00 (2019) 000–000

on the grain boundaries of the Nickel coupon exposed for 5 days at 350 °C to the 6 times more oxidizing environment with respect to the Ni/NiO transition.

All CERT curves at 350 °C exhibited oscillations of load, which are typical for dynamic strain ageing. Tests with the slower extension rate achieved higher strengths. The maximum stress at the minimum cross-section of the tapered tensile specimen was higher in comparison with published values. For the room temperature test the maximum stress was 612 MPa (table UTS value is 567 MPa). For the specimen tested at 350 °C the maximum stress was 458 MPa (table UTS value is 421 MPa). Both CERT curves measured in the H2-steam environments showed greater elongation.

Plastic deformation of varying extents appeared as slip bands emerging from the surface. Several ductile cracks initiated from the slip bands. The FEGSEM observation of the failed specimens revealed typical ductile fracture dimples.

Acknowledgements

This work has been undertaken within the SUSEN Project (established in the framework of the European Regional Development Fund (ERDF) in the project CZ.1.05/2.1.00/03.0108 and of the European Strategy Forum on Research Infrastructures (ESFRI) in the project CZ.02.1.01/0.0/0.0/15_008/0000293, which is financially supported by the Ministry of Education, Youth and Sports - project LM2015093 Infrastructure SUSEN. The support of Horizon 2020 MEACTOS GA No. 755151 is also acknowledged.

References

Raja, V. S. and Shoji, T., 2011. Stress corrosion cracking, Theory and practice, Woodhead Publishing Limited Shoji, T., 2003. Progress in the Mechanistic Understanding of BWR SCC and Its Implication to the Prediction of SCC Growth Behavior in Plants,

Proc. 11th Int. Conf. Environmental Degradation of Materials in Nuclear Systems, Stevenson, pp. 588-599. Couvant, T., Moulart, P., Legras, L., Bordes, P., Capelle, J., Rouillon, Y., Balon, T., 2006. PWSCC of austenitic stainless steels of heaters of

pressurizers, 6th Symposium Fontevraud: SFEN. Turnbull, A., Mingard, K., Lord, J. D., Roebuck, B., Tice, D.R., Mottershead, K.J., Fairweather, N.D., Bradbury, A.K., 2011. Sensitivity of stress

corrosion cracking of stainless steel to surface machining and grinding procedure, Corrosion Science 53, pp. 3398–3415. Couvant, T., Legras, L., Herbelin, A., Musienko, A., Ilevbare, G., Delafosse, D., Cailletaud, G., Hickling, J., 2009. Development of Understanding

of the Interaction between Localized Deformation and SCC of Austenitic Stainless Steels Exposed to Primary PWR Environment, 14th Int. Conf. on Environmental Degradation of Materials in Nuclear Power Systems, Virginia Beach.

Persaud, S. Y., Korinek, A., Huang, J., Botton, G. A., Newman, R. C., 2014. Internal oxidation of Alloy 600 exposed to hydrogenated steam and the beneficial effects of thermal treatment, Corrosion Science 86, 108–122.

Economy, G., Jacko, R. J., Pement, F. W., 1987. Igscc Behavior of Alloy 600 Steam-Generator Tubing in Water or Steam Tests above 360 C,” Corrosion, vol. 43, no. 12, pp. 727–73.

Yu, J., Xue, L J., Zhao, Z. J., Chi, G. X., Parkins, R.N., 1989. Determination of Stress Corrosion Crack Initiation Stress and Crack Velocities using Slowly Strained Tapered Specimens, Fatigue Fracture of Engineering Materials and Structures 12, pp. 481-493.

Berger, S., Kilian, R., Ritter, S., Ehrnstén, U., Bosch, R.-W., Perosanz Lopez, F.-J., 2016. Mitigation of crack initiation in LWRs (MICRIN+), Paper No. 65952, Eurocorr, Monpellier.

Janoušek, J., Scenini, F., Volpe, L., Hojná, A., Trojan, T., 2018. Instrumentation for SCC testing in low pressure superheated hydrogen steam environments, IOP Conference Series: Materials Science and Engineering, Volume 461, Issue 1.

Scenini, F., Newman, R.C., Cottis, R.A., Jacko, R.J., 2005. Alloy 600 oxidation studies related to PWSCC, Proceedings of the 12th International Symposium on Environmental Degradation of Materials in Nuclear Power System - Water Reactors, Minerals, Metals and Materials Society/AIME, pp.891–902.

Page 9: Environmentally-Assisted Cracking of Type 316L Austenitic ...

Jaromír Janoušek et al. / Procedia Structural Integrity 17 (2019) 440–447 447 Jaromír Janoušek/ Structural Integrity Procedia 00 (2019) 000–000 7

Fig. 6. Secondary electron micrographs of: (a) polished surface at the minimum cross-section after testing at a rate of 2×10-6 ms-1 in the H2-steam environment of R = 1/6 at 350 °C; and (b) Ground surface - cracks initiating from slip bands.

The appearance of the Nickel coupon surface exposed for 5 days at 350°C to the 6 times more oxidizing environment with respect to the Ni/NiO transition is shown in Figure 7. These secondary electron micrographs showed oxide particles localized on grain boundaries.

Fig. 7. Secondary electron micrographs of the surface of the pure nickel (99.5 wt%) coupon in the H2-steam environment of R = 1/6 at 350 °C for 5 days exposure.

4. Conclusion

A new low pressure superheated H2-steam system was used to perform Constant Extension Rate tests at room temperature and an elevated temperature of 350 °C with rates of 2×10-6 ms-1 and 2×10-8 ms-1. The special corrosion apparatus was built in cooperation with the University of Manchester and Škoda JS for the MEACTOS EU project. The following conclusions can be drawn from the test results: One day and five day exposures prior to loading the 316L steel specimen in both H2-steam environments at 350 °C did not lead to the initiation of EAC cracks. However, FEGSEM evaluation revealed some oxide particles localized

8 Jaromír Janoušek / Structural Integrity Procedia 00 (2019) 000–000

on the grain boundaries of the Nickel coupon exposed for 5 days at 350 °C to the 6 times more oxidizing environment with respect to the Ni/NiO transition.

All CERT curves at 350 °C exhibited oscillations of load, which are typical for dynamic strain ageing. Tests with the slower extension rate achieved higher strengths. The maximum stress at the minimum cross-section of the tapered tensile specimen was higher in comparison with published values. For the room temperature test the maximum stress was 612 MPa (table UTS value is 567 MPa). For the specimen tested at 350 °C the maximum stress was 458 MPa (table UTS value is 421 MPa). Both CERT curves measured in the H2-steam environments showed greater elongation.

Plastic deformation of varying extents appeared as slip bands emerging from the surface. Several ductile cracks initiated from the slip bands. The FEGSEM observation of the failed specimens revealed typical ductile fracture dimples.

Acknowledgements

This work has been undertaken within the SUSEN Project (established in the framework of the European Regional Development Fund (ERDF) in the project CZ.1.05/2.1.00/03.0108 and of the European Strategy Forum on Research Infrastructures (ESFRI) in the project CZ.02.1.01/0.0/0.0/15_008/0000293, which is financially supported by the Ministry of Education, Youth and Sports - project LM2015093 Infrastructure SUSEN. The support of Horizon 2020 MEACTOS GA No. 755151 is also acknowledged.

References

Raja, V. S. and Shoji, T., 2011. Stress corrosion cracking, Theory and practice, Woodhead Publishing Limited Shoji, T., 2003. Progress in the Mechanistic Understanding of BWR SCC and Its Implication to the Prediction of SCC Growth Behavior in Plants,

Proc. 11th Int. Conf. Environmental Degradation of Materials in Nuclear Systems, Stevenson, pp. 588-599. Couvant, T., Moulart, P., Legras, L., Bordes, P., Capelle, J., Rouillon, Y., Balon, T., 2006. PWSCC of austenitic stainless steels of heaters of

pressurizers, 6th Symposium Fontevraud: SFEN. Turnbull, A., Mingard, K., Lord, J. D., Roebuck, B., Tice, D.R., Mottershead, K.J., Fairweather, N.D., Bradbury, A.K., 2011. Sensitivity of stress

corrosion cracking of stainless steel to surface machining and grinding procedure, Corrosion Science 53, pp. 3398–3415. Couvant, T., Legras, L., Herbelin, A., Musienko, A., Ilevbare, G., Delafosse, D., Cailletaud, G., Hickling, J., 2009. Development of Understanding

of the Interaction between Localized Deformation and SCC of Austenitic Stainless Steels Exposed to Primary PWR Environment, 14th Int. Conf. on Environmental Degradation of Materials in Nuclear Power Systems, Virginia Beach.

Persaud, S. Y., Korinek, A., Huang, J., Botton, G. A., Newman, R. C., 2014. Internal oxidation of Alloy 600 exposed to hydrogenated steam and the beneficial effects of thermal treatment, Corrosion Science 86, 108–122.

Economy, G., Jacko, R. J., Pement, F. W., 1987. Igscc Behavior of Alloy 600 Steam-Generator Tubing in Water or Steam Tests above 360 C,” Corrosion, vol. 43, no. 12, pp. 727–73.

Yu, J., Xue, L J., Zhao, Z. J., Chi, G. X., Parkins, R.N., 1989. Determination of Stress Corrosion Crack Initiation Stress and Crack Velocities using Slowly Strained Tapered Specimens, Fatigue Fracture of Engineering Materials and Structures 12, pp. 481-493.

Berger, S., Kilian, R., Ritter, S., Ehrnstén, U., Bosch, R.-W., Perosanz Lopez, F.-J., 2016. Mitigation of crack initiation in LWRs (MICRIN+), Paper No. 65952, Eurocorr, Monpellier.

Janoušek, J., Scenini, F., Volpe, L., Hojná, A., Trojan, T., 2018. Instrumentation for SCC testing in low pressure superheated hydrogen steam environments, IOP Conference Series: Materials Science and Engineering, Volume 461, Issue 1.

Scenini, F., Newman, R.C., Cottis, R.A., Jacko, R.J., 2005. Alloy 600 oxidation studies related to PWSCC, Proceedings of the 12th International Symposium on Environmental Degradation of Materials in Nuclear Power System - Water Reactors, Minerals, Metals and Materials Society/AIME, pp.891–902.