Download as pdf or txt
Download as pdf or txt
You are on page 1of 266

Environmentally Assisted Fatigue (EAF)

Knowledge Gap Analysis


Update and Revision of the EAF Knowledge Gaps

2018 TECHNICAL REPORT

13321451
13321451
Environmentally Assisted Fatigue
(EAF) Knowledge Gap Analysis
Update and Revision of the EAF Knowledge Gaps
3002013214

Final Report, September 2018

EPRI Project Manager


K. Ahluwalia

All or a portion of the requirements of the EPRI Nuclear


Quality Assurance Program apply to this product.

ELECTRIC POWER RESEARCH INSTITUTE


3420 Hillview Avenue, Palo Alto, California 94304-1338 ▪ PO Box 10412, Palo Alto, California 94303-0813 ▪ USA
800.313.3774 ▪ 650.855.2121 ▪ askepri@epri.com ▪ www.epri.com

13321451
DISCLAIMER OF WARRANTIES AND LIMITATION OF LIABILITIES
THIS DOCUMENT WAS PREPARED BY THE ORGANIZATION(S) NAMED BELOW AS AN ACCOUNT
OF WORK SPONSORED OR COSPONSORED BY THE ELECTRIC POWER RESEARCH INSTITUTE,
INC. (EPRI). NEITHER EPRI, ANY MEMBER OF EPRI, ANY COSPONSOR, THE ORGANIZATION(S)
BELOW, NOR ANY PERSON ACTING ON BEHALF OF ANY OF THEM:
(A) MAKES ANY WARRANTY OR REPRESENTATION WHATSOEVER, EXPRESS OR IMPLIED, (I)
WITH RESPECT TO THE USE OF ANY INFORMATION, APPARATUS, METHOD, PROCESS, OR
SIMILAR ITEM DISCLOSED IN THIS DOCUMENT, INCLUDING MERCHANTABILITY AND FITNESS
FOR A PARTICULAR PURPOSE, OR (II) THAT SUCH USE DOES NOT INFRINGE ON OR
INTERFERE WITH PRIVATELY OWNED RIGHTS, INCLUDING ANY PARTY'S INTELLECTUAL
PROPERTY, OR (III) THAT THIS DOCUMENT IS SUITABLE TO ANY PARTICULAR USER'S
CIRCUMSTANCE; OR
(B) ASSUMES RESPONSIBILITY FOR ANY DAMAGES OR OTHER LIABILITY WHATSOEVER
(INCLUDING ANY CONSEQUENTIAL DAMAGES, EVEN IF EPRI OR ANY EPRI REPRESENTATIVE
HAS BEEN ADVISED OF THE POSSIBILITY OF SUCH DAMAGES) RESULTING FROM YOUR
SELECTION OR USE OF THIS DOCUMENT OR ANY INFORMATION, APPARATUS, METHOD,
PROCESS, OR SIMILAR ITEM DISCLOSED IN THIS DOCUMENT.
REFERENCE HEREIN TO ANY SPECIFIC COMMERCIAL PRODUCT, PROCESS, OR SERVICE BY
ITS TRADE NAME, TRADEMARK, MANUFACTURER, OR OTHERWISE, DOES NOT NECESSARILY
CONSTITUTE OR IMPLY ITS ENDORSEMENT, RECOMMENDATION, OR FAVORING BY EPRI.
THE FOLLOWING ORGANIZATION, UNDER CONTRACT TO EPRI, PREPARED THIS REPORT:
Amec Foster Wheeler

THE TECHNICAL CONTENTS OF THIS PRODUCT WERE NOT PREPARED IN ACCORDANCE WITH
THE EPRI QUALITY PROGRAM MANUAL THAT FULFILLS THE REQUIREMENTS OF 10 CFR 50,
APPENDIX B. THIS PRODUCT IS NOT SUBJECT TO THE REQUIREMENTS OF 10 CFR PART 21.

NOTE
For further information about EPRI, call the EPRI Customer Assistance Center at 800.313.3774 or
e-mail askepri@epri.com.

Electric Power Research Institute, EPRI, and TOGETHER…SHAPING THE FUTURE OF ELECTRICITY
are registered service marks of the Electric Power Research Institute, Inc.
Copyright © 2018 Electric Power Research Institute, Inc. All rights reserved.

13321451
ACKNOWLEDGMENTS

The following organization, under contract to the Electric Power Research Institute (EPRI),
prepared this report:
Amec Foster Wheeler
Booths Park, Chelford Road
Knutsford, Cheshire
WA16 8QZ
England
Principal Investigator
D. Tice
Reviewers
G. Quirk
N. Platts
A. Davison
Authors
A. McLennan
P. Gill

This report describes research sponsored by EPRI. EPRI staff members J. Smith and G. Stevens
contributed to the report’s development.

This publication is a corporate document that should be cited in the literature in the following
manner:
Environmentally Assisted Fatigue (EAF) Knowledge Gap Analysis: Update and Revision of the
EAF Knowledge Gaps. EPRI, Palo Alto, CA: 2018. 3002013214.
iii
13321451
13321451
ABSTRACT

Designers attempting nuclear plant life extension and new build are required to demonstrate safe
operation of the plant over its intended lifetime. This typically involves the application of
established fatigue initiation design curves. However, the existing rules and associated guidance
can predict very high cumulative usage factors for some components when environmentally
assisted fatigue (EAF) effects are considered. Furthermore, the high calculated usage factors
predicted are not reflected in operational experience, and there have been no known fatigue
failures to date attributed to low cycle thermal fatigue enhanced by EAF.
To understand the applicability of the currently available fatigue rules and EAF guidance, EPRI
identified gaps in the industry’s understanding of EAF. This was documented in the Knowledge
Gap Analysis report published in 2011 (EPRI product 1023012) and the Roadmap report
published in 2012 (EPRI product 1026724). Those documents highlighted the technical areas
where uncertainty existed and identified the research and development needed to address them.
This resulted in 47 Knowledge Gaps, which were reviewed by nuclear industry experts who
allocated priorities to the various gaps.
Since the Knowledge Gap report was published, there have been a number of significant research
activities worldwide, and a number of alternative methodologies for EAF assessment have been
proposed. Some of these activities specifically addressed some of the previously identified
Knowledge Gaps.
This report provides a detailed review of publicly available research results published since the
original Knowledge Gaps were developed in 2011, and provides an update to the Knowledge
Gaps. It also identifies that significant additional laboratory data has been generated for light
water reactor environments since the time of the prior report.
It should be noted that the need for a plant-representative test using plant-relevant geometries
and loading histories has been identified. As a result, there is now considerable worldwide
interest, supported by an international collaborative group of interested parties, in proceeding to
develop and perform an EAF test that is more representative of a plant component and
experiences plant-like transients. Such a test would provide significant additional insights and a
means of validating any proposed assessment methods.
The existing 47 Knowledge Gaps were reviewed and a revised, shorter list of 17 outstanding
issues that warrant further study was developed. The rationale for the new Knowledge Gap list
was to streamline the existing gaps by removing redundancy and overlap, as well as focusing the
remaining gaps on the major issues as identified by the nuclear industry.
Keywords
Crack growth EAF Environmentally assisted fatigue
Fatigue assessment Knowledge gap Light water reactor

v
13321451
13321451
EXECUTIVE SUMMARY

Deliverable Number: 3002013214


Product Type: Technical Report
Product Title: Environmentally Assisted Fatigue (EAF) Knowledge Gap Analysis:
Update and Revision of the EAF Knowledge Gaps

PRIMARY AUDIENCE: Designers and nuclear power plant operators


SECONDARY AUDIENCE: International nuclear research community

KEY RESEARCH QUESTION


The original Knowledge Gap Assessment (EPRI report 1023012) was produced in 2011 and provided a list
of gaps in the nuclear industry’s understanding of environmentally assisted fatigue (EAF). Since that
document was published, significant additional work has been performed by research institutes worldwide to
gain further understanding of this degradation mechanism, with the objective of finding effective and less
conservative methods with which to assess the fatigue lives of nuclear plant components. In light of the
significant amount of research done on this topic, the present report was prepared in order to provide an
update and refinement of the existing Knowledge Gaps (as they were previously defined in EPRI report
1023012).

RESEARCH OVERVIEW
The deleterious effect of light water reactor (LWR) environments on materials tested in the laboratory is well
known. However, it is unclear how the laboratory results, which used small-scale specimens under fully
reversed loading, apply to actual nuclear plant components. EAF currently poses a significant issue for nuclear
plant design for both new build and plant life extension, because of the high fatigue usage factors predicted
by current methodologies that incorporate environmental factors (Fen). Due to the issues presented by EAF,
a significant amount of research has been conducted by the international community. In 2011/2012, work was
done to assess the industry’s understanding of EAF and present a list of Knowledge Gaps with which to guide
international research for maximum benefit to the industry. The Knowledge Gap document found 47 gaps that
related to fatigue initiation in air and LWR environments, fatigue crack growth in LWR environments, and EAF
assessment methodologies. Given the large amount of work done on EAF by the nuclear industry since 2011,
the need to update and review the EAF Knowledge Gaps was identified. This report details the work done
regarding the previous Knowledge Gaps, highlights overlap and redundancy between the gaps, and revises
the gaps based on the information collected.

KEY FINDINGS
A detailed review of research since the original Knowledge Gaps were developed in 2011 was conducted and
found that significant additional laboratory data has been developed in LWR environments. The major focus
of this recent work was on austenitic stainless steels in a PWR primary coolant environment. Much of this
testing has been under conditions intended to more closely simulate plant operating conditions, in terms of
factors such as complex loading transients, thermo-mechanical loading, and surface finish. However, test
data on plant representative geometries and representative component sizes are still limited. The
observations from these advanced experimental studies provide a partial explanation for the observed
discrepancy between the predictions from Fen-based models and plant component field experience. Improved
predictive models have been developed that provide a better description of the observed behavior in these
recent laboratory tests. However, full validation for plant-representative component geometries and loading
histories is not available.
vii
13321451
EXECUTIVE SUMMARY

Based on a review of these data, the existing 47 Knowledge Gaps have been reviewed and revised, and a
shorter list of outstanding issues that warrant further study has been developed. This revised list is as follows,
with the relevant original Knowledge Gaps (KGs) listed in parentheses:
A. General issues
1. Further studies on both fatigue endurance and fatigue crack growth to support improved EAF
assessment models and address discrepancies between predictions from EAF models and plant
procedures (KGs 2 and 4)
2. Universal stakeholders’ EAF database with appropriate screening of data (KG 46)
B. Fatigue endurance data in air and fatigue design curves
1. Further work to address material variability effects and, where appropriate, develop material-
specific fatigue design curves (KG 9)
2. Revised transference factors on life and stress for design fatigue curves (KGs 3 and 12)
3. Data, improved understanding, and standardization of fatigue assessment methods for welds
(KGs 37 and 38)
C. LWR environment effects on fatigue design
1. Data deficiencies relating to defining Fen equations, especially temperature and strain rate
thresholds (KGs 1, 5, 10, 13, 20, 21, 34, and 47)
2. Improved EAF design methods that incorporate differences between transference factors
between LWR environments and air (KG19)
3. Improved assessment methods that recognize strain amplitude effects on Fen (KG 7)
4. Effects of mean stress on fatigue life and EAF (KGs 6, 11, 14, and 44)
D. Influence of complex and non-isothermal loading transients on fatigue life in LWR environments
1. Improved assessment methodologies under complex and non-isothermal loading more relevant
to plant components (KGs 8, 18, 39, 40, 41, 42, and 43)
2. Validation data under plant-relevant loading to support new assessment methods, including
applicability to alloys other than stainless steel (KGs 15, 16, and 17)
3. Improved total life methodologies to incorporate influence of strain gradients for thermal shock
transients (KG 16)
E. Fatigue crack growth in LWR environments
1. Improvements to reference crack growth curves for austenitic stainless steels, incorporating
threshold and R ratio effects (including negative R), and extension to BWR (normal water
chemistry and hydrogen water chemistry) conditions (KGs 22, 23, 24, 25, and 31)
2. Reference crack growth curves for carbon and low-alloy steels in BWR environments (KGs 29
and 30)
3. Improved reference crack growth curves for nickel-based alloys (KGs 26, 27, and 32)
4. Assessment methods for crack growth under plant-relevant loading that recognize weighting of
damage through the loading cycle (KGs 28 and 45)
F. Prototypical and component testing (KGs 33, 35, and 36)

viii
13321451
EXECUTIVE SUMMARY

WHY THIS MATTERS


This work provides a summary of the research done since the publication of the original Knowledge Gap
document. This report is intended to provide an independent statement of the state-of-the-art for EAF
research. It also highlights the links between the different research techniques and assessment methods to
facilitate better collaboration and progress on the topic of EAF.

HOW TO APPLY RESULTS


The work presented in this product could be used to help justify and focus research programs that are
important to the individual members. Furthermore, this document shows the links between experimental
research and development of assessment methods, which could enable EPRI members to better identify
collaboration opportunities to maximize the impact of their research.

LEARNING AND ENGAGEMENT OPPORTUNITIES


• International research organizations may find the report useful for possible future direction for
environmentally assisted fatigue research and testing.

EPRI CONTACTS: Kawaljit (Al) Ahluwalia, Technical Executive, kahluwal@epri.com


Jean Smith, Principal Technical Leader, jmsmith@epri.com
Gary Stevens, Technical Executive, gstevens@epri.com

PROGRAMS: Primary Systems Corrosion Research (PSCR), Program 41.01.01; Boiling Water Reactor
Vessel and Internals Project (BWRVIP), Program 41.01.03; Pressurized Water Reactor Materials Reliability
Program (MRP), Program 41.01.04

IMPLEMENTATION CATEGORY: Reference–Technical Basis

Together...Shaping the Future of Electricity®

Electric Power Research Institute


3420 Hillview Avenue, Palo Alto, California 94304-1338 • PO Box 10412, Palo Alto, California 94303-0813 USA
800.313.3774 • 650.855.2121 • askepri@epri.com • www.epri.com
© 2018 Electric Power Research Institute (EPRI), Inc. All rights reserved. Electric Power Research Institute, EPRI, and
TOGETHER...SHAPING THE FUTURE OF ELECTRICITY are registered service marks of the Electric Power Research Institute, Inc.

13321451
13321451
CONTENTS

ABSTRACT ..................................................................................................................................V

EXECUTIVE SUMMARY ............................................................................................................VII

1 INTRODUCTION .................................................................................................................... 1-1


1.1 Background .................................................................................................................. 1-1
1.2 Roadmap Hypotheses ................................................................................................. 1-2
1.2.1 Hypotheses ......................................................................................................... 1-2
1.2.2 Recommendations, Option 1 and 2..................................................................... 1-4
1.3 Objectives of the Update and Revision of the EAF Knowledge Gaps ......................... 1-4
1.4 Report Structure .......................................................................................................... 1-5

2 FATIGUE ENDURANCE DATA IN AIR ................................................................................. 2-1


2.1 Introduction .................................................................................................................. 2-1
2.2 Fatigue Data Generation and Mean Data Curves........................................................ 2-1
2.2.1 Knowledge Gap 9 (Previously Classified as Category C, Low Priority) .............. 2-1
2.2.2 Knowledge Gap 37 (Previously Classified as Category C, Low Priority) ............ 2-4
2.3 Transference Factors ................................................................................................... 2-6
2.3.1 Knowledge Gap 12 (Previously Classified as Category C, Medium Priority) ...... 2-7
2.3.2 Knowledge Gap 3 (Previously Classified as Category C, High Priority) ........... 2-11
2.4 Fatigue and Environmental Fatigue Databases ......................................................... 2-13
2.4.1 Knowledge Gap 46 (Previously Classified as Category C, High Priority) ......... 2-13
2.5 Summary of Fatigue Endurance Data and Design Curves in Air ............................... 2-14

3 LWR ENVIRONMENTAL FATIGUE ....................................................................................... 3-1


3.1 Introduction .................................................................................................................. 3-1
3.2 Conventional Cyclic Loading Fatigue Testing .............................................................. 3-2
3.2.1 Knowledge Gap 19 (Previously Classified as Category C, Medium Priority) ...... 3-2
3.2.2 Knowledge Gaps 10, 13 and 47 .......................................................................... 3-4

xi
13321451
3.2.3 Knowledge Gap 20 (Previously Classified as Category C, Medium Priority) .... 3-11
3.2.4 Knowledge Gap 7 (Previously Classified as Category C, High Priority) ........... 3-13
3.2.5 Knowledge Gap 34 (Previously Classified as Category A, High Priority) .......... 3-15
3.2.6 Knowledge Gap 21(Previously Classified as Category C, Low Priority) ........... 3-15
3.3 Complex Waveform Testing....................................................................................... 3-17
3.3.1 Knowledge Gap 17 (Previously Classified as Category C, High Priority) ......... 3-17
3.3.2 Knowledge Gap 18 (Previously Classified as Category B, High Priority) .......... 3-18
3.3.3 Knowledge Gap 11 (Previously Classified as Category A, Low Priority) .......... 3-23
3.3.4 Knowledge Gap 14 (Previously Classified as Category A, Medium Priority) .... 3-26
3.4 Non-Isothermal Loading ............................................................................................ 3-27
3.4.1 Knowledge Gaps 15 And 16 ............................................................................. 3-27
3.5 Prototypical and Component Testing ......................................................................... 3-32
3.5.1 Knowledge Gap 33 (Previously Classified as Category A, High Priority) .......... 3-32
3.5.2 Knowledge Gap 36 (Previously Classified as Category C, High Priority) ......... 3-36
3.6 Summary of LWR Environmental Fatigue.................................................................. 3-38

4 ENVIRONMENTAL FATIGUE CRACK GROWTH................................................................. 4-1


4.1 Introduction .................................................................................................................. 4-1
4.2 Conventional Cyclic Loading Fatigue Testing .............................................................. 4-2
4.2.1 Knowledge Gap 22 (Previously Classified as Category C, Medium Priority) ...... 4-2
4.2.2 Knowledge Gap 24 (Previously Classified as Category C, Medium Priority) ...... 4-4
4.2.3 Knowledge Gaps 25 and 31 ................................................................................ 4-5
4.2.4 Knowledge Gaps 29 and 30 .............................................................................. 4-10
4.2.5 Knowledge Gap 26 (Previously Classified as Category C, Medium Priority) .... 4-11
4.2.6 Knowledge Gap 27 (Previously Classified as Category C, Medium Priority) .... 4-12
4.2.7 Knowledge Gap 32 (Previously Classified as Category C, Low Priority) .......... 4-14
4.3 Complex Waveforms and Non-Isothermal Loading ................................................... 4-16
4.3.1 Knowledge Gap 28 (Previously Classified as Category A, High Priority) .......... 4-16
4.4 Summary of Crack Growth Data ................................................................................ 4-21

5 ASSESSMENT METHODS .................................................................................................... 5-1


5.1 Introduction .................................................................................................................. 5-1
5.1.1 Summary of Mechanistic Understanding of Fatigue and EAF ............................ 5-1
5.1.2 Review of Key Concepts ..................................................................................... 5-2
5.1.3 Fatigue Initiation Assessment ............................................................................. 5-3

xii
13321451
5.1.4 Fatigue Crack Growth Assessment..................................................................... 5-4
5.1.5 Summary of Current Fatigue Assessment Methods ........................................... 5-4
5.2 Fatigue Initiation .......................................................................................................... 5-9
5.2.1 Knowledge Gap 2 (Previously Not Classified) .................................................... 5-9
5.2.2 Knowledge Gap 4 (Previously Classified as Category B, High Priority) ............ 5-12
5.2.3 Knowledge Gap 1 (Previously Classified as Category C, High Priority) ........... 5-13
5.2.4 Knowledge Gap 5 (Previously Classified as Category C, High Priority) ........... 5-14
5.3 Fatigue Initiation: Complex Waveform Analysis......................................................... 5-15
5.3.1 Knowledge Gap 8 (Previously Classified as Category A, Medium Priority) ...... 5-15
5.3.2 Knowledge Gaps 39, 41, 42 and 43 .................................................................. 5-17
5.3.3 Knowledge Gap 6 and 44.................................................................................. 5-21
5.4 Fatigue Crack Growth ................................................................................................ 5-23
5.4.1 Knowledge Gap 23 (Previously Classified as Category C, High Priority) ......... 5-23
5.4.2 Knowledge Gap 28 (Previously Classified as Category A, High Priority) .......... 5-25
5.5 Welds ......................................................................................................................... 5-28
5.5.1 Knowledge Gap 38 (Previously Classified as Category C, Medium Priority) .... 5-28
5.6 Geometric Features ................................................................................................... 5-30
5.6.1 Knowledge Gap 40 (Previously Classified as Category C, Medium Priority) .... 5-30
5.7 Other Knowledge Gaps ............................................................................................. 5-31
5.7.1 Knowledge Gap 35 (Previously Classified as Category B, High Priority) .......... 5-31
5.7.2 Knowledge Gap 45 (Previously Classified as Category C, Medium Priority) .... 5-32
5.8 Summary ................................................................................................................... 5-34

6 SUMMARY OF OBSERVATIONS AND PROPOSED UPDATE TO KNOWLEDGE


GAPS ......................................................................................................................................... 6-1
6.1 Introduction .................................................................................................................. 6-1
6.2 Status of Original Knowledge Gaps ............................................................................. 6-2
6.2.1 General Issues .................................................................................................... 6-2
6.2.2 Fatigue Endurance Data in Air and Fatigue Design Curves ............................... 6-2
6.2.3 LWR Environment Effects on Fatigue Design ..................................................... 6-5
6.2.4 Influence of Complex and Non-Isothermal Loading Transients on Fatigue
Life in LWR Environments ................................................................................................ 6-8
6.2.5 Prototypical and Component Testing ................................................................ 6-10
6.2.6 Fatigue Crack Growth ....................................................................................... 6-10
6.2.7 EAF Assessment Methods ................................................................................ 6-13

13321451
7.1 Introduction .................................................................................................................. 7-1
7.2 General Issues ............................................................................................................. 7-2
7.2.1 Discrepancies Between Predictions from EAF Models and Plant Procedures ... 7-2
7.2.2 Universal Stakeholders’ Database ...................................................................... 7-3
7.3 Fatigue Endurance Data in Air and Fatigue Design Curves ........................................ 7-3
7.3.1 Material Variability and Material-Specific Fatigue Design Curves....................... 7-3
7.3.2 Transference Factors for Design Fatigue Curves ............................................... 7-4
7.3.3 Weld Metal Data and Assessment ...................................................................... 7-4
7.4 LWR Environment Effects on Fatigue Design ............................................................. 7-4
7.4.1 Data Deficiencies Defining Fen Equations, Parameters and Thresholds ............. 7-4
7.4.2 Applicability of Air Fatigue Design Transference Factors to LWR
Environments.................................................................................................................... 7-5
7.4.3 Strain Amplitude Effects on Fen ........................................................................... 7-6
7.4.4 Effects of Mean Stress on Fatigue Life and EAF ................................................ 7-7
7.5 Influence of Complex and Non-Isothermal Loading Transients on Fatigue Life in
LWR Environments ............................................................................................................... 7-7
7.5.1 EAF Endurance Data Under Complex and Non-Isothermal Loading .................. 7-7
7.5.2 Treatment of Cyclically Variable Parameters in Fen Equation ............................. 7-8
7.5.3 Influence of Strain Gradients for Thermal Shock Transients .............................. 7-9
7.6 Fatigue Crack Growth in LWR Environments .............................................................. 7-9
7.6.1 Reference Crack Growth Curves for Austenitic Stainless Steels ........................ 7-9
7.6.2 Reference Crack Growth Curves for Carbon and Low Alloy Steels .................... 7-9
7.6.3 Reference Crack Growth Curves for Nickel-Based Alloys ................................ 7-10
7.6.4 Crack Growth Under Plant-Relevant Loading ................................................... 7-10
7.7 Prototypical and Component Testing ......................................................................... 7-11

8 CONCLUSIONS ..................................................................................................................... 8-1

9 REFERENCES ....................................................................................................................... 9-1

A TRANSLATED TABLE OF CONTENTS .............................................................................. A-1


Français (French) ................................................................................................................. A-3

日本語 (Japanese) ............................................................................................................. A-19


Español (Spanish) .............................................................................................................. A-33

한국어 (Korean) ................................................................................................................. A-49

xiv
13321451
LIST OF FIGURES

Figure 5-1 SNW weighting function ......................................................................................... 5-20

xv
13321451
13321451
LIST OF TABLES

Table 3-1 Fen Parameters and Their Threshold Values for Ferritic Steels ................................. 3-7
Table 3-2 Parameters for EFEM Stainless Steel Fen Equation .................................................. 3-8
Table 3-3 Fen parameters and their threshold values for austenitic steels. ................................ 3-9
Table 3-4 Fen parameters and their threshold values for nickel-based alloys. ......................... 3-10
Table 6-1 Factors on Life Applied to the Mean Fatigue ε–N Air Curve to Account for the
Effects of Various Material, Loading, and Environmental Parameters. ............................ 6-15
Table 8-1 Listing of New Knowledge Gaps ................................................................................ 8-3

xvii
13321451
13321451
1
INTRODUCTION

1.1 Background
Established fatigue initiation life curves, such as those given in Section III of the American
Society of Mechanical Engineers (ASME) Boiler and Pressure Vessel Code [1], provide the
design basis for nuclear power plant (NPP) components. These curves do not explicitly account
for the effect on fatigue initiation life of the high-temperature water environments to which
components of the reactor coolant system in light water reactors (LWRs) are typically exposed.
Low-Cycle Fatigue (LCF) data, obtained by laboratory testing of small, polished specimens,
have demonstrated that substantial reductions in fatigue life may occur in LWR environments
compared with testing in air. The Nuclear Regulatory Commission (NRC), in Regulatory Guide
1.207 [2], has prescribed assessment guidance for fatigue initiation life of new NPP components
in the U.S. This guidance, the technical basis for which is described in the report
NUREG/CR-6909 Rev 1 [3], was developed based on research conducted by Argonne National
Laboratory (ANL). ANL’s work included a review of small specimen fatigue initiation life test
data and refined an environmental factor, Fen, approach originally proposed by EPRI [4], by
which the fatigue design lifetime is reduced in LWR environments compared to air. ASME has
also proposed Code Cases for assessing fatigue life in LWR environments, either using an
approach similar to that prescribed by the NRC, or using fatigue curves fitted to the experimental
data in water environments.
In addition to the environmental effect on fatigue life, crack growth test data also indicate that
significant environmental enhancement of fatigue crack growth rates can occur under LWR
environmental conditions. Although reference crack growth curves or Code Cases are available
in the ASME Section XI Code for some materials or reactor environments, this is not the case for
all relevant structural materials. ASME Code Case N-809 [5] was developed for austenitic
stainless steels in Pressurized Water Reactor (PWR) environments.
Application of Regulatory Guide 1.207, which is intended to account for environmental effects
on fatigue life, may result in prediction of high cumulative fatigue usage factors for some LWR
components, as compared with calculations based on extant design codes. These results may
represent a significant challenge in validating safe long-term NPP operation. However, these
calculated usage factors do not appear to be consistent with the field experience where there have
been no reported LCF fatigue failures of components in existing LWR plants. Most of the
failures which have occurred have been attributed to unanticipated transients or high cycle
thermal mixing, for example at tee junctions. As a result, a need was identified to resolve
uncertainties and identify Knowledge Gaps to improve the understanding and treatment of EAF
for lifetime justification of the adequacy of fatigue lives for LWR components. In response to
this need, a Knowledge Gap (KG) Analysis was performed and published in December 2011
which identified the specific technical areas where uncertainties existed at the time [6]. In that

1-1
13321451
Introduction

analysis, the status of existing research, test data and design code developments were reviewed to
identify KGs concerning EAF for ferritic steels, austenitic stainless steels and nickel-based alloys
in PWR and Boiling Water Reactor (BWR) water environments. A total of forty-five individual
KGs were identified as a result of the KG Analysis.
Subsequently, the KGs were reviewed by two nuclear industry-wide groups of experts, the EPRI
EAF Expert Panel, which met twice in 2011, and a separate Focus Group convened by EPRI in
2012, both of which assigned priorities to the various gaps. Two additional KGs were also
identified during the later review. A listing of the gaps ranked by priority (high, medium and
low) is provided in a later report, published in December 2012 [7]. This report also proposed a
number of hypotheses which might contribute to the apparent mismatch between the predictions
of the models based on laboratory data and plant experience. A Roadmap for future research was
also developed, together with suggested research activities which were considered to be needed
to resolve each KG.
Since the KG reports were published, there have been a number of significant relevant research
activities completed worldwide. These activities have been aimed at improving the
understanding of the differences between predictions of environmental fatigue behavior based on
laboratory data and plant component experience. A number of alternative methodologies for
EAF life assessment have also been proposed which are aimed at refining the NRC methods
described in NUREG/CR-6909 Rev 0 [8] and the related ASME Code Cases. A revised version
of NUREG/CR-6909 was issued for public comment in 2014, and the final revised document
(NUREG/CR-6909 Rev 1) [3] was subsequently published to address stakeholder feedback on
areas of concern identified in the original Fen models and the draft revised report.

1.2 Roadmap Hypotheses


The KG Analysis Roadmap document [7] sought to suggest a sequence by which the KGs could
be addressed. This document also presented further detail on the work packages needed to
resolve each KG. The intention of the Authors was to prioritize the KGs to maximize the short
term benefits to designers. Over a longer time period, the prioritized sequence of KGs was
intended to provide the mechanistic understanding to support new assessment procedures,
additional testing and fatigue management programs. To aid prioritization, the Roadmap
document formulated seven hypotheses which may contribute to the apparent discrepancies
observed between the predictions of fatigue assessment models developed from laboratory data
on small specimens and the in-service plant experience. The identified KGs were then assigned
to relevant hypotheses. Based on how relevant the hypotheses were to the industry, the
hypotheses were themselves categorized and prioritized. These hypotheses are reproduced in the
following subsections along with their respective prioritizations.

1.2.1 Hypotheses
Category A
The loading conditions used in the laboratory testing may be different than the loading
experienced in LWR components. For example, in almost all the environmental fatigue tests, the
testing was done at constant temperature with strain controlled load cycling. In reality, fatigue
cycling in power plant components is predominantly due to temperature transients.

1-2
13321451
Introduction

Hypothesis 1: Cyclically Variable Parameters in a Thermally Induced Stress Cycle Reduce or


Negate the Environmental Influence on Fatigue (Category A).
Priority Applicable Knowledge Gaps
High 15, 28 and 33
Medium 8, 42 and 43
Low N/A

Hypothesis 2: Compressive Stress Does Not Contribute to the Corrosion Fatigue Damage
Mechanism (Category A).
Priority Applicable Knowledge Gaps
High N/A
Medium 6 and 14
Low 11

Category B
There may be conservatisms in the current ASME Code fatigue analysis procedures that bound
any adverse effect due to environments.
Hypothesis 3: Conservatism Due to the Use of Bounding Transients for Design Purposes is
Sufficient to Accommodate Environmental Enhancement of Fatigue Damage (Category B).
Priority Applicable Knowledge Gaps
High 4 and 35
Medium N/A
Low N/A

Hypothesis 4: Conservatism in the Current Treatment of Non-Contiguous Cycles for Design


Purposes May Partly Account for Environmental Influences on Fatigue (Category B).
Priority Applicable Knowledge Gaps
High 18
Medium N/A
Low N/A

Category C
Conservatism is introduced through the use of inadequate material data and the calculation
methods derived from them which do not fully represent relevant plant conditions.
Hypothesis 5: Conservatism is Introduced in Plant Assessment Through the Use of Available
Material Data which is Insufficiently Comprehensive, in Terms of the Parameters Considered
and the Range of those Parameters, to Adequately Represent Realistic Plant Conditions
(Category C).
Priority Applicable Knowledge Gaps
High 3, 5, 13 and 46
Medium 12, 19, 20, 22, 24, 26, 27, 29 and 38
Low 9, 10, 21, 25, 32, 34 and 37

1-3
13321451
Introduction

Hypothesis 6: Conservatism is Introduced by the Calculation Methods Recommended for the


Determination of Fen Factors which are Largely Unsubstantiated and Do Not Adequately
Consider the Relevant Parameters and their Time-Dependent Influences (Category C).
Priority Applicable Knowledge Gaps
High 1, 7, 16, 17, 23, 36, 39, 41 and 47
Medium 40, 44 and 45
Low N/A

Category D
Inadequate mechanistic understanding leads to conservative assessment procedures or
inappropriate application of procedures to some components or plant conditions.
Hypothesis 7: Improved Mechanistic Understanding Would Identify Circumstances Where the
Application of the Fen Factor Approach is not Required (Category D).
Priority Applicable Knowledge Gaps
High 31
Medium 30
Low N/A

1.2.2 Recommendations, Option 1 and 2


Due to the large number of KGs covering a wide range of EAF topics that could involve
significant levels of cost and effort, the Roadmap document [7] presented two options for
resolving key deficiencies in EAF understanding. The first option essentially demands that all the
high-priority gaps are dealt with in the first instance, and the medium and lower priority KGs
updated only as necessary. Option 1 was not recommended by the Roadmap analysis due to
prohibitively high costs and long timescales.
Option 2 prioritizes the KGs by which hypothesis is the most significant. The significance of the
hypothesis is judged based on the perceived benefit from resolving it and applicability of the
hypothesis to plant components. On this basis, the Roadmap document suggests that Hypotheses
1, 3 and 6 are to be tested in parallel with each other. Hypotheses 5 and 7 were described as
important, but they would incur high costs and need long timescales to resolve. These hypotheses
were recommended to be tested over a longer timescale than Hypotheses 1, 3 and 6. Progress on
Hypothesis 4 was thought to require significant advance in the mechanistic understanding of
EAF and, therefore, it was not recommended until mechanistic understanding had significantly
improved. Hypothesis 2 was thought to be of limited applicability to plant components and was
not recommended.

1.3 Objectives of the Update and Revision of the EAF Knowledge Gaps
The aim of the current report is to review each of the KGs in the light of new data or information
which has become available since publication of the KG report in 2011. For each of the 47 KGs
in KG Analysis Roadmap report, the KG definition and research development need was restated
along with a table containing a reference to the relevant section in the original KG analysis

1-4
13321451
Introduction

document. Relevant new information is summarized for each KG and the status of each was
determined by considering the identified publically-available research. In this context, the gaps
are categorized as follows:
• Closed: In this case, it is judged either that sufficient information is now available and that
there is no longer a significant lack of knowledge in this area, or that further work to enhance
knowledge on this topic is not judged likely to yield significant benefit in the context of
improving assessment methodology for EAF.
• Open: The work performed to-date towards this KG is insufficient to provide the information
necessary to consider it resolved.
The second part of this report provides a simplified and revised list of KGs based on the
information obtained during the update. The revision and simplification exercise aims to remove
repetition of the KGs and to combine KGs that address closely related issues. This listing also
includes any new KGs that have emerged since the 2011/2012 reports were issued.

1.4 Report Structure


In the 2012 report, the KGs were categorized into several groups. The main groups were ‘S-N
data’, ‘crack growth data’, ‘reconciliation of test data with field experience’ and ‘fatigue
initiation life assessment methods’. Other topics covered ‘multiaxial loading’, ‘weldments’ and
‘inspection requirements’. Several KGs were grouped separately under ‘anomalous positions’
and ‘further application of existing knowledge’. A subsequent review of the KGs and these
groupings by a number of experts produced the following modified groupings: ‘fatigue curves
and initiation’, ‘environmental effects on S-N behavior’, ‘mean stress effects’, assessment
methods’, ‘crack growth’, ‘plant-relevant loading and prototypical testing’ and ‘inspection/real
plant conditions’.
The organization of the original KGs was revised for this report. The KGs themselves were
developed as a matter of course during the review of existing literature in the original KG
Analysis report [6]; therefore, the numbering sequence had no specific significance. The
Roadmap report [7] organized the KGs in two ways: by priority or by applicable hypothesis.
Furthermore, the priority set by the Roadmap report was based on the priorities allocated by two
expert groups, the EPRI EAF Expert Panel which met twice in 2011 to discuss the draft report,
and an EPRI focus group, which was specifically organized to discuss the KGs and met in San
Diego in 2012. Different priorities were assigned by these two groups; the two sets of priorities
are, for the most part equivalent, but in some cases they do differ.
It was judged that for the current report the best way of organizing the KGs was into the general
groups of: fatigue endurance data in air (Chapter 2); LWR environmental fatigue (Chapter 3);
environmental fatigue crack growth (Chapter 4); and assessment methods (Chapter 5). This
overall structure would allow for the majority of overlap and redundancy to be identified within
these specific topics. Following this, it was viewed that identifying the relationships and
consistent themes between the KGs would be far easier. It was felt that, by assessing the KGs in
this order and manner, the simplification and revision task would be better accomplished and
justified. Furthermore, this structure would make using this document, to track the KGs from

1-5
13321451
Introduction

their original definitions and progress to the newly revised ones, easier for the nuclear
communities. This has an additional advantage that a group interested in a particular type of
testing or assessment KG would be able to parse this document and access the current state of it
with relative ease.
Chapter 2, which looks at KGs related to fatigue endurance in air, first deals with data generation
and mean air curves, then transference factors used to derive fatigue design curves from mean
data curves, followed by the KG regarding a fatigue database. The next three chapters (3 to 5)
are structured in a similar way; firstly, they address any simple waveform-loading KGs, followed
by complex waveform testing, non-isothermal KGs, and plant and component testing KGs.
Chapter 6 summarizes the advances against the original KGs, identifies the redundancies and
overlaps found within the KGs, and highlights the relationships between the KGs. This sets the
framework for Chapter 7, which presents the revised KGs in a more streamlined and user-
friendly form that shows where collaboration between disciplines is required to close out the
identified KGs. These chapters also identify new emerging issues and their research needs.
Chapter 8 provides conclusions together with a tabulated list of the new KGs and identified
research required to address them.

1-6
13321451
2
FATIGUE ENDURANCE DATA IN AIR

2.1 Introduction
The KGs discussed in this section of the report are aimed at gathering fatigue life (S-N) data in
air for a wider range of loading conditions and materials than have been studied to-date, as well
as further understanding on the transference factors which are used to develop a design fatigue
curve from a mean air fatigue curve obtained from conventional cyclic fatigue tests in a
laboratory air environment. The transference factors are otherwise referred to as translation
coefficients or adjustment factors. As pointed out by Cooper [9], these factors are intended to
make smooth, small specimen data relevant to real pressure vessels and components and should
not be considered as safety factors.
The KGs in this section are grouped into two main categories: generation of new fatigue data
which may form the basis for new mean data curves and transference factors between mean data
and design curves. This grouping of the KGs follows the philosophy in NUREG/CR-6909 [3, 8]
that the design fatigue curve for an air or inert environment is developed first, and that an
environmental enhancement factor is then applied for components operating in light water
reactor environments. A number of studies suggest that this approach may be excessively
conservative because the transference factors used to develop the air design curve differ between
air and LWR conditions, or the laboratory specimen curves are overly conservative for use with
actual plant components. The influence of LWR environments on fatigue endurance is discussed
in Section 3.

2.2 Fatigue Data Generation and Mean Data Curves


The KGs contained within this section specifically relate to the lack of adequate S-N data for
some grades of material or availability of data under certain loading conditions (e.g., high cycle
fatigue). These data, if available in sufficient amounts, could be used to construct new mean air
curves, validate existing curves, or test the validity of or revise the factors used to translate the
mean air curves into design curves.

2.2.1 Knowledge Gap 9 (Previously Classified as Category C, Low Priority)


Knowledge Gap Definition
A generic stainless steel inert fatigue design curve for stainless steels may prove difficult to
apply since it appears to be too conservative in the high cycle fatigue regime for some materials.
Material-specific, high cycle design fatigue curves may be preferred in some cases.
Research and Development Need
If there is a need to refine the inert fatigue curve for specific material grades, additional high
cycle fatigue initiation life data will be required.

2-1
13321451
Fatigue Endurance Data In Air

Ancillary Information from 2011 KG Report


Chapter Section Subsection(s) Page(s) Applicable Materials
9 Critical Review of Design Revised Inert Fatigue 9-1 to 9-2 Stainless Steel
Code Developments Design Curves

Background
The fatigue design curves presented in NUREG/CR-6909 [3, 8] for ferritic steels differ from
those presented in ASME Section III. The fatigue curves in NUREG/CR-6909 and ASME
Section III for austenitic stainless steels are identical since the 2009 addenda, and they are based
on transference factors of 12 on life and 2 on stress. Although the new and old ASME stainless
steel curves are quite similar in the low-cycle fatigue region, the more recent stainless steel
curves result in significantly reduced fatigue lives compared to the earlier ASME Section III
curves in the high cycle fatigue regime (i.e., above about 105 cycles). This lowered region is
caused by the fatigue threshold for stainless steel expressing heat-to-heat sensitivity. The
implication of this is that, for some heats with a higher than average fatigue limit, the use of a
mean data curve covering many materials may introduce additional unnecessary conservatism.
Therefore, there could be considerable benefit to designers in producing materials-specific
design curves for stainless steels. This conservatism was also noted by the NRC [10].
Review of Activities Relevant to KG 9
Several fatigue endurance research programs [11, 12, 13] are taking steps towards producing
their own specific stainless steel curves relevant to materials used in their national nuclear
programs. However, the quantity of data required in both the low and high cycle fatigue regimes
to justify a material-specific curve makes it difficult to close this KG. Also, some of the observed
scatter in S-N data may not necessarily be a grade-to-grade variability, but may reflect other
factors such as material strength. Additionally, competition from other research areas, such as
defining the effect of complex waveforms or hold times, has meant that the focus of many of the
cited research programs has not been on generating heat-specific mean air curves.
Schuler et al. [14] demonstrated that the stabilized stainless steels used in German plants showed
substantially different fatigue behavior to the mean air data line presented in NUREG/CR-6909
[3] in the high cycle fatigue region. Other studies conducted in the UK and France [15, 16] have
also reported significant heat-to-heat variability, even for conventional grades of stainless steel
(304, 316 and their low carbon (L) variants). The greatest observed differences between stainless
grades are in the high cycle region, reflecting differences in the fatigue endurance limits at
ambient temperature. The cause of the variability between different material heats and grades is
therefore still uncertain.
Using data produced by multiple institutions in room and high-temperature air, Schuler was able
to show that the titanium and niobium stainless steels used in the studies had better performance
at ambient temperature in the 1x105-1x107 cycle range than predicted by the NUREG/CR-6909
and current ASME Section III air curves. This performance was attributed to the stabilized
stainless material having a higher fatigue endurance limit at ambient temperature than assumed
in ASME Section III. However, they also found that differences in fatigue lives between room
temperature and high-temperature air data were significant, and two best fit curves were required
to fit these datasets. Elevated temperatures, in the range of 240 °C to 350 °C, were observed to
have a detrimental effect on fatigue life of the stainless steel grades studied, in that fatigue lives

2-2
13321451
Fatigue Endurance Data In Air

in the 1x105-1x106 cycle range dropped below those predicted by the NUREG/CR-6909 mean air
curve. This is a significant observation, since NUREG/CR-6909 included data at temperatures up
to 400 °C in the development of the fatigue best fit air curve for stainless steel. The resulting two
mean data curves and their respective design curves have been adopted into the German KTA
3201.2 code [17].
The fatigue endurance data produced by these data generation test programs has typically been
assessed using either the Langer, Coffin Manson or Basquin type models. The UK [15], France
[18], Germany [14; 19], Japan [11] and the USA [3] favor the use of the Langer model which is
based on total strain/stress amplitude and takes the form of the: Langer equation, as follows:
𝒍𝒍𝒍𝒍(𝑵𝑵) = 𝑨𝑨 − 𝑩𝑩 ∙ 𝒍𝒍𝒍𝒍(𝜺𝜺𝒂𝒂 − 𝑪𝑪) Eq. 2-1

The parameters A and B are constants, N is the number of cycles to failure, εa is the strain
amplitude and the C parameter represents the fatigue limit of the material.
In contrast to those countries, some work performed in Hungary [20] has used a plastic strain
based Coffin-Manson approach. The work by Trampus et al. [20] argued that because fatigue
damage is caused by plastic strain, plastic strain based or strain energy methods were a sensible
alternative to other strain based models. Additionally, the current European standard EN 13445
advocates a Basquin model for welded components but a Langer model incorporating tensile
strength for unwelded components [21]. This represents a significant difference between the
European methodology and those used by ASME or the Japanese Society of Mechanical
Engineers (JSME).
Differences in how individual countries draw conclusions on their data, as well as how the data
are produced, make comparisons between the data and NUREG/CR 6909 difficult. The
INCEFA+ collaborative project will, as a matter of course, make significant progress towards
this KG by the participation of multiple European countries. Although assessment methods and
codes being developed by some of the different participating countries for regulatory purposes
may differ, the intentions within the INCEFA+ project are to perform fatigue tests in air and
PWR water on a common material and to develop a fatigue assessment procedure based on
analysis of the data produced by the project, together with data which may be made available to
the project from external organizations [12].
Summary
In light of the above discussion of the work done towards KG 9, there has not been enough
concerted effort to produce heat-specific design curves which can take advantage of material-
specific benefits. Ideally, to close this gap, additional stainless steel material-specific curves
would need to be adopted by safety standards organizations or regulators to demonstrate that the
case of stabilized stainless steel is not unique. KG 9, therefore, remains open.

2-3
13321451
Fatigue Endurance Data In Air

2.2.2 Knowledge Gap 37 (Previously Classified as Category C, Low Priority)


Knowledge Gap Definition
Data on weld metal are more limited than on parent materials but, where studied, appear to be
bounded by parent data.
Research and Development Need
More heat specific data on the behavior of weld metal may be required.
Ancillary Information from 2011 KG Report
Applicable
Chapter Section Subsection(s) Page(s)
Materials
9 Critical Review of Design Treatment of 9-12 to 9-13 All
Code Developments Weldments

Background
The method used by ASME Section III assumes that the fatigue lives of weldments are bounded
by parent plate material. The general trend is for weld material to have a higher strength than the
parent material and, consequently, fracture toughness is lower. In this sense, it could be inferred
that fatigue endurance properties of weld material should be superior to parent material.
However, a weld feature can have additional properties that reduce the fatigue lifetime of the
component. This includes specific geometrical features which are dealt with using a stress
concentration factor (see KG 38 in Section 5.5.1). The weld metal data analyzed in NUREG/CR
6909 Rev 1 [3] shows that, for ferritic and austenitic steels, fatigue lives were on average slightly
lower than the parent material but are still encompassed by the scatter band for the various
grades of steel. However, it is also stated that there are only limited data on the fatigue properties
of weld metals to compare to those of the parent material, especially for LWR environments. The
Knowledge Gap Analysis [6] and Road Map document [7] also highlight that the lack of S-N
data on the environmental fatigue life of weld metals represents a KG.
Review of Activities Relevant to KG 37
Similar Metal Welds
Fatigue testing in air on niobium stabilized stainless steel (X6CrNiNb 18-10) welds has been
carried out in Germany by Areva, TU Darmstadt and MPA. Langschwager et al. have focused on
the comparison of base metal, weld metal and butt welded joints in terms of the times for LCF
initiation [22; 23]. The LCF tests were performed at room and high-temperature air (200 °C and
350 °C). Langschwager’s work reported in 2014 [22] measured the N5 fatigue initiation lifetime
(i.e., number of cycles for a 5 % load drop) for cylindrical and flat specimens. They observed
that weld metal specimens (with no weld interface in the specimen gauge length) were harder
and had shorter fatigue lives than the corresponding base metal, which is similar behavior to that
observed in NUREG/CR 6909 Rev 1 [3]. However, the weld material was still well bounded by
the KTA design curve. This work also studied a flat fatigue specimen with a weld/parent metal
interface in the controlled gauge length. In this case, the primary crack occurred specifically in
the base metal due to the higher yield strength of the weld material giving rise to a local strain
concentration in the base material. The fatigue initiation life for this weld interface specimen was
found to be up to 10 times shorter than that of a base metal specimen due to this strain

2-4
13321451
Fatigue Endurance Data In Air

concentration. Further work involving a comparison of numerical solutions with optically


measured strain fields was planned to resolve the problems surrounding testing specimens with
weld interfaces in the gauge length.
A weld feature can have additional properties that reduce the fatigue lifetime of the component.
It can be expected, therefore, that specimens containing a weld interface within the gauge length
would produce shorter lifetimes than those without. The term ‘metallurgical notch’ has been used
to describe a local region in a component or specimen which is of different strength from the
bulk material [24]. In the UK R5 procedure [25], the fatigue strength reduction factor, therefore,
includes both geometrical and mechanical components, but the procedure does not give any
guidance on environmental effects.
Due to the experimental issues of having a metallurgical notch in the gauge length of fatigue
endurance specimens encountered in previous work, Langschwager et al. [23] used digital image
correlation to define the local and global strains in specimens containing weld interfaces. This
allowed the strain to adjust, depending on failure location, to a strain comparable to that
experienced by round bar specimens made from uniform material. For example, a welded
specimen was cycled at applied strain amplitudes of 0.36 % resulting in a lifetime 10 times
shorter than expected. However, by accounting for local strain in the base metal due to the weld,
the local strain amplitude was increased to 0.5 % which resulted in a fatigue life within the data
range between the weld and base metal. This work also presented high-temperature results on
welded specimens using the local strain correction method described above. These results
showed that, at 350 °C, weld interface specimens failed before comparable base metal or weld
metal specimens, but were contained within the KTA design curve for austenitic stainless steel.
In the work of Bosch et al. [24], metallurgical notches were considered using a finite element
local approach method. The conclusion was that the disparity between predictions and
experiment could be rationalized by using a geometric strain enhancement factor (see Section
5.5.1). A modification to the Ke (Sn) parameter was proposed to ensure that lifetime predictions
are lower than experimental values.
The work done on this KG supports the ASME approach of applying a stress concentration
factor to parent material properties when analyzing weldments (See the discussion regarding
KG8 in Section 5.5.1).
Dissimilar Metal Welds
Schuler [26; 27; 28; 29] also investigated the fatigue life behavior in air of this same stabilized
stainless steel in the form of dissimilar metal welds (DMWs) and cladding material from a
Reactor Pressure Vessel (RPV) and main coolant line. In this work, specimens were taken from
across the weld with the gauge length of the cylindrical specimens being positioned in various
locations, i.e., austenitic stainless or carbon steel base metal, buttering, or weld metal.
Additionally, specimens were also taken from the transitions between stainless steel base metal
and weld metal, weld metal and buttering, and buttering and carbon steel base metal. This work
showed broadly similar behavior to that observed by Langschwager [23]. LCF tests found that, in
general, weld metal and buttering gave initiation lives below the mean data curve (KTA) for
stabilized stainless steel in air and, in some cases, outside of the scatter band, whereas the data
for cladding material were found to fall within the mean data curve scatter band for stabilized
austenitic stainless steel. For the DMWs, the mean data curve could be used for specimens with
transitions from buttering to ferritic steel or buttering to weld metal. However, specimens

2-5
13321451
Fatigue Endurance Data In Air

containing a transition from austenitic base metal to weld metal gave initiation values well below
the mean data curve for austenitic stainless steels and produced consistent failures in the base
metal section of the test specimens. The strain concentration introduced in the base metal by the
presence of a higher strength weld material was cited as the cause for these observations.
EN13445—Weldments
The fatigue design curves shown in EN 13445 [21] deal specifically with welds and were not
explicitly described in the initial KG Analysis document [6]. EN 13445 contains a design curve
for unwelded material (applicable to both ferritic and austenitic steels) and a separate set of
design curves relating to weld metals. The fatigue limits of both welded and unwelded materials
for constant amplitude loading are described as a cut-off limit based on a stress amplitude below
which damage does not accumulate. For welded materials, this occurs at 5x106 and 2x106 cycles
for non-welded materials. In the case of variable amplitude loading that includes stress ranges
greater than that defined for the fatigue limits, separate equations are used to define the portions
of the curves below these cycle limits.
Summary
The experiments run by Schuler and Langschwager highlight the experimental problems with
testing across welds. Any abrupt change in material properties in the specimen gauge length
appears to have a significant effect on the life obtained for the specimen, which is dependent on
test specimen geometry and may not be representative of actual plant component behavior where
through-wall plastic strains would not normally be present. This phenomenon has been described
as a metallurgical notch and, without the input of Finite Element Analysis (FEA) or a method of
defining the local strains throughout the specimen, the data generated using this type of specimen
cannot be readily analyzed. If further S-N data are to be produced using specimens with a weld
line in the parallel gauge length, then the methods described above will need additional
investigation. However, the initial experimental data produced using these methods showed that,
although at high-temperatures the weld interface specimens failed slightly earlier that those of
base or weld metal, they were bounded by the relevant design curves.
The KG Analysis report stated that the mechanism of corrosion fatigue initiation is not fully
understood so it cannot be assumed that the existing methods for treating weldments are
adequate. In light of the above discussion and work done towards KGs 37 and 38, the amount of
S-N data available for determining if weld metal is in fact bounded by parent material are
insufficient. In addition, no relevant studies in LWR environments have been identified. KG 37,
therefore, remains open.

2.3 Transference Factors


The KGs contained within this section relate to the factors applied to a mean data curve to
convert it into a fatigue design curve. They focus heavily on the methodology detailed in ASME
Section III, of which a past version forms the basis of the Japanese standard.

2-6
13321451
Fatigue Endurance Data In Air

2.3.1 Knowledge Gap 12 (Previously Classified as Category C, Medium Priority)


Knowledge Gap Definition
Conservatism is included in NUREG/CR-6909 Rev 0 concerning the derivation of the
adjustment factor of 12, which is used to relate test endurance data in air to component
endurance data. Insufficient data exist concerning the values for the individual factors which are
combined and the means by which they should be combined.
Research & Development Need
Further test data are likely to be required to relate smooth specimen fatigue initiation life to
engineering component fatigue initiation life by accounting for material variability, size effects,
surface finish, and loading history.
Ancillary Information from 2011 KG Report
Chapter Section Subsection(s) Page(s) Applicable Materials
9 Critical Review of Review of 9-2 to 9-3 All
Design Code NUREG/CR-6909
Developments Fen Factor Approach

Background
ASME Section III describes a procedure for creating fatigue design curves by adjusting mean air
data curves by factors of 20 or 12 on life and 2 on stress/strain, whichever results in a greater
influence on fatigue life. KG 12 is concerned with the factor on life; the factor on stress is
covered by KG 3, discussed below.
For both ferritic and austenitic steels, ASME Section III originally set adjustment factors of 20
on cycles to transform the mean air fatigue data curves into the fatigue design curves. The same
value was recommended for ferritic steels, austenitic stainless steels and nickel-based alloys
(such as Alloy 600). The factor of 20 was made up of several transference subfactors (the values
of these subfactors are given in parentheses) which accounted for data scatter (2.0), size effects
(2.5) and a factor which included aspects such as surface finish and industrial environments (4.0)
[9; 30]. NUREG/CR-6909 [3; 8] applied statistical analysis to the available data which showed
that the factor of 20 was overly conservative and suggested that a value of 12 be used instead.
The value of 12 in NUREG/CR 6909 includes sub-factors covering material variability and data
scatter, surface finish, size and loading history. NUREG/CR 6909 Rev. 1 supported the retention
of a factor of 12 on life as the adjustment factor to be consistent with the adjustment factor used
by the ASME Code fatigue design curves, even though further analysis suggested a factor of
about 10 was probably more appropriate.
It is noted that ASME Section III currently uses a factor of 12 for the fatigue curve for austenitic
stainless steels and nickel-based alloys, but retains a factor of 20 for the carbon and low alloy
steel fatigue curve. The reasons for retaining the factor of 20 is unclear. In addition, ASME
Section III continues to use a single, combined fatigue curve for carbon and low alloy steels,

2-7
13321451
Fatigue Endurance Data In Air

whereas NUREG/CR-6909 defines separate fatigue curves for these two material groupings. Out
of the four parameters that NUREG/CR 6909 Rev. 1 identified (surface finish, size effect,
loading sequence and data scatter) the factor covering data scatter was statistically evaluated in
detail with resulting values ranging from 2.1 to 2.8 [3; 13].
The KG Analysis report highlighted several issues with the values that had been calculated for
the subfactors that make up the total adjustment factor on life [6]. Some examples of these are
that (i) the subfactor of 1.0 to 2.0 for loading history may not be appropriate to all plants, (ii) a
size effect factor has been included despite NUREG/CR-6909/ASME Section III stating that it is
not required for rough surfaces [8], and (iii) the surface finish factor may be dependent on
environment. The key barrier to setting reasonable transference factors was identified as the lack
of consistent data with clear conclusions regarding the transference factors themselves.
It should be noted that KG 12 relates specifically to the air design curve. The approach of
NUREG/CR 6909 to calculation of fatigue life in LWR environments is to use the air fatigue
design curve together with a multiplicative environmental factor, Fen. Clearly, if the transference
subfactors for any of the parameters used in NUREG/CR 6909 to develop the air fatigue design
curve (data scatter, surface finish, size and loading history) are different in LWR environments
than in air, then the environmental fatigue life will be incorrectly predicted. Thus, whilst the
current KG 12 relates specifically to the air fatigue curve, the discussion relating to the impact of
the environment on the surface finish factor is contained within KG 19. As discussed below and
in Section 3, KG 19 identified some evidence for such a difference in the case of the surface
finish transference factor and recommended further work to confirm this.
Review of Activities Relevant to KG 12
Data Scatter and Material Variability
The effect of material variability and data scatter was included in the factor of 12 on life in
NUREG/CR-6909 to ensure that the fatigue design curve adequately covered 95% of the
available test data with 95% confidence. This effect was assessed by using best-fit curves to data
obtained from individual heats of material. The resulting constants were then ordered and
estimates of the cumulative distributions produced. These distributions were then fit to log-
normal curves and the standard deviations were determined. A range of values between 2.1 and
2.8, representing the 95th percentile of the data, were proposed for this transference subfactor by
NUREG/CR-6909 to cover austenitic and ferritic steels [3]. The work done by EdF and Areva on
stainless steels in support of the French EAF assessment methodology corroborates the value
proposed by NUREG/CR-6909 for use in EAF assessments [18]. Asada et al. [19] also
conducted an analysis of data scatter and material variability in support of the new Japanese
fatigue design curves. Asada presented transference factors, corresponding to the 95%
confidence level, of 2.154 and 2.487 for carbon/low alloy and stainless steels, respectively. It
can, therefore, be said that the data scatter and material variability transference factor proposed
by NUREG/CR-6909 is supported by the general literature.
Surface Finish
The majority of fatigue endurance tests that are undertaken in air, forming the basis for the
current design curves, are based on the use of specimens with a highly polished surface.
However, most plant components have significantly rougher surface conditions, which are
recognized to have a deleterious effect on the fatigue endurance performance of steels and nickel

2-8
13321451
Fatigue Endurance Data In Air

based alloys tested in air [3]. To take account of this effect, NUREG/CR-6909 Rev 0 [8] imposed
a transference subfactor, of between 1.5 and 3.5. The surface finish subfactor in NUREG/CR-
6909 is based on air data, but it was recognized at the time of the 2011 KG Report that some data
available in the literature suggested that the effect of surface finish in LWR environments may
be smaller. This is the subject of KG 19 which is discussed in the next section.
Work focusing on defining the magnitude of the surface finish factor applicable to carbon/low
alloy and stainless steels in air has been, in general, inconclusive. Wilhelm et al. (20) and Asada
et al. [21] have, independently, reanalyzed available austenitic and ferritic steel data
investigating surface finish effects in air. Wilhelm found that no equation could adequately fit all
the international data and advised that a factor of 2.5 be used. Asada, by using multiple
screenings of the data, was able to fit the ferritic steel data to two equations - one for ultimate
tensile strengths of greater than 800 MPa and one for less than 800 MPa - which contained the
maximum surface roughness (Rmax). For Rmax values of 50 µm, these equations yield a value
for the surface finish factor of 1.37 for a material with a UTS of 861 MPa (such as ASTM A193
B7, a low alloy bolting/fastener material) and 1.32 for steels with UTS less than 800 MPa.
Furthermore, Asada states that the analysis showed that the impact of surface finish on austenitic
steels was negligible. However, Asada observed that several means of describing surface
roughness, which were not strictly comparable to each other, were used in the literature, which
increased the uncertainty over the exact effect of surface finish on fatigue life. This aspect of
defining the effect of surface finish in air and a LWR primary coolant environment on the fatigue
life of stainless steel is being addressed within the European collaborative research project
INCEFA+ by testing a common material with the specimens finished to a consistent manner
across all the involved institutes [12].
The values for the surface finish factor in NUREG/CR 6909 are based on an equation which
utilizes the root mean squared value of surface roughness (Rq) [3; 8] and supported by the data
presented in NUREG/CR-6815 [22]. However, the data on surface roughness presented in
NUREG/CR-6815 are very limited and, as suggested in the industry comments in NUREG/CR
6909 Rev 1 [3], require expansion to encompass a wider variety of roughness values and more
data points. Furthermore, NUREG/CR-6909 recognizes the paucity of surface finish data with
which to draw conclusions. It is also recognized that the value applicable to ferritic steels (factor
of 2 to 3) may be lower than for austenitic steels (factor of 3) [3; 22]. This observation is
inconsistent with the approach adopted for EN 13445 [23], which has a specific term for material
strength in the calculation, and the work by Asada discussed above [21]. In contrast to the
observation in NUREG/CR-6909 that surface finish has less of an effect for ferritic than
austenitic material, both EN 13445 and the work by Asada indicate that the surface finish factor
should increase with the strength of the material. The inconsistencies between the re-analyzed
data and the inclusion of material strength in the equations of some established work highlight
the need for more research into this area of the fatigue design curves.
Size Effects
The size effects parameter is intended to account for the expected difference in fatigue life
between a small test specimen and a full size component. One of the aspects covered by this
parameter is the likely increase in Rt (maximum roughness height) when dealing with a larger
specimen and its resulting impact on fatigue life. This aspect raises the idea that the size effect
and surface roughness factors are linked, whereas NUREG/CR 6909 [3 states that, when

2-9
13321451
Fatigue Endurance Data In Air

considering rough surfaces, it is not appropriate to consider the effects of component size.
Another issue covered by size effects is the different specimen types that generate fatigue
endurance data. NUREG/CR 6909 states that the effect of specimen size on fatigue life is not
significant when the fatigue curve it would be applied to is based on axial strain-controlled data.
Additionally, no size effects were observed for smooth specimens tested using plain bending or
rough specimens using axial loading. NUREG/CR-6909 further states that the effect of specimen
size is only applicable to rotating bending test specimens where the fatigue limit was observed to
decrease by 25% as the specimen size was increased from 2 mm to 16 mm [3]. The impact that
the “size effect” has on the fatigue limit of rotating bending test specimens is generally explained
as being due to the variation of stress gradient for different sizes and an increase in statistical
variation of material properties which define the fatigue strength [24]. Despite making the above
comments, the original NUREG/CR-6909 [8] report suggested a value ranging from 1.2 to 1.4
for this transference subfactor. However, it was later revised to 1.0 to 1.4 in the Rev. 1 document
[3] in recognition of data supporting the statement that the impact of size effects on fatigue life is
negligible.
The position taken by the Japanese Design Fatigue Curve (DFC) Subcommittee, when
investigating the magnitude of the size effect parameter for carbon/low alloy and stainless steels,
is that it is negligibly small [24]. When the Subcommittee considered size effects on rotary
bending and axial tension/compression tests, they found that a factor of less than 1.1 was enough
and, therefore, the incorporation of a size effect factor was deemed unnecessary. They also found
that in axial tension/compression tests, the fatigue lives tended to be longer for larger specimen
diameters.
Loading Sequence
The loading sequence adjustment subfactor relates to data generated from tests that used a
variable amplitude methodology; for example, a set of high stress/strain amplitude cycles
followed by lower stress/strain amplitude cycles or vice versa. The most damaging cases were
reported to be for high stress cycles followed by low stress cycles. In these cases,
microstructurally small cracks, generated under high stress/strain loading, were observed to grow
to engineering sizes even when subsequently cycled at amplitudes below the fatigue limit. Due to
the possibility of a component being subjected to variable loading, a transference subfactor in the
range of 1.2 to 2.0 was applied in NUREG/CR-6909 Rev 0 [8]. The lower value was revised
down to 1.0, in NUREG/CR-6909 Rev 1 to reflect data showing that in some cases the effect of
loading sequence was negligible [3].
Although considerable work has been put into plant-relevant and complex waveforms
(considered further in KG 18), the information that could relate to the loading sequence subfactor
has not been put into the context of the transference subfactor. Neither EdF nor the Japanese
DFC [11] have made claims against loading history in their updates to the current fatigue design
curve methodologies [21; 24], although Le Duff et al. have suggested that a loading history
benefit could be claimed for thermal transients [16; 25]. Alternative approaches to address
complex transient loading have been developed more recently and are discussed in Section 5.

2-10
13321451
Fatigue Endurance Data In Air

Derivation of the Overall Adjustment Factor on Life in NUREG/CR-6909


NUREG/CR-6909 Rev 1 [3] provided further clarification of the method behind the adoption of
an adjustment factor of 12 on life as opposed to the original value of 20 suggested in ASME
section III. Monte Carlo simulations using the values, provided in the same document, for the
subfactors for data scatter and material variability (2.1-2.8), surface finish (1.5-3.5), size effects
(1.0-1.4), and loading sequence (1.0-2.0) were performed for each. From the results of these
simulations the value of the adjustment factor value corresponding to the 95th percentile, as
opposed to a median value, was found to be about 10. However, to maintain consistency with
ASME Code practices, a value of 12 was retained in NUREG/CR-6909 Rev. 1, although it was
also stated that further conservatism could be removed by using the factor of 10 suggested by the
analysis. The use of the latter value for the adjustment factor on life is supported by a proposal to
use it in the French RCC-M code [26].
It should be noted that the Monte Carlo simulations used in NUREG/CR-6909 assumed the
independence of the individual adjustment factors that were used to obtain a value for the factor
on life. This was due to the amount of data defining these subfactors being insufficient to
determine correlations between them. The potential for some of the subfactors to be correlated
was recognized in NUREG/CR-6909 Rev 1 [3] and supports the need for this KG to remain
open.
Summary
The majority of new data that have been collected relative to this KG have investigated whether
transference subfactors may differ in LWR environments compared to air. The main focus has
been on the impact on the surface finish transference subfactor with a smaller amount of effort
applied to generating data that support the data scatter subfactor. These aspects are discussed in
Section 3. In the context of air fatigue, recent analyses suggest that the value of 12 for austenitic
stainless steel may be slightly pessimistic and it may be possible to justify a factor of 10 as stated
in the recently published NUREG/CR-6909 Rev 1 and recently proposed for incorporation into
the French RCC-M Code [26].
More significantly, the disparity between the values of 20 (ASME) and 12 (NUREG/CR-6909)
for ferritic steels warrants further consideration, as well as the adoption of separate fatigue
curves for carbon and low alloy steels in ASME Section III. When the current amount of data is
considered as a whole, it is clear that more data would be required to further underpin the
individual transference subfactors. KG 12, therefore, should remain open.

2.3.2 Knowledge Gap 3 (Previously Classified as Category C, High Priority)


Knowledge Gap Definition
No comment is given on how the factor of 2 on stress was derived or why it is retained for both
air and water environments. This issue requires resolution since the technical basis for design
codes should be clearly understood.
Research & Development Need
Further review of the data underlying this methodology is warranted.

2-11
13321451
Fatigue Endurance Data In Air

Ancillary Information
Chapter Section Subsection(s) Page(s) Applicable Materials
9 Critical Review of Review of 9-1 to 9-11 All
Design Code NUREG/CR-6909
Developments Fen Factor Approach

Background
As discussed above, ASME Section III describes a procedure for creating fatigue design curves
by adjusting mean air data curves by a factor of 2 on stress/strain when this would result in a
greater influence than a factor of 12 on life. Although the adjustment factor on life was described
and justified in NUREG/CR-6909 [8], the factor of 2 to be applied to stress/strain was retained
for consistency with ASME Section III. Presently, the basis for ASME’s original selection of this
factor of 2 is not documented or justified. Revision 1 of NUREG/CR-6909 [3] recommends that,
before revised life factors of other than 12 are considered, that it be done in concert with the
reconsideration of an appropriate factor for stress/strain.
Review of Activities Relevant to KG 3
NUREG/CR-6909 Rev 1 [3] presented a more detailed description of the basis of the factor of 2
on stress (strain) than was provided in the original report [8]. In the Rev 1 report, it was stated
that the factor of 2 had been adopted to be consistent with ASME Code practices. It was
explained that this adjustment factor accounted for: data scatter, size, surface finish, loading
sequence, the effects of temperature and strain rate on secondary hardening, and dynamic strain
aging. It was stated that these effects are not cumulative and that the adjustment factor of 2 was
suitable to account for the dominant effect on the high cycle fatigue region (stated to be a
combination of material and environmental conditions). Chopra highlighted that this adjustment
factor may be conservative for some materials and that its scope and magnitude could be refined
if more data were available.
The current Japanese standard is based on the same transference factors on stress (strain) adopted
by ASME Section III, with factors of 20 and 12, respectively being adopted for carbon/low alloy
and stainless steels, and a factor of 2 on stress or strain. The Japanese DFC Subcommittee was
set up in 2011 to develop new design fatigue curves. Asada [11] reported on the work of the
subcommittee and indicates that, rather than deriving an overall transference factor (as was done
in NUREG/CR-6909 by Monte Carlo analysis), the subcommittee has the intention to explicitly
treat each subfactor (e.g., mean stress, surface finish, etc.) in a manner similar to that adopted by
EN13445. The developed design curves are intended to apply up to 1x108 cycles, with the
fatigue limit and high cycle fatigue design curves beyond 1x108 cycles being defined separately.
The DFC Subcommittee’s current plans suggest that the design fatigue limit from KHKS-0220
[38] will be used for low alloy and stainless steels. This new method is based on a mean curve
using air data at temperatures below 200 °C, which is then adjusted using factors covering mean
stress, surface condition and data scatter to produce the design curve. Current proposals indicate
that, for stainless steels in LWR conditions, both the stress and cycle transference factors only
need to include the effects of data scatter [11], and that no adjustments for mean stress [39],
surface finish [34] or size [36] are required. This leads to a transference factor of 1.34 on stress
for the proposed stainless steel fatigue design curve. In contrast, for carbon and low alloy steels,

2-12
13321451
Fatigue Endurance Data In Air

the DFC Subcommittee has indicated that the stress adjustment factor requires contributions
from mean stress, surface finish and data scatter. The DFC’s proposal provides a means of
calculating the factors relating to mean stress and surface finish but applies fixed factors for data
scatter (1.22 and 1.34 for ferritic and austenitic stainless steels, respectively).
In contrast to the DFC Subcommittee’s approach, EdF is continuing with the methods described
in ASME Section III and NUREG/CR-6909, but they are revising some of the transference factor
values. The approach to the factor on stress advocated by EdF described the derivation of a
factor of 1.4 which was calculated from the statistical analysis of high cycle fatigue data
produced using a more restricted range of steel grades (304L, 316L) relevant to French plants
[13; 40]. EdF used four different statistical approaches to validate the use of this adjustment
factor consisting of prediction intervals, order statistics and Wilks percentiles, maximum
likelihood and quantile regression. The effect of the EdF proposal would be to raise the air
fatigue design line in the high cycle regime to a level similar to the earlier ASME Section III
design curve. A subsequent analysis also included Finnish data on stabilized stainless steel
grades (321 and 347) and observed that a factor of 1.4 on stress covered these materials also
[40]. This approach, like the DFC subcommittee’s approach to stainless steel, only considers the
contribution from data scatter on the stress adjustment factor.
Summary
The recently published NUREG/CR-6909 Rev 1 document highlights that the adjustment factor
of 2 on stress (strain) was kept at this value to maintain consistency with ASME code practices.
Chopra stated that this adjustment factor was not revised in NUREG/CR-6909 due to the paucity
of data and the need to account for dynamic strain aging, and the effects of temperature and
strain rate on secondary hardening. This emphasizes the need for further data to properly define
the scope and magnitude of the adjustment factor on stress/ strain. The work by the DFC
Subcommittee and EdF suggest the current adjustment factor on stress is potentially conservative
under fully reversed (R = -1) loading, but not necessarily in all circumstances. However, there
still remains a distinct lack of data directly assessing the contribution to the stress adjustment
factor from size, surface finish and loading sequence. In light of the above discussion and work
done towards KG 3, outstanding uncertainties still remain with regard to the stress adjustment
factor used by ASME Section III and retained in NUREG/CR-6909 and, therefore, KG 3 remains
open.

2.4 Fatigue and Environmental Fatigue Databases

2.4.1 Knowledge Gap 46 (Previously Classified as Category C, High Priority)


Knowledge Gap Definition: Development of a Stakeholders’ Testing Database
A universal stakeholders’ testing database should be established to provide a consistent basis for
developments around the world.
Research and Development Need
Some gaps identify the need for further analysis of existing data. Other gaps identify the need to
add to existing data, while others identify a need to make all data available to researchers around
the world. To support these needs, a comprehensive and up-to-date database of worldwide
fatigue and EAF data is required. In addition, expert data screening for relevancy and validity
2-13
13321451
Fatigue Endurance Data In Air

should be considered. This will identify specifically where material data gaps exist and provide a
basis for coherent international collaboration. The data base should be continually updated as
new data become available.
Ancillary Information
This gap did not appear in the original Gap Analysis report (1023012) [6] but was recommended
by the International EAF Focus Group meeting in San Diego in 2012 and added to the KG listing
in the Roadmap report (1026734) [7].
Background
KG 46 was raised by the EPRI EAF Focus Group at their meeting in 2012 because of the fact
that much of the data in air and light water reactor environments on which ANL based the
NUREG/CR-6909 [8] environmental fatigue methodology is not available to the international
community. This lack of availability means that independent verification of the analysis
underpinning NUREG/CR-6909 is currently not possible. Although some data have been
obtained through digitization and data sharing, the majority of data used to construct
NUREG/CR-6909 are still not available to other organizations for detailed analysis. The analysis
in NUREG/CR-6909 combines data from both solid bar and hollow (tubular) specimens. In Rev.
1, it is stated that, although the former showed a lower mean life than solid specimens, the
difference was not statistically significant. A similar statement has been made by
Asada et al. [41], but is contradicted by Twite et al. [42]. The lack of availability of a fatigue life
database is a concern to the international community as the cost impact of NUREG/CR-6909 to
new build and plant life extension has been significant. There is also a need to collate and
consider additional data, e.g., significant new data generated outside of the U.S.
Summary
Due to the lack of a complete data set forming the basis of the assessment curves in
NUREG/CR-6909 [3; 8] being available to the international community, KG 46 should remain
open. The development of such a database covering both air and LWR environments is
considered high-priority.

2.5 Summary of Fatigue Endurance Data and Design Curves in Air


The KGs relating to fatigue endurance air data discussed within this section are all relevant to
Hypothesis 5 (defined as a Category C hypothesis) by the Roadmap document [7] (see
introduction for further detail). The Roadmap document does not recommend pursuing this
hypothesis over other hypotheses (such as Hypotheses 1, 3, or 6) due to the high cost and time
scale predicted to resolve it. Additionally, the Roadmap suggested that work towards Hypotheses
1, 3 and 6 would be a more effective use of research effort compared to dealing with Hypothesis
5. This is not to suggest that work towards Hypothesis 5 should be ignored; rather, it should be
carefully considered on a case by case basis, and the risk/reward balance properly considered.
Hypothesis 5 highlights KGs 3 and 46 as having a high-priority. However, these gaps have not
received the greatest level of attention with respect to the other gaps contained within Hypothesis
5. For example, KG 9 (defined as a low priority gap) and KG 12 (medium priority gap) have
arguably received more attention and effort than KG 3 and 46. With respect to KG 3, the data
underpinning the adjustment factor of 2 on stress/strain (used in ASME Section III,

2-14
13321451
Fatigue Endurance Data In Air

NUREG/CR-6909 and JSME) are not readily available and the basis of the factor of 2 is not fully
described nor justified. Nevertheless, NUREG/CR-6909 Rev 1 describes what this factor is
supposed to account for and highlighted that the lack of data available to inform this adjustment
factor prevented further investigation. From this perspective, KG 3 ties in with KG 46 which
focuses on establishing a database of international data. Specific efforts to analyze more
restricted datasets relevant to French, German or Japanese plants have attempted to define their
own values for the adjustment factor on stress/strain using statistical analysis but very little data
are available to assess other contributors such as surface condition or loading sequence. As a
result of this work, alternative adjustment factors between 1.22 and 1.4 have been suggested,
depending on whether the material is austenitic or ferritic.
KG 46 was not identified in the original 2011 KG Analysis report, but was identified and
described in the 2012 Roadmap document. It described the creation of a “stakeholders’ testing
database” that could provide a consistent basis for developing new methodologies. Although
discussions are ongoing and plans are being presented and considered, as yet no such database
exists. Potential hurdles for the creation of this database are the cost of maintenance and the
terms on which the data will be accessed by potential users or contributors. Testing organizations
are reluctant to release the results of large investments in testing without remuneration or some
form of confidentiality. A need for validation of contributed data was also identified. Despite
difficulties so far encountered in setting up such a database, the benefits of such a database
would be substantial since one of the underlying issues slowing down work on EAF is the lack of
access to quality data with which to test new assessment methods and interrogate existing ones.
KG 12 deals with the adjustment factor on life that is applied to the mean air curve to generate a
fatigue design curve. The bases for the adjustment factor on life in both ASME Section III and
NUREG/CR-6909 are significantly more detailed and documented compared to the adjustment
factor on stress/strain. As a result of this, it is easier for the nuclear industry to determine the risk
vs. reward scenario for putting effort into reanalyzing and justifying a new life adjustment factor.
The better understanding of the origin of the adjustment factor on life allows analysts to target
specific areas that are thought to be overly conservative. Additionally, this work ties into work
done in LWR environments which seeks to determine if the transference subfactors that make up
the adjustment factor on life—for example, for surface finish which is the largest contributor to
this factor—are appropriate. This has meant that a significant proportion of the work done
towards KG 12 has focused on being able to describe the impact of surface finish on fatigue life
of ferritic and austenitic steels by means of equations involving material strength and surface
roughness. Some further work has investigated size effects and loading sequences, but the work
done along these lines is very limited.
The progress towards KG 9 relating to material-specific design curves appears to have been
obtained mainly passively as a result of work done in pursuit of other KGs (such as KG 19). This
has occurred due to the need for researchers to compare their results in LWR conditions to mean
air curves that properly describe their material, which are often claimed to have better properties
than the mean air curve described in ASME Section III. Some work has been done that directly
contributes to this Gap, but similar to the work performed towards this Gap as described above, it
has been generated to take advantage of the better material properties of the materials used in
specific plants.

2-15
13321451
Fatigue Endurance Data In Air

Perhaps somewhat surprisingly, considering the sentiments of the 2012 Roadmap report, KG 37
(a low priority Gap) has received significant effort from researchers in Germany. These groups
have put effort into the testing and analysis of weld metals and specimens containing weld
interfaces. The driving force behind such work is the complexity of performing adequate analysis
on welds and the level of demonstration required to give confidence in the answer. The work
done towards this gap has reinforced the idea that weld metal is bounded by the parent metal, and
it has progressed some of the way towards generating an adequate testing methodology for
investigating specimens with weld interfaces in the gauge length.
Significant advancement has been made towards some of the Gaps discussed within this section,
but only a relatively small amount of effort has been put towards the high-priority Gaps
described by Hypothesis 5 in the 2012 Roadmap document. Insufficient advancement has been
achieved towards the Gaps in this section to close any of them. As indicated by the Roadmap
document, research effort may be better spent elsewhere on other Gaps, but it should be noted
that this needs to be carefully considered depending on the needs of operators, designers and
regulators.

2-16
13321451
3
LWR ENVIRONMENTAL FATIGUE

3.1 Introduction
This section of the report relates to fatigue endurance data generation in LWR environments and
the implications for methods used to assess fatigue endurance life of components exposed to
high-temperature water environments in operating LWR plants. Considerable reductions in the
number of cycles to failure have been observed when comparing conventional fatigue endurance
(S-N) tests conducted in light water reactor environments to similar effects in air, especially
when the tests are performed at low strain rates. This has prompted a significant international
effort to improving the fatigue life predictions for components subjected to LWR plant
conditions. The method that has been widely adopted involves the use of a life reduction factor
(Fen) that accounts for environmental effects. The environmental correction factor is a function of
a number of parameters, including strain rate, temperature, the oxygen content of the water and
material (carbon/low alloy steel, stainless steel, or nickel-based alloys; for carbon/low alloy
steels, the sulfur content is also an influencing variable). The effect of the Fen factors is to amend
the lifetimes predicted by the S-N curve in air to account for the deleterious impact of LWR
environments.
The use of the Fen approach has been found to predict high cumulative usage factors (CUFs) for a
number of plant components in PWRs and BWRs subjected to design transient combinations. A
significant portion of recent research has therefore been aimed at understanding the reasons for
the disparity between plant experience and the predicted CUF values (see KGs 2 and 4 in
Section 5.2). In particular, there has been an increasing focus on testing under loading conditions
more relevant to plant operating conditions and transients. The discussion in this section is
divided into four main topic areas:
• Generation of test data in LWR environments using conventional triangular or sawtooth
loading and implications for fatigue life using the Fen approach or alternative methods.
• Isothermal EAF testing using complex waveforms more relevant to plant.
• Non-isothermal EAF testing including thermo-mechanical fatigue and thermal shock testing.
• Prototypical EAF testing and tests on component-like features.
The aim of most of these tests is to provide data to support alternative assessment methodologies.
Although some discussion of the implications of testing for assessment methods is provided in
this section, a more detailed overview of the development of assessment methods for EAF is
provided in Section 5.

3-1
13321451
LWR Environmental Fatigue

3.2 Conventional Cyclic Loading Fatigue Testing

3.2.1 Knowledge Gap 19 (Previously Classified as Category C, Medium Priority)


Knowledge Gap Definition
Further S-N tests are warranted to confirm the apparently differing influence of surface
roughness between air and water environments. This may justify a reduction in the design
margin applicable for components in water environments.
Research & Development Need
Further S-N tests are warranted to consider the influence of surface roughness between air and
water environments.
Ancillary Information
Chapter Section Subsection(s) Page(s) Applicable Materials
9 Critical Review of Design Review of NUREG/CR-6909 9-3 All
Code Developments Fen Factor Approach

Background
The KG Analysis report [6] observed that limited data, predominantly on stainless steel,
suggested that the effect of surface finish on the fatigue life of metals in LWR environments
could be less than that observed in air. This could lead to excessive conservatism when
combining Fen and the surface finish transference factor [6; 43]. This observation was based on
two sets of observations: firstly, limited ANL data on ferritic steels that show a factor of 3 life
reduction for air environments [35], but, a negligible effect in oxygenated BWR conditions [8],
and, secondly, Areva data in a PWR environment which indicated that the surface finish
detriment for stainless steel was about half that in air [16]. The results are consistent with the
observation that shear notches, which are more evident on the surface of fatigue specimens tested
in air and effectively contribute to surface roughness, were less prevalent on specimens subjected
to a PWR environment. However, shear notches have also been reported in stainless steel
exposed to high-DO environments relevant to BWR normal water chemistry [8].
Review of Activities Relevant to KG 19
More recent data obtained on austenitic stainless steel from the French [44; 43; 45] and British
[15] national programs support the previous observation that the surface finish transference
factor in PWR water is significantly less than the value used in air. The earlier work by le Duff et
al. [16] that indicated a significantly smaller influence of surface finish in PWR environment
than in air has more recently been extended [43; 45]. The work done by Poulain et al. [45] found
the reduction factor (polished/ground) for 300 °C air is 1.7 and the reduction factor in PWR
water was 1.6. These results have been taken into account in the developing Fen integrated
assessment methodology that has been accepted as a Règle en Phase Probatoire (RPP)
[Probationary Phase Rule] in the French RCC-M code, and has recently been proposed as an
ASME Code Case (see Section 5). Stairmand et al. [15] have observed similar effects in that the
surface finish factor is a factor of 4 in air and 2 in PWR water. However, the two studies
(Poulain et al. and Stairmand et al.) differ with regard to the nature of the rough surfaces used
and the absolute values of the factors in both air and water. Asada et al. [34], after analyzing the

3-2
13321451
LWR Environmental Fatigue

stainless steel data available in the literature, concluded that a surface finish adjustment factor for
PWR environments was negligibly small. Asada’s work further analyzed literature fatigue
endurance data produced on ferritic steels. In this reanalysis, Asada found that the effect of
surface finish on fatigue life of carbon and low alloy steels was best described by an equation
involving tensile strength and maximum surface roughness. Using the equation described by
Asada [34] values between 1 and ~1.65 can be obtained for ferritic steels.
Although there is some disagreement over the precise value of the surface finish adjustment
factor between the different investigators, the majority of available information suggests that the
value currently used by ASME Section 3 and NUREG/CR-6909 for austenitic stainless steel
could be reduced for application to PWR environments using the NUREG/CR-6909 Fen
methodology by a factor of up to two. The reduced effect of surface finish in high-temperature
deaerated water environments compared to air is consistent with mechanistic understanding of
the crack nucleation process. In air, nucleation involves the formation of persistent slip bands
leading to short surface shear notches that produce a surface roughening effect, and is central to
the nucleation of fatigue cracks. This process is, to some extent, bypassed by increased surface
roughness. The shear notch features are less obvious in a PWR water environment [46], and
there is far less evidence of initial propagation along persistent slip bands. This provides a
mechanistic basis for the lessened effect of initial surface roughness in deaerated LWR water
than in air. However, the observation of shear notch initiation in oxygenated water [8] suggests
that further testing is required to determine whether or not the observed reduction in surface
finish transference factor between air and PWR coolant is applicable to BWR NWC conditions.
Since the size effect in NUREG/CR-6909 is due to the higher probability of finding crack
initiation sites on a component with a large volume rather than a small size laboratory specimen,
it has been postulated that the effect will be reduced for a rough surface [31]. In recognition of
this, EdF, in their RPP submission, have combined the size and surface roughness transference
factors into a single factor called component effect [44; 18]. Using this methodology, EdF has
specified a range between 1.5 and 2.5 for the component effect parameter in LWR environments.
When compared to the ranges for surface roughness (1.5-3.5) and size (1.0-1.4) specified in
NUREG/CR-6909 Rev 1 [3], EdF’s new range represents a reduced degree of conservatism and
therefore potentially significant benefit to EAF assessments in LWR environments.
Summary
Several studies have indicated that the influence of surface condition on fatigue life is
considerably greater in air than in high temperature water. Since surface condition is one of the
contributors to the transference factor on life, the differences between its effects in these two
environments could be responsible for conservatism when the air design curve is combined with
the Fen factor for environment. Additionally, some other components of the air transference
factor may be inapplicable in air such as the effect of an “industrial environment” which is
included in the so-called surface finish and environment factor, but is also accounted for
separately using the Fen approach for LWR environments. The implication of these observations
is that the air design curve effectively includes some aspects of the environmental effect. An
“allowable Fen” could therefore be defined which could be used to reduce the value of Fen as

3-3
13321451
LWR Environmental Fatigue

calculated by NUREG/CR-6909. Such an approach was proposed by Le Duff et al. [47] and
Métais et al. [48], and subsequently developed into the Fen-integrated approach discussed in
Section 5.2.1 as well as the Fen threshold methodology currently under consideration by ASME
[49].
The benefit of recognizing that some of the environmental effects on fatigue life are already
included in the ASME design curve so that the value of Fen can be reduced in fatigue usage
factor evaluations is obvious but requires further study. Most recent studies are on stainless steel
in PWR environments and relevance to other alloys and to BWR environments may need to be
addressed. It is therefore recommended that KG19 should remain open.

3.2.2 Knowledge Gaps 10, 13 and 47


Knowledge Gap Definition
Knowledge Gap 10 (Previously Classified as Category C, Low Priority)
There is a lack of data concerning the definition of material-specific thresholds for the
occurrence of EAF. This is likely to contribute importantly to the identification of components
for which consideration of environmental effects is not required. This is particularly so for
stainless steel.
Knowledge Gap 13 (Previously Classified as Category C, High Priority)
Comprehensive test data to define the environmental enhancement factor and encompassing the
full range of relevant parameters as independent variables are not available.
Knowledge Gap 47 (Previously Not Classified)
The need exists to provide guidance on circumstances where the approach of NUREG/CR-6909
is not appropriate because of DO levels.
Research & Development Need
Knowledge Gap 10
High cycle fatigue initiation threshold data may be required for different grades of stainless steel.
The possibility of heat-to-heat variability may also need to be considered. The thresholds may
relate most significantly to the strain range below which consideration of environmental effects
is not necessary. Temperature thresholds should also be considered.
Knowledge Gap 13
Further test data are required to cover a wider range of the independent variables of temperature
and strain rate, which are relevant to the Fen factor definition.
Knowledge Gap 47
Some BWR plants operate a strict control of DO, minimizing concentrations using hydrogen
water chemistry (HWC) and noble metal chemical additions (NMCA). The discontinuous
influence of DO in the NUREG/CR-6909 Fen algorithms for carbon and low alloy steels is
insufficiently refined to take advantage of closely controlled DO levels in reducing Fen values.
Fen factors defined as continuous functions of DO without thresholds are required. For austenitic

3-4
13321451
LWR Environmental Fatigue

stainless steels, the environmental effect is reduced in high oxygen conditions. The inclusion of a
DO dependent expression for O* (transformed dissolved oxygen parameter) in this case would
be of benefit to those BWRs operating normal water chemistry (NWC).
Ancillary Information
Knowledge Applicable
Chapter Section Subsection(s) Page(s)
Gap Materials
10 and 13 Review of
Critical Review
NUREG/CR-
9 of Design Code 9-1 to 9-3 All
6909 Fen Factor
Developments
Approach
47 This gap did not appear in the original Gap Analysis report
(1023012) [6] but was recommended by the International EAF
Stainless Steel
Focus Group meeting in San Diego in 2012 and added to the
Knowledge Gap listing in the Roadmap report (1026734) [7].

Background
The KG Analysis report [6] identified a lack of data supporting the Fen parameters and their
thresholds as defined in NUREG/CR-6909 Rev 0 [8]. NUREG/CR-6909 Rev 1 [3] was published
after the original KG Analysis document, and used updated fatigue databases to redefine the
equations for calculating values of Fen for ferritic steels, austenitic steels and nickel-based alloys.
The fit to the new statistical model presented in Rev 1 has resulted in the Fen parameters and their
thresholds being altered. The definition of these parameters, and their thresholds, represents a
significant part of accounting for the effect of EAF on typical plant materials. For example,
NUREG/CR 6909 (Rev 0 and Rev 1) both incorporate a strain amplitude threshold for
carbon/low alloy steel (0.07 %) and austenitic stainless steels (0.1 %) below which LWR coolant
environments do not have an impact on fatigue life. Therefore, some components made from
materials with a high fatigue limit would be subject to EAF assessment unnecessarily [6]. This
was suggested to be especially true for austenitic stainless steels.
With respect to the definition of the parameters themselves, NUREG/CR-6909 [3; 8] assumed
that the environmental factors which impact upon the fatigue life of carbon/low alloy or stainless
steels are compounded from their individual effects. The KG Analysis [6] highlighted the
absence of data across the full bounding range of the transformed factors. Additionally, the
dependency of these factors on each other was considered to require further consideration. It was
concluded that, without additional data, the method proposed by NUREG/CR 6909 to calculate
Fen by multiplying the environmental factors could not be fully verified.
As previously stated, NUREG/CR-6909 Rev 1 [3] contains an updated data set and revised
thresholds for the Fen parameters (T*, ε̇*, S* and O*) for austenitic and ferritic steels. KGs 10
and 13 are concerned with alterations to the definition and thresholds of the Fen parameters
respectively. Therefore, this section will only state where changes to NUREG/CR 6909 Rev 0
have been made or where relevant work published in the public domain has suggested
alterations.

3-5
13321451
LWR Environmental Fatigue

Review and Summary of Activities Relevant to KGs 10 and 13


Ferritic Steels
NUREG/CR-6909 Rev 1 [3] introduced several changes to the Fen expression for ferritic steels.
Firstly, a unified Fen equation for carbon and low alloy steels was produced, secondly, several
changes to the magnitudes and/or the way in which the parameters are calculated were made and,
finally, alterations were made to the threshold values of the Fen parameters that are shown in
Table 3-1. These alterations were made based on the better fit they produced to the increased
data set available to ANL for Rev 1.
The lower bound of the sulfur Fen parameter was removed in NUREG/CR-6909 Rev 1 due to the
observation that the fatigue lives of some heats with sulfur contents less than 0.005 wt.% were
still affected. Additionally, the thresholds based on Dissolved Oxygen (DO) were also removed
for the transformed sulfur parameter, but, the reasoning behind this is unclear. The calculation of
the sulfur parameter itself has also changed from what is described in NUREG/CR-6909 Rev 0
(shown in Table 3-1) to a linear equation based on sulfur content for weight percentages of sulfur
less than the threshold value (0.015 wt.%). This parameter changes to a fixed value of 3.47 when
the threshold value is exceeded. Furthermore, NUREG/CR-6909 Rev 1 highlighted the lack of
adequate data describing the impact of sulfur content on fatigue lives sufficiently well enough to
define a functional form beyond the assumed linear relationship and upper or lower bound for
this parameter.
The upper limit of the temperature Fen parameter was revised down from 350 °C to 325 °C in
NUREG/CR-6909 Rev 1 due to the lack of data above 325 °C and to provide a reasonable bound
to cover LWR operating conditions. The calculation of the temperature Fen parameter has
changed to reflect the analysis of the extended data set, and can take a maximum value of 1.316
at 325 °C. NUREG/CR 6909 Rev 1 states that for ferritic steels there is insignificant data at
temperatures above 290 °C performed in LWR environments and that the limit for the
temperature parameter was extended to 325 °C to cover reasonable operating conditions. The
lack of high temperature data covering reasonable plant operating conditions represents a
significant Knowledge Gap in itself, but given that in some instances pressurizer temperatures
can be ~345 °C this means that the proposed upper limit of 325 °C may not be adequate to cover
all areas of the plant. Therefore, more data are required to adequately define where the effect of
temperature saturates and, hence, the upper limit of the temperature Fen parameter for ferritic
steels.
The additional data included in NUREG/CR-6909 Rev 1 showed that there was still an impact of
environment on the fatigue life of ferritic steels below 0.001 % s-1. Therefore, the saturation
limit applied to the strain rate Fen parameter was decreased from 0.001 % s-1 to 0.0004 % s-1.
The upper limit for the effect of strain rate was also increased to 2.2 % s-1 on the basis of the
new data. These alterations also resulted in a change to the calculation of the parameter within
the bounding limits to normalize the value to the upper limit after which the impact of strain rate
on fatigue life was thought to be negligible. Furthermore, dynamic strain aging was also
discussed for LWR environments in Rev 1, and the shorter-than-predicted fatigue lives for 0.001
% s-1 tests in high-temperature water were attributed to this phenomenon. However, the
reduction in fatigue life vs. the prediction observed was stated to be less than a factor of two and,
therefore, could be accounted for by bounding data scatter rather than developing a specific
factor to deal with dynamic strain aging.

3-6
13321451
LWR Environmental Fatigue

Table 3-1
Fen Parameters and Their Threshold Values for Ferritic Steels
Fen
Parameter Rev 0 Rev 1
Limit
Sulfur S*=0.015 (DO>1.0 ppm) S*=2.0+98·S (S≤0.015 wt.%)
S =0.001 (DO≤1.0 ppm and S≤0.001 wt.%)
*
S*=3.47 (S>0.015 wt.%)
S*=S (DO≤1.0 ppm and 0.001<S≤0.015
wt.%)
S =0.015 (DO ≤1.0 ppm and S>0.015 wt.%)
*

Temperature T*=0 (T≤150 °C) T*=0.395 (T<150 °C)


T =T-150 (150 °C<T≤350 °C)
*
T =(T-75)/190 (150 °C≤T≤325 °C)
*

Oxygen O*=0 (DO≤0.04 ppm) O*=1.49 (DO≤0.04 ppm)


O =ln(DO/0.04) (0.04 ppm<DO≤0.5 ppm)
*
O =ln(DO/0.009) (0.04 ppm≤DO≤0.5
*

O*=4.02 (DO>0.5 ppm) ppm)


O*=4.02 (DO>0.5 ppm)
Strain Rate ε̇*=0 (ε̇>1 % s-1) ε̇*=0 (ε̇>2.2 % s-1)
ε̇*=ln(ε̇) (0.001 % s-1≤ ε̇<1 % s-1) ε̇*=ln(ε/2.2)
ε̇*=ln(0.001) (ε̇<0.001 % s-1) (0.0004 % s-1≤ ε̇<2.2 % s-1)
ε̇ =ln(0.0004/2.2) (ε̇<0.0004 % s-1)
*

Stainless Steels
The original Fen equation in NUREG/CR-6909 Rev 0 [8] could not result in a value of 1 even
when the threshold criteria for EAF were not satisfied. This was addressed in NUREG/CR 6909
Rev 1 [3] by removing a constant in the equation, this change is shown in Equation 3-1 and
Equation 3-2. The removal of the 0.734 constant allows the Fen equation to result in values of
unity in cases where EAF is considered to be negligible, for example where temperatures are less
than 100 °C.
′ ′ 𝑶𝑶′
𝑭𝑭𝒆𝒆𝒆𝒆 = 𝒆𝒆𝟎𝟎.𝟕𝟕𝟕𝟕𝟕𝟕−𝑻𝑻 𝜺𝜺̇ (Rev. 0) Eq. 3-1
−𝑻𝑻∗ 𝜺𝜺̇ ∗ 𝑶𝑶∗
𝑭𝑭𝒆𝒆𝒆𝒆 = 𝒆𝒆 (Rev. 1) Eq. 3-2

The lower bound of the stainless steel temperature parameter was changed from 150 °C to
100 °C based on the revised assessment carried out using the expanded database in NUREG/CR
6909 Rev 1 [3]. The upper threshold of the temperature parameter for stainless steel was also
reduced from 350 °C to 325 °C. To reflect the changes in the upper and lower bounds of the
temperature Fen parameter, the equation determining the value of this parameter was also
changed. Although this was done to better describe the increased data available, it should be
noted that the equation for the transformed temperature predicts a value of 0.9 at temperatures of
325 °C. This implies that at 325 °C, the effect of temperature on the value of Fen has not
saturated and for components, such as pressurizers which can operate at up to 345 °C, the value
of Fen may not be conservative. This is in contrast to NUREG/CR-6909 Rev 0 which states that
the transformed temperature parameter reaches unity at temperatures in excess of 325 °C.
The original Fen parameter, in NUREG/CR-6909 Rev 0, characterizing the effect of dissolved
oxygen on the fatigue lives of stainless steels was assigned a constant value for all dissolved
oxygen levels. However, Rev 1 has added upper and lower limits for this parameter based on the

3-7
13321451
LWR Environmental Fatigue

amount of dissolved oxygen and if the material is considered to be sensitized. The new limits
mean that the impact of dissolved oxygen on fatigue life is a maximum for low dissolved oxygen
levels or a combination of high dissolved oxygen and either sensitized high carbon wrought
stainless steel or cast material. This revision is based on the observation that the fatigue lives of
stainless steels were either equivalent or longer for dissolved oxygen levels greater than 0.1 ppm
unless the material was sensitized or cast [3].
There is a notable difference between the calculated Fen value for stainless steels in BWR
environments determined using the Japanese Environmental Fatigue Evaluation Method (EFEM)
[50] or NUREG/CR 6909 Rev 1 [3]. NUREG/CR-6909 Rev 1 suggested that the cause of the
discrepancy could be the differing ways in which the DO and temperature dependencies were
approached for the two models. The EFEM claims that the data used to form the basis of this
method were insufficient to demonstrate an effect of DO on the fatigue lives of stainless steel
specimens in either PWR or BWR water chemistries. Furthermore, the EFEM approach suggests
that there are no temperature thresholds for stainless steels in BWR conditions and no minimum
threshold, whereas, in PWR conditions, an equation is provided for temperatures ≤325 °C and a
constant value for temperatures above this. This indicates a lack of consensus in the international
community regarding the calculation of Fen values for stainless steels and requires further
investigation. The EFEM equation and parameters for stainless steel are provided below,
Equation 3-3 and Table 3-2, for information; where Fen is the environmental factor, C is a
constant (0.992 for BWR and 3.910 for PWR), ε̇* is the transformed strain rate and T* is the
transformed temperature:
∗ )×𝑻𝑻∗
𝑭𝑭𝒆𝒆𝒆𝒆 = 𝒆𝒆(𝑪𝑪−𝜺𝜺̇ Eq. 3-3

Table 3-2
Parameters for EFEM Stainless Steel Fen Equation

BWR Equation PWR Equation


ε̇*=ln(2.69) where ε̇ >2.69 % s-1 ε̇*=ln(49.9) where ε̇ >49.9 % s-1
ε̇*=ln(ε̇) where 0.0004 % s-1 ≤ε̇≤ 49.9 % s-1
ε̇*=ln(ε̇) where 0.00004 % s-1 ≤ε̇≤ 2.69 % s-1
For stainless steel except cast stainless steel.
ε̇*=ln(ε̇) where 0.00004 % s-1 ≤ε̇≤ 49.9 % s-1
ε̇*=ln(0.00004̇) where 0.00004 % s-1 <ε̇
For cast stainless steel
ε̇*=ln(0.0004) where ε̇ <0.0004 % s-1
T*=0.000969xT
For stainless steel except cast stainless steel.
ε̇*=ln(0.00004) where ε̇ <0.00004 % s-1
For cast stainless steel.
T*=0.000782xT where (T≤325 °C)
T*=0.254xT where (T>325 °C)

The upper limit applicable to strain rate was altered between the two versions of NUREG/
CR 6909 from 0.4 % s-1 to its current value of 7 % s-1. NUREG/CR-6909 Rev 1 stated that this
change was based on the inclusion of the additional data in the database. However, the plots
presented in NUREG/CR-6909 Rev 1 show data up to only 0.4 % s-1 or 0.8 % s-1 for low DO
and high-DO environments respectively, but no data are shown for higher strain rate values to
support the change in limit from 0.4 % s-1 up to 7 % s-1. The alteration to the upper bound of the

3-8
13321451
LWR Environmental Fatigue

Fen parameter resulted in an adjustment to the normalization applied to the calculation of this
parameter in between the upper and lower bounds.
Table 3-3
Fen parameters and their threshold values for austenitic steels. Note an editorial change
between Rev 0 and Rev 1 labels the transformed parameters with an asterisk for all
materials.
Fen Parameter
Rev 0 Rev 1
Limit
Temperature T’=0 (T≤150 °C)
T*=0 (T<100 °C)
T =(T-150)/175 (150 °C<T≤325 °C)

T =(T-100)/250 (100 °C≤T≤325 °C)
*
T’=1 (T≥325 °C)
Oxygen O*=0.29 (DO<0.1 ppm)
O =0.29 (DO≥0.1 ppm and sensitized high carbon
*
O’=0.281 (All DO levels)
wrought and cast material)
O*=0.14 (DO≥0.1 ppm and non-sensitized material)
Strain Rate ε̇*=0 (ε̇>7 % s-1)
ε̇’=0 (ε̇>0.4 % s-1)
ε̇*=ln(ε̇/7)
ε̇’=Ln(ε̇/0.4)
(0.0004 %·s-1≤ε̇<7 % s-1)
(0.0004 % s-1≤ε̇<0.4 % s-1)
ε̇*=ln(0.0004/7)
ε̇ =ln(0.0004/0.4) (ε̇<0.0004 % s-1)

(ε̇<0.0004 % s-1)

Further work has been done on the threshold values for stainless steel by De Baglion et al. [43].
Through the study of fatigue endurance in vacuum, air and PWR conditions, the complex
relationship between the test conditions, environmental conditions and the fatigue life (number
of cycles to failure) of stainless steel specimens was highlighted. It was confirmed that in an inert
environment, either decreasing the strain rate or raising the temperature in isolation did not
negatively influence fatigue life of the stainless steel specimens. Interestingly, in the same
environment, the combination of low strain rates and high temperatures had a deleterious effect
on fatigue life, due to dynamic strain aging. As expected, the negative impact of a PWR
compared to an inert environment, was demonstrated, with decreasing strain rate resulting in a
significant decrease in the fatigue life of the specimens. These observations demonstrate the
complexity that will be present when predicting fatigue lives from test data produced in
aggressive environments. It was concluded that, due to the interactions between the parameters
themselves and separately with the environment, it was unreasonable to predict fatigue life
simply by multiplying each available independently evaluated parameter together.
Nickel-Based Alloys
NUREG/CR-6909 Rev 1 [3] highlights a lack of data for nickel-based alloys with which to
define the thresholds for the nickel-based alloy Fen parameters. For the temperature parameter,
data were available from 50 °C to 325 °C and showed that environmental effects were
insignificant below 50 °C and saturated at 325 °C. The increased amount of nickel-based alloy
fatigue data led to the introduction of a lower bound to the temperature parameter in Rev 1,
Table 3-4. As a result of the change to the limits of the temperature Fen parameter the calculation
of its value has also changed. The calculation now reflects the smaller active range of this
parameter but still saturates (goes to unity) at temperatures of 325 °C. However, NUREG/CR-
6909 Rev 1 also states that due to the “very limited fatigue data for Ni Cr Fe alloys, the effects of

3-9
13321451
LWR Environmental Fatigue

temperature on the fatigue lives of these materials were also assumed to be similar to those
observed in austenitic stainless steels”.
The dissolved oxygen Fen parameter was redefined in NUREG/CR-6909 Rev 1 which states that
the value for the transformed DO parameter is 0.14 and DO levels less than 0.1 ppm. At DO
levels of 0.1 ppm or greater, the transformed DO parameter decreases to 0.06. In Rev 0 this
threshold was defined in terms of either NWC BWR or PWR/HWC BWR chemistries rather than
a defined DO threshold to distinguish between chemistry regimes. The improved fit to the
additional data has also resulted in a change to the magnitude of the dissolved oxygen parameter.
Table 3-4
Fen parameters and their threshold values for nickel-based alloys. Note an editorial change
between Rev 0 and Rev 1 labels the transformed parameters with an asterisk for all
materials.
Fen
Parameter Rev 0 Rev 1
Limit
T’=T/325 (T<325 °C) T*=0 (T<50 °C)
Temperature
T’=1 (T≥325 °C) T =(T-50)/275 (50 °C≤T≤325 °C)
*

O’=0.09 (NWC BWR Water) O*=0.06 (DO≥0.1 ppm)


Oxygen ’
O =0.16 (PWR or HWC BWR water) O*=0.14 (DO<0.1 ppm)
ε̇*=0 (ε̇>5 % s-1)
ε̇’=0 (ε̇>5 % s-1)
ε̇*=ln(ε̇/5)
ε̇’=Ln(ε̇/5)
Strain Rate (0.0004 %·s-1≤ε̇≤5 % s-1)
(0.0004 % s-1≤ε̇<5 % s-1)
ε̇*=ln(0.0004/5)
ε̇ =ln(0.0004/5) (ε̇<0.0004 % s-1)

(ε̇<0.0004 % s-1)

Despite presenting specific expressions for Fen values for nickel-based alloys, NUREG/CR-6909
Rev 0 and Rev 1 both state that the database available for these alloys is not sufficient to
properly define the Fen parameters and their thresholds. An example of this is the lack of
sufficient data to define a threshold strain amplitude below which environmental effects become
negligible. NUREG/CR 6909 Rev 1 specifically states that work on nickel-based alloys should
be recommended for any future research activities to capture the interaction between temperature
and dissolved oxygen. It should be noted that NUREG/CR-6909 Rev 1 [3] recommends that the
austenitic stainless steel design curve should be used in conjunction with the nickel-based alloys
Fen equation to predict the fatigue life of these metals in LWR environments.
Summary
NUREG/CR-6909 Rev 0 and Rev 1 both highlight a lack of data inhibiting the determination of
threshold values for the Fen parameters for all materials. Typically, sufficient data spanning the
entire range of values are not readily available and consideration of values near thresholds are
not examined. An example of this is the saturation temperature for ferritic steels at 325 °C that is
due to a cut off in available data above this temperature. Furthermore, it is stated for all materials
that the definition of the strain amplitude threshold for the onset of environmental effects is very
limited due to the length of test times involved in obtaining data in LWR conditions at such low
strain amplitudes. Therefore, due to the lack of sufficient data covering a range of conditions
relevant to the determination of Fen parameter thresholds, this KG remains a significant gap.

3-10
13321451
LWR Environmental Fatigue

It is clear that the relationship between the individual Fen parameters does affect the fatigue life
of stainless steel specimens. The evolution of dynamic strain aging, for example, links the strain
rate and temperature parameters to the detriment of the fatigue life of the specimen. Furthermore,
the interaction of these parameters and the environment is also complex and not fully defined.
Therefore, significant work remains to be done to fully map out the interdependencies of the
parameters themselves, as well as the impact of the environment on them, before the
methodology presented in NUREG/CR-6909 can be fully verified. In light of the above
discussion and work done towards KG 10 and13, the amount of S-N data available to determine
if methodology used by NUREG/CR-6909 in the Fen expression is valid under all possible
conditions is insufficient. As a result, KGs 10, 13 and 47 should be combined and remain open.

3.2.3 Knowledge Gap 20 (Previously Classified as Category C, Medium Priority)


Knowledge Gap Definition
This KG relates specifically to austenitic stainless steels. Several issues are identified:
1. NUREG/CR-6909 acknowledges conservatism in its model regarding the influence of DO
level. This particularly applies to some grades of stainless steel in high-DO water. Further
refinement of the model to recognize an effect of DO (i.e., a difference between PWR and
BWR/ Normal Water Chemistry (NWC)) may be warranted.
2. There are no data on the influence of DO in PWR water containing boric acid and lithium
hydroxide (i.e., under transient conditions).
3. Only limited data are available for BWR Hydrogen Water Chemistry (HWC). Although
PWR data may be bounding, this remains to be established.
4. The effect of steel sulfur content on fatigue initiation life of stainless steel has not been
established—it has a substantial influence on fatigue crack growth rate.
Research & Development Need
Issue (1) is currently being revisited in ongoing NRC work and there may be sufficient data to
revise the model for BWR NWC. If insufficient data are available to justify a change, further
tests are warranted to better establish the influence of DO level on the fatigue life of austenitic
stainless steels. There is a need to establish if a benefit is likely from more data regarding (2) and
(3). Item (4) warrants study since this effect may be significant.
Ancillary Information
Chapter Section Subsection(s) Page(s) Applicable Materials
9 Discussion of Laboratory Data 10-5 to 10-6 Stainless Steel
Knowledge Gaps

Background
Since it has been clearly established that the fatigue life of stainless steels in LWR environments
depends on corrosion potential which is a function of dissolved oxygen, the use of a single value
for the O’ parameter in the initial version of NUREG/CR-6909 [8] in stainless steel may have

3-11
13321451
LWR Environmental Fatigue

been erroneous. The KG Analysis document [6] further identifies a lack of data on the impact of
dissolved oxygen in PWR environments containing lithium hydroxide and boric acid, which may
occur during startup and shutdown transients. Furthermore, the authors of the KG Analysis
report considered that insufficient data were presented in NUREG/CR 6909 Rev 0 to substantiate
the claim that PWR data bounded stainless steel fatigue behavior under BWR HWC conditions.
It was also recommended that the impact of sulfur on fatigue initiation in stainless steels be
investigated, since a beneficial effect on crack growth had been reported under some conditions.
Review of Activities Relevant to KG 20
Issue 1): Influence of DO on the Fen Expression for Stainless Steel
Although NUREG/CR-6909 Rev 0 [8] reported that lower fatigue lives for stainless steel were
measured in low DO environments when compared to high-DO water conditions, the DO
parameter in NUREG/CR-6909 Rev 0 was set to a single value irrespective of DO level.
NUREG/CR-6909 Rev 1 [3], however, recognized this issue and provided two values for use in
the Fen expression: 0.29 for low-DO environments and 0.14 for high-DO environments.
Issue 2): S-N Data investigating the Influence of DO on Fatigue Life in Boric Acid and Lithium
Hydroxide Containing PWR Environments.
No further work since the KG Analysis document was issued has been identified that directly
relates to this part of the KG. However, since oxygen has been shown to be beneficial for fatigue
life in high-temperature water (as discussed above), it is likely to be conservative to assume that
behavior under transient oxygen conditions is bounded by PWR data.
Issue 3): Data Supporting the Use of PWR Data to Bound BWR HWC Applications.
No further work since the KG Analysis document was issued has been identified that directly
relates to this part of the KG. However, work done on stainless steels by Schuler et al. [51] in
400 ppb DO BWR water demonstrates that the data generated on stabilized stainless steel is
conservatively described by the NUREG/CR-6909 Rev 0 mean water curve.
Issue 4): The Impact of Sulfur Content on the Fatigue Life of Stainless Steel S-N Specimens.
A small amount of work has been identified in the public domain since the publication of the
original KG Report [6] that addresses the impact of sulfur content on the fatigue life of stainless
steels. Asada et al. [52] performed fatigue life tests in PWR primary conditions on 316 stainless
steel and found that hold periods of 100 and 1000 hours at maximum strain had no effect on the
fatigue life. This study was based on a limited data set of four points and using 316 stainless steel
with a sulfur content of 0.001 %, which was below the threshold for retardation defined by Tice
et al. [53] in crack growth studies. Therefore, hold times would not be expected to have an
impact upon the materials fatigue life in these tests. However, even for a high sulfur material,
there may be no observable extension in fatigue life. This may be due to sulfur being washed out
of small cracks during initiation before the sulfur can improve corrosion resistance in the first
instance. In the later stages of fatigue life, where sulfur can become trapped within a crack, there
would be little environmental enhancement due to the high effective stress intensity factors
(ΔKs). A discussion of the effect of sulfur on crack growth testing is presented within KG 28 in
Section 4.

3-12
13321451
LWR Environmental Fatigue

Summary
This KG relates to several different issues. Regarding, Issue (1), the introduction of additional
DO terms for the stainless steel Fen addresses the concerns regarding the original Fen equations
not accounting for the longer fatigue lives for specimens tested in NWC BWR environments.
Very few S-N data are available to support items (2), (3) and (4). KG 20 should therefore remain
open. However, Issues 1, 2 and 4 of this Gap should be considered for merging with Gaps 10 and
13.

3.2.4 Knowledge Gap 7 (Previously Classified as Category C, High Priority)


Knowledge Gap Definition
Mechanistic understanding leads to the expectation that the degree of environmental
enhancement of fatigue damage should depend on strain amplitude. This is not consistent with
the Fen factor approach.
Research & Development Need
Further analysis of available test data is required to determine the extent to which the
environmental enhancement factor is a function of strain rate. If significant, an alternative to the
Fen approach would be preferred.
Ancillary Information
Chapter Section Subsection(s) Page(s) Applicable Materials
9 Critical Review of Dependence of the 9-4 to 9-5 All
Design Code Fen Factor on Strain
Developments Amplitude

Background
NUREG/CR-6909 Rev 1 [8] describes Fen as a function of strain rate, temperature, dissolved
oxygen and sulfur content (the latter only applies to ferritic steels). It is generally understood that
crack nucleation and Stage 1 and 2 crack growth involve different mechanisms, and the
proportion of life represented by each stage changes as fatigue transitions from low to high strain
amplitudes. Therefore, from a mechanistic standpoint, it is not unreasonable to expect a
dependence of Fen on strain amplitude. However, largely due to the paucity of high-temperature
water data within the high cycle regime, NUREG/CR-6909 makes the assumption that the effect
of the high-temperature water environment (relative to room temperature air) is independent of
the strain amplitude. This is equivalent to assuming the same effect of environment on crack
nucleation as is observed on short and long crack growth.
Review of Activities Relevant to KG 7
Evidence for a dependence of Fen on strain amplitude, with Fen increasing with decreasing strain
amplitude, has been observed through limited testing on both ferritic and stainless steels using
multiple testing methods. This includes work done by Spatig et al. [54] (stress control) and Platts
et al. [55] (strain-control) for stainless steel and Kammerer et al. [56] for low alloy steel. This
position is further supported by the work using strain control methods on ferritic and stabilized
stainless steels by Herter et al. [51]. The latter study suggests a greater effect of the environment
on nucleation than growth since nucleation dominates fatigue life at low strain amplitudes. Other

3-13
13321451
LWR Environmental Fatigue

factors such as an impact of the environment on the fatigue limit may also be significant in the
apparent strain amplitude dependence of Fen. Although NUREG/CR 6909 Rev 1 acknowledges
this dependence stating “These results indicated that environmental effects may be more
pronounced at low strain amplitudes compared to high strain amplitudes”, it nevertheless
continues to adopt a strain range independent definition of Fen based on paucity of data and ease
of use of the resulting equations. By not considering this strain amplitude dependence of Fen,
NUREG/CR 6909 may be unduly conservative at high strain amplitudes but potentially non
conservative at low strain amplitudes towards the high cycle regime.
For stainless steels, the total life approach described by Mann et al. [57] uses a combination of
alternative crack initiation curves and standard fracture mechanics assessments to calculate the
lifetime of a test specimen. The alternative initiation curves were produced as a result of J-
integral crack growth assessments performed on the striation counting data obtained from
cylindrical fatigue test specimens in air and water. These curves could be used to determine the
number of cycles required to initiate a 250 µm crack, with the remainder of life determined by
the use of fracture mechanics. Mann et al. demonstrated that, between 250 µm and failure (i.e.,
short to long crack growth), there is good agreement in the crack growth rates determined by
striation counting and established crack growth relationships. This enabled the number of cycles
to a 250 µm crack to be determined from conventional fatigue life data. However, by 250 µm
there is a substantial increment of crack growth such that the impact of the environment on
nucleation cannot be determined directly from this work. In contrast, Kamaya [58] reports good
agreement between fatigue crack growth tests and striation counting in fatigue life specimens
based on an effective stress intensity factor, ΔKeff, growth law between 20 µm and specimen
failure and concludes that there is not a significant impact of the environment on nucleation. In
the earlier work of O’Donnell [59] on ferritic steels, nucleation was defined as 50 µm, the point
at which continuum mechanics became applicable. By defining nucleation as 20 µm for most
austenitic stainless steels, Kamaya extends back beyond the validity of continuum mechanics
into microstructurally small crack growth.
Additional support for an effect of the LWR environment on nucleation for austenitic steels
comes from a quantitative application of the R5 methodology to the NUREG/CR-6909 air and
water curves. Applying the R5 fatigue life partition equations (which have been validated in air
but not high-temperature water) to the NUREG/CR-6909 curve (which is intended to apply to
ambient or high temperature air) produces a nucleation life (to 20 µm) for 0.6 % strain amplitude
of 1759 cycles. However, for a 300 °C PWR environment, NUREG/CR-6909 indicates a fatigue
life at failure less than this for even relatively fast strain rates of 0.3 % s-1 or less. This strongly
suggests that fatigue nucleation is affected by the environment and, therefore, contradicts the
conclusions of Kamaya above.
Whilst it is likely that the impacts of the environment on nucleation and growth are not identical
(as assumed in the Fen approach), the available data and mechanistic understanding suggest that
there should be a significant effect of the environment on nucleation.
Summary
KG 7 is concerned with the lack of understanding of how LWR environments impact the
nucleation of a fatigue crack and the subsequent stages of crack growth. There are several
different mechanisms—proposed in the literature—to explain corrosion fatigue that would fit the

3-14
13321451
LWR Environmental Fatigue

observations indicating a strain amplitude dependence of Fen. However, there are insufficient
data and understanding to effectively describe the relationship and to clarify which is the
responsible EAF mechanism. Additionally, the information in the literature is contradictory
regarding the environmental effects on nucleation and improved understanding of this aspect is
required to support the total life prediction model (see also discussion of the R5 assessment
methodology in Section 5). With this in mind KG 7, therefore, remains open.

3.2.5 Knowledge Gap 34 (Previously Classified as Category A, High Priority)


Knowledge Gap Definition
Very few data are available to establish the influence of PWR secondary water on EAF.
Research & Development Need
Generation of relevant data may be appropriate, e.g., for steam generator shells or tubing, but a
benefit needs to be established first.
Ancillary Information
Chapter Section Subsection(s) Page(s) Applicable Materials
3 Fatigue Initiation Life Austenitic Stainless Steels 3-16 to 3-17 All

Background
At the time of writing the original KG analysis [6] it was unclear if, unlike the EFEM procedure,
the equations in NUREG/CR-6909 Rev 0 were intended to be applicable to PWR secondary
environments. Furthermore, the data supporting the use of the either the EFEM or NUREG/CR-
6909 procedures in secondary environments was unclear. The gap analysis report, therefore,
identified a need to generate EAF data to establish the applicability of these procedures to
secondary chemistry environments.
Review and Summary of Activities Relevant to KG 34
The publication of NUREG/CR-6909 Rev 1 [3] has clarified that the equations and procedures
contained within that document are intended to be applicable to PWR secondary environments.
However, no further work has been undertaken to generate fatigue life data in secondary
environments since the publication of the KG Analysis [6] and, therefore, KG 34 remains open.

3.2.6 Knowledge Gap 21(Previously Classified as Category C, Low Priority)


Knowledge Gap Definition
The existing data for nickel-based alloys (Alloys 600, 182, and 82) are limited. NUREG/CR-
6909 provides Fen factors for nickel-based alloys based on these data and Regulatory Guide
1.207 recommends that these factors be applied to the new austenitic stainless steel air curve.
Data on Alloy 690 and its weld metals are very limited.
Research & Development Need
Further data on nickel-based alloys would lead to refinement of the Fen values; however, the
effects are significantly less than for stainless steel and, in practice, Stress Corrosion Cracking
(SCC) in Alloy 600 is likely to be more significant in PWR primary coolant, especially for

3-15
13321451
LWR Environmental Fatigue

components which operate at the highest temperatures. However, this is unlikely to be the case
for Alloy 690, which is much more resistant to SCC. More data to confirm behavior for Alloy
690 (and weld metals) may therefore be warranted.
Ancillary Information
Chapter Section Subsection(s) Page(s) Applicable Materials
10 Discussion Discussion of 10-5 to 10-6 Nickel-Based Alloys
Knowledge Gaps

Background
It was highlighted in both NUREG/CR-6909 Rev 0 and Rev 1 [3; 8] that the amount of data
available was very limited for some nickel-based alloys. In comparing nickel-based alloy data to
austenitic stainless steel it was observed that nickel-based alloys behaved in a similar manner to
austenitic stainless steels with respect to fatigue life in air. With this in mind, both NUREG/CR
6909 and Regulatory Guide 1.207 recommended that the austenitic stainless steel air curves
could be used for nickel-based alloy Fen evaluations. NUREG/CR-6909 Rev 1 further stated that,
due to the small amount of data available for nickel-based alloys in LWR environments, the
austenitic stainless steel Fen expression could also be used for nickel-based alloys instead of the
material-specific expression provided.
Review of Activities Relevant to KG 21
The KG Analysis report [6] highlighted that research into nickel-based alloys under PWR
conditions tends to focus on SCC in Alloy 600 rather than generating S-N data to support the use
of a specific Fen expression or mean air curve. In practice, the use of Alloy 600 is being phased
out for new plants in favor of Alloy 690 and very few additional data points have been added
since NUREG/CR 6909 Rev 0 was released. Some limited data are now available on Alloy 690.
Jang et al. [60] produced LCF data on Alloy 690 and the weld metal 52M in air and LWR
conditions which demonstrated that there was no discernible difference in the fatigue life of the
two nickel-based alloys in either environment. There was a reduction in fatigue life in LWR
conditions relative to air and an effect of strain rate, both of which were similar to the effects
seen for lower chromium alloys such as Alloy 600 and 182, as reported by ANL and Higuchi et
al. [61], but smaller than for stainless steel. It was concluded that the predictive models for
nickel-based alloys proposed in NUREG/CR-6909 and by the JSME [61] were reasonable in
both air and water. Additionally, the data produced in the study by Jang et al. supported the
discussion in NUREG/CR-6909 Rev 1 by demonstrating that the impact of environment on the
fatigue life of nickel-based alloys was significantly less than that observed for austenitic stainless
steels. Overall the new nickel-based alloy data indicate that the approach of Regulatory Guide
1.207 Rev 1 [2] and the recommendation in NUREG/CR-6909 Rev 1 [3] to use the austenitic
stainless steel curves and Fen expression for nickel-based alloy assessments are conservative.
Summary
In light of the above discussion and work done towards KG 21 the amount of S-N data available
to support a refined Fen expression for nickel-based alloys remains insufficient. However, the
data that have been produced indicate that nickel-based alloys will be bounded by stainless steel
data. Additionally, no work has been found that shows that the available assessment curves are
non-conservative. KG 21, therefore, remains open.

3-16
13321451
LWR Environmental Fatigue

3.3 Complex Waveform Testing

3.3.1 Knowledge Gap 17 (Previously Classified as Category C, High Priority)


Knowledge Gap Definition
NUREG/CR-6909 recommends a “modified rate approach” for which a unique Fen factor is
determined for each cycle. Only very limited test data are available to substantiate the modified
rate approach or the use of partial Fatigue Usage Factors (FUF).
Research & Development Need
Test data to validate the “modified rate approach” or an alternative procedure are required.
Ancillary Information
Chapter Section Subsection(s) Page(s) Applicable Materials
9 Critical Review of Design The Use of Average 9-7 to 9-8 All
Code Developments Fen Factors

Background
Currently NUREG/CR-6909 [3; 8] recommends a “modified rate approach” for calculating cycle
specific Fen factors. In essence the cycle is divided into a number of strain range increments and
a Fen value calculated for each of these sections. These factors are then weighted according to the
ratio of the increment strain range to the total range before being summed to give the total Fen
factor. The basis of this KG is the lack of experimental data to underpin this method of
calculating Fen which relates to the amount of justification behind the adoption of this approach
in NUREG/CR-6909 [8]. There are sufficient data available from the international community to
demonstrate that the modified strain rate approach is better than using a straightforward average
strain rate for the whole cycle and that the approach is suitable for simple waveform shapes such
as sine or triangular [62]. The issue with the modified strain rate approach is that it does not
account for the observation that the position of a slow rising portion of a complex waveform
within a transient cycle influences the degree of environmental enhancement it imparts to the
total damage in the cycle [16]. This means that the modified strain rate approach can then
become overly conservative for complex and/or plant-realistic loading.
Review of Activities Relevant to KG 17
Hirano et al. [63] investigated the effect of sine wave loading (continuously changing strain rate)
on carbon steel specimens in BWR conditions using hollow specimens. This work showed that
the predictions of the modified strain rate approach and the impact on Fen showed good
agreement with the experimental data. Testing on 316L stainless steel was performed by Hong et
al. [64] in 310 °C air and LWR conditions. The waveform in this work comprised one strain rate
for the first 75 % of the strain amplitude followed by another from 75 % to 100 %; the falling
rates mirrored this configuration. Hong et al. also concluded that the predictions of fatigue life
obtained from the modified strain rate approach were reasonable when compared to experimental
evidence. However, several other studies addressing the impact of variable strain rates within
plant-relevant cycles showed the inadequacy of the modified strain rate approach when the
waveforms are asymmetric [16; 65; 66; 52; 67]. In cases such as these, the environmental
influence on fatigue damage is unevenly distributed about the cycle depending on value of the

3-17
13321451
LWR Environmental Fatigue

strain rate and where it is in the cycle with respect to strain. In particular, slow strain rate ramps
close to the upper, tensile portion of the cycle had a significantly greater influence on reducing
fatigue life than slow ramps in the lower tensile, or compressive, parts of the cycle. These new
variable strain rate test data are discussed further below in the context of KG18. Damage-based
models which directly deal with Fen have been proposed; these models support KGs 39, 41 and
42 and are described further in Section 5.2.1.
Summary
The modified strain rate approach has been shown to be sufficient to predict the failure of
laboratory specimens subject to simple symmetrical waveforms in which the environmental
impact on fatigue damage is evenly distributed throughout the cycle. However, it has been shown
that for complex cycles more relevant to plant transients the environmental effect on fatigue
damage is not constant over the cycle, and current averaging methods for obtaining the effective
strain rate are not appropriate. Alternative models, such as the weighted Fen approach [44; 66],
have been proposed and appear to show improved predictive capability and, therefore, significant
promise in optimizing conservatism in assessments of fatigue in LWR environments.
Nevertheless, these models are in their infancy and will require a significant amount of data to
validate them. Consideration should also be given to their relevance to other materials such as
carbon and low alloy steels, for which supporting data will be required. KG 17, therefore,
remains open.

3.3.2 Knowledge Gap 18 (Previously Classified as Category B, High Priority)


Knowledge Gap Definition
There is no basis available for defining the treatment of non-contiguous cycle pairs with regard
to both crack initiation and growth in LWR environments. This is because of a lack of
mechanistic understanding on which to formulate rules and a lack of test data with which to
validate them.
Research & Development Need
Mechanistic understanding is required as a basis for formulating rules for the treatment of non-
contiguous cycles. Validation testing with complex transient loading should be performed and
compared with predictions of current assessment methods for EAF.
Ancillary Information
Chapter Section Subsection(s) Page(s) Applicable Materials
9 Critical Review of Design Combined Cycles 9-8 to 9-9 All
Code Developments

Background
Plant loading histories are made up of definable and individual transients; however, the design
transients are often conservative and the order in which these transients occur may be unknown.
Therefore, plant designers commonly generate a cycle representing the most severe stress range
possible by combining transient pairs. The Fen factors for each part of the constructed cycle
would then be need to be calculated and the maximum value applied to the usage factor based on
the total strain range of the whole cycle. It should be noted that the transient pairs used to create

3-18
13321451
LWR Environmental Fatigue

the bounding cycle could potentially be separated by a significant amount of time. Therefore, the
application of a single Fen factor to the transient pair may be unreasonable. This is particularly
true if the hold time between the transient pairs had an influence on the overall fatigue damage.
There are further complications when applying the Fen expression in NUREG/CR 6909 [8]to
transient pairs derived from plant monitoring. Firstly, it is unclear how to define the strain rate
for these cycles, and secondly the cycle with the highest stress range may not have the greatest
Fen factor. This means that after the environmental influence is accounted for in the usage factor
the cycle with the highest stress range may not be the most damaging.
Review of Activities Relevant to KG 18
There have been a number of tests carried out in the last few years studying the influence of
complex loading cycles or cycles including hold times which support this KG. These studies
have informed the development of modified assessment procedures that provide improved
methods for the calculation of strain rate for complex cycles; these procedures are discussed in
Section 5.2.1 in the context of KGs 39, 41 and 42. The following discussion addresses the
experimental work.
Variable Strain Rate Testing
Le Duff et al. [16] compared the effect of complex waveforms on the fatigue life of stainless
steel specimens. The complex waveform was related to typical thermally induced stresses, due to
down and up shocks. Variants of the complex waveform, in which the sequence of different
portions of the complex waveforms was rearranged to determine the impact on a calculated Fen,
were also investigated. The tests were performed isothermally on austenitic stainless steel in
simulated PWR primary coolant at 300 °C. This work demonstrated that fatigue environmental
factors were significantly lower when the slow rise was located in the compressive part of the
waveform than in the tensile portion. Furthermore, for all four transients the environmental
enhancement was less than the value of Fen predicted by NUREG/CR 6909. The observations
made by Le Duff [16] have been further reinforced by work published by Seppanen et al. [65]
and Currie et al. [66]. The work presented in these two studies consists of a 4-stage waveform
where the strain rate of each stage can be varied. The positive loading ramp rates were varied
with different strain rates applied in the compressive and tensile portion of the cycle. Initial data
from this waveform type are broadly consistent with the work of Le Duff, and indicates that a
rapid rise from minimum to maximum strain is far less damaging than when the slow ramp is
located in the mean to maximum strain portion of the cycle, which gives the most significant life
reduction. Seppanen et al. used a 4-stage waveform based on a safety injection system (SIS)
transient akin to that used in Le Duff’s work. Two distinct dual strain rate waveforms (0.01 % s-
1 and 0.0004 % s 1) were used with the slow rate applied to either the ramp from mean to
maximum or from minimum to mean. Currie et al. also used a range of multistage waveforms
including simulated SIS transients as well as multi-stage linear ramps with fast and slow rise
times positioned in different portions of the cycle. All these data support the conclusion that Fen
predictions using NUREG/CR-6909 are too severe due to the strain rate in the Fen calculation
being based on an average of the rising ramp rate. Seppanen et al. [65] suggested that Fen should
instead be calculated based on plastic strain rate as opposed to the average across the cycle.
Analysis of data presented by Currie et al. using these waveforms confirms this approach but
suggests that Fen is proportional to the square of plastic strain rate rather than directly
proportional as suggested by Seppanen et al. Furthermore, Currie et al. advocate the use of a

3-19
13321451
LWR Environmental Fatigue

weighted damage model, referred to as the Strain-Life-Weighted (or Weighted Fen) Model,
which is simpler to apply than the plastic strain rate approach that requires load-displacement
hysteresis data to predict the development of plastic strain during the cycle. Both models appear
to provide the basis for optimizing conservatism for plant-relevant transients. Further discussion
is provided in Section 5.2.1.
Effect of Hold Times
Solin et al. [19; 68] investigated the implications of the hold periods between non-contiguous
plant load cycles for fatigue life, mostly on stabilized stainless steel. Testing was carried out in
room temperature air using either triangular or sinusoidal waveforms with a range of strain
amplitudes and strain rates. The holds were applied at different intervals in the cycles or at zero
load for various time periods, at temperatures up to 450 °C. This was intended to simulate
primary piping subjected to periodic thermal fatigue cycles separated by extended periods at
operating temperature; the higher temperature enabled the use of much shorter hold times
compared to plant. This work demonstrated three points. Firstly, holds, regardless of where in the
cycle they are applied, have a positive effect on fatigue life at low strain amplitudes but little
effect at strain amplitudes above 0.4%. Secondly, an increased cyclic hardening effect was
observed when cyclical loading was restarted after a hold. Thirdly, the positive effect of hold
times on the fatigue life of the specimens was shown to increase as the length of the holds
became longer. More limited testing was carried out on Types 304L and 316L stainless steels,
and showed generally similar behavior. These results reinforce and expand on earlier work by
Solin et al. [69] in which a link between the increase in cyclic hardness (observed after a hold
period) and fatigue life extension was found.
The impact of hold times on the transferability of laboratory data to plant operation was further
explored in a collaboration between VTT, Rolls-Royce, Areva GmbH, E.ON and the University
of Manchester (ADFAM project), and reported by Karabaki et al. [70]. This study extended the
temperature of the hold periods down to temperatures relevant to LWRs (325 °C) but also sought
to establish a mechanistic understanding to complement the empirical relationship between hold
times and fatigue life. As with the study by Solin et al., cyclical loading was performed in air at
ambient temperature with the holds at elevated temperature (325 - 450 °C) at zero stress. Similar
to the work discussed above, the data showed that hold periods produced extended fatigue lives
compared to standard tests up to strain amplitudes of 0.4 %.
Asada et al. [52] performed testing on hollow specimens in a PWR primary environment at
325 °C, using sawtooth loading at a strain amplitude of 0.6 % and a strain rate of 0.001 % s-1 for
the strain increasing region. A hold period of either 100 h or 1000 h was imposed after 30 % of
the cycles-to-failure for a test run to completion. No influence of holds on fatigue life was
observed, which is consistent with the tests in air with hot holds performed by Solin et. al., and
Karabaki et al. [70] who observed no benefit of holds for strain amplitudes above 0.4 %.
Rudolph et al. [29] also investigated the impact of hold times on stainless steel specimens tested
in air (240 °C) and PWR (240 °C) primary water. In this case, hold periods of 72 hours at 290 °C
were applied at 30 % and 70 % of average specimen life (as calculated from reference
specimens). These tests were performed at 0.35 % and 0.6 % strain amplitudes. Like Asada et al.
and Solin et al. these tests were not performed at low enough strain amplitudes to expect a
significant impact of hold times on fatigue life and, therefore, it is unsurprising to note that
Rudolph et al. also observed no hold time effect.

3-20
13321451
LWR Environmental Fatigue

With respect to mechanistic understanding of the effect of hold times on the fatigue life of
stainless steels, the work done by the ADFAM collaboration theorized that a thermally-
dependent mechanism was involved [70; 71; 72]. This theory was supported by observations of
strain relaxation in the specimens during hot holds and a threshold duration for holds to obtain a
significant life extension. Striation counting on tested specimens demonstrated that hot holds had
no impact on long crack growth. Furthermore, no significant differences in bulk hardness and the
dislocation structures before and after holds was observed. It was postulated that hot holds could
affect fatigue crack nucleation and/or short crack growth, possibly through an effect on the build-
up of localized stress that is a precursor to crack nucleation. Additional test data and
microstructural analysis would be required to confirm this hypothesis.
Based on the above results, Solin et al. [19] proposed so called ‘bonus factors’ to account for the
enhancement to life when steady state operation exists between fatigue transients (further
described in the KG 2 discussion). The benefit was ascribed to intermediate annealing, and it was
suggested that a similar effect could be expected in an operating plant. Reese et al. [73]
suggested that the main benefit of hold times under these conditions was in the high cycle fatigue
region, due to a higher endurance limit. They concluded that the data were equivalent to a
reduction factor of between 1.0 and 1.12 on strain amplitude. In a later paper, Reese et al. [74]
extended this work further using numerical models to more accurately describe the hold time
effects and concluded that the data corresponded to a reduction in fatigue usage factor of up to
32 %.
Another method to assess lifetimes in NPP components subject to holds was proposed by
Karabaki et al. [70]. In this work an Fhold factor was proposed which is then multiplied by the
Fen to give a modified life reduction factor, Freal. Through a comparison of test results
containing holds with standard S-N tests, Fhold values of between 0.11 and 0.84 were
determined, the former being applicable at very low strain amplitude. Karabaki et al. stated that,
whilst this approach to include a benefit for hold effects had been incorporated into the German
KTA code in 2013 [17], the value of Fhold remains set at unity until a better understanding of
the phenomenon becomes available [70]. It was concluded that further investigation of the
mechanisms behind the observed life extension due to holds and more data were required before
definitive values of Fhold could be adopted.
Work using blunt notched compact tension specimens by Seifert et al. [75; 76] has investigated
the impact of hold times at minimum, mean or maximum loads on fatigue initiation and short
crack growth of stainless steel in BWR HWC and NWC environments. It was found that the two
environments produced differing responses. In the HWC environment, an increase in the so-
called “technical” fatigue initiation life (which includes short crack growth up to about 250 m)
was found in the presence of hold periods at mean or maximum load, but not at minimum load.
This increase appeared to saturate for hold times longer than 12 hours. It was suggested that this
was because any microcracks would only be exposed to the environment during holds at loads at
or above mean load. However, no effect of holds was observed for subsequent crack growth. In
BWR NWC, hold times were found only to have a significant effect on sensitized material, for
which times to fatigue initiation were reduced when compared with cyclic loading. This was
attributed primarily to a contribution from SCC.

3-21
13321451
LWR Environmental Fatigue

The effect of hold times on the fatigue life of carbon steel has been studied by Brown et al. [77],
using blunt notch CT specimens. Experiments were carried out in room temperature air, 300 °C
air and high-temperature oxygenated water (2 ppm Li, 11-14 ppm DO). This work defined the
N25 failure criterion as crack lengths of 500 µm at ambient temperatures and 300 µm at 300 °C.
The results of the tests conducted were compared to the predictions in NUREG/CR-6909 on this
basis. In air, the data showed a reasonable agreement with the fatigue life values predicted using
NUREG/CR-6909. Comparisons between air data collected at ambient temperature or 300 °C
suggested that hold periods had no observable impact, irrespective of temperature. Tests
conducted in the oxygenated water environment showed better agreement with NUREG/CR
6909 at shorter rather than longer rise times, with more data scatter being observed at longer rise
times. When hold times were introduced, it was found that holds implemented at minimum load
showed no effect but at maximum load the specimens required considerably more cycles to
initiate and failed to reach the target 300 µm growth. Brown et al. postulated that a reduction in
the stress intensity factor at the tip of the crack, due to crack tip blunting resulting from extended
exposure to the environment during holds at maximum load, may be responsible for the
increased life.
Summary
Since the publication of the KG report in 2011, a significant amount of work has been carried
out, mainly on stainless steels, towards the understanding of the impact of the environment on
fatigue life under complex load cycling. Testing has addressed multistage cycles with fast and
slow rise time portions and cycles simulating the cyclic variations in stress/strain produced by
thermal transients (but performed under isothermal conditions). Testing has also been performed
using trapezoidal loading with long hold times at elevated temperature. This was intended to
simulate the time between the transients used to make up non-contiguous cycles. However, most
of the latter testing was in air, and the limited standard S-N data available in water were at high
strain amplitude where no effect was observed.
It has been clearly demonstrated that the degree of damage in a complex transient occurs when
the slower rise time portion of the transient is in the high positive strain portion of the transient.
As a result, the NUREG/CR-6909 Fen methodology, combined with the use of the modified rate
approach to calculate strain rate during the cycle, significantly over-predicts damage for
transients that are representative of thermal shocks. Alternative assessment approaches, using a
weighted damage approach based on the level of plastic strain, appear to have improved
predictive capability. At strain amplitudes ≤0.4 %, a small but significant benefit has also been
identified in laboratory S-N testing in air for the period of operation between the transients
during which the material is held at a constant elevated temperature and load. This effect appears
to be related to the cyclical hardening behavior of the material directly after the hold.
Mechanistic studies suggest that the effect of hold periods is greatest on crack nucleation and/or
short crack growth since striation counting methods showed no impact of holds on long crack
growth. Despite significant progress on this topic, there is a need for additional confirmatory
work under a wider range of testing conditions. The majority of recent work is on austenitic
steels in a PWR primary environment and relevance to other materials and environments
warrants consideration. KG 18 should therefore remain open. It is noted that there is a close
interaction with KGs 15 and 16 which incorporate the additional complication of temperature

3-22
13321451
LWR Environmental Fatigue

variation during cycling which is an important aspect of many plant transients which are thermal
in origin. KGs 39, 41, 42 and 43 address EAF assessment methodologies for complex and non-
isothermal transients and are discussed in Section 5.2.

3.3.3 Knowledge Gap 11 (Previously Classified as Category A, Low Priority)


Knowledge Gap Definition
There is likely to be a significant influence of mean stress on EAF that is not adequately
quantified by existing test data.
Research & Development Need
There is a need to develop additional data to establish the influence of positive and negative
mean stress.
Ancillary Information
Applicable
Chapter Section Subsection(s) Page(s)
Materials
9 Critical Review of Design Code Review of NUREG/CR- 9-2 All
Developments 6909 Fen Factor Approach\
Mean Stress Correction

Background
ASME Section III specifies that high cycle fatigue data should be corrected for the effects of
mean stress on fatigue life and recommends the use of a modified Goodman approach. In the low
cycle fatigue regime, plasticity allows the material to shake down (component achieves structural
stability with the absence of ratcheting) to load ratios of -1 and so relieve the mean stress. This
would make a mean stress correction for the low cycle regime on a material such as austenitic
stainless steel, where the yield stress is lower than its fatigue limit, inappropriate. In contrast, for
metals such as carbon steel, where the yield stresses are excess of their respective fatigue limits,
the mean stress would persist without the presence of significant plasticity. NUREG/CR-6909
[8], states that a modified Goodman approach for mean stress in austenitic stainless steels was
applied when deriving the design curves, but no details are provided concerning the methodology
[6].
The potential implications of mean stress on the fatigue limit of austenitic stainless steels raise
interesting questions over the strain amplitude Fen limit for both the carbon/low alloy and
austenitic stainless steel models in NUREG/CR-6909. If mean stress can reduce the fatigue limit
of a material, which the Fen strain amplitude and cycle number threshold appears to be based on,
then mean stress could have a significant impact on corrosion fatigue. Based on the available
data at the time, the KG Analysis document [6] concluded that there were insufficient data to
characterize the impact of this aspect of mean stress and that there was a significant need to
determine what relationship there was, if any, between mean stress and EAF.

3-23
13321451
LWR Environmental Fatigue

Review of Activities Relevant to KG 11


KG 11 relates specifically to the need for additional data to quantify mean stress effects on
fatigue and EAF. Assessment aspects are addressed by KG 6 and KG 44, see Section 5.3.3.
Air Environment
Due to the lack of a straightforward way of applying mean stress in a low cycle fatigue regime,
which typically uses strain control methods, testing any hypothesis in this area is complicated.
Previously, the effect of persistent mean stress on fatigue life has been viewed as positive if
compressive or negative if tensile [54; 78]. By using a strain control methodology that allowed
the applied strain amplitude to ratchet, Vincent et al. [78; 79], were able to obtain a persistent
tensile mean stress in S-N tests on stainless steel in air. Under these conditions, their work
indicated that a persistent mean stress reduced the fatigue life of the test specimens by 30 %.
Kamaya [80] took the opposite approach and applied a mean stress in stress control while
monitoring the change in strain amplitude. This method allowed the resulting ratcheting of the
strain amplitude to be recorded and the tests to be compared to strain controlled tests. This study
showed a reduction in fatigue life for experiments conducted using positive mean stresses that
was due to the increase in strain amplitude due to ratcheting. However, both pre-strained and
ratcheting tests showed similar fatigue lives for the same effective strain amplitude. Work by
Spatig et al. [54] on Type 316L stainless steel in air, using solid fatigue specimens under load
control, found that both tensile and compressive mean stresses produced a life extension, which
is in contrast to what was previously believed.
LWR Environment
Mean stress experiments in high-temperature water add additional complications to the
experimental set-up, which bring further difficulties in producing data to address this KG. The
requirement to use simulated LWR coolant means that the standard extensometry used in fatigue
endurance air tests cannot be applied to solid specimens. The use of hollow specimens enables
the use of standard extensometry but at the expense of using a specimen geometry that is
different to that used to produce the ASME III mean air curve. Spatig et al. used hollow tubular
specimens to study the impact of mean stress in load-controlled experiments on 316L stainless
steel [54; 81]. When compared to Spatig’s air tests the LWR tests produced a more complex
behavior. Here, the application of compressive and 50 MPa mean stresses produced an increase
in fatigue life; whereas, mean stresses of 10 and 20 MPa resulted in a decrease. The potential
issues with directly comparing solid and hollow specimens also need to be considered, as
discussed below.
Spatig also assessed the impact of mean stress on Fen for the 316L material studied. The analysis
found that the calculated Fen factor showed a slight increase when the stress amplitude decreased
from 190 MPa to 170 MPa for zero mean stress. For positive mean stresses, it was found that the
Fen factor increased with decreasing stress amplitude much more strongly; these observations
were based on very limited data. KG 7 (Section 3.2.4) discusses the stress/strain amplitude
dependence of Fen in more detail.
A considerable proportion of the S-N data used to generate the NUREG CR/6909 water curves
were generated from hollow tubular specimens. The cyclic plastic strain behavior of hollow test
specimens has been investigated by Lee et al. [82], where stainless steel solid and hollow
specimens were cycled using strain control in room temperature air. The specimens had gauge

3-24
13321451
LWR Environmental Fatigue

lengths of 20 mm and outer diameters of 8 mm with the hollow specimens having a 4 mm inner
bore. When comparing the results for tests done with strain amplitudes of 0.3 % to 0.7 % they
found that hollow specimens had noticeably shorter lives than those of solid specimens where the
difference was greatest at high strain amplitudes and was negligible at amplitudes of 0.3 %. At
strain amplitudes of 0.7 % the hollow specimen lives were shorter than the solid specimens by a
factor of 1.5. Lee et al. analyzed their data using two-dimensional FEA modelling to investigate
the impact that the specimen geometries had on the stress and strain distributions around the
crack tip. This work showed that the internal strains were 1.6 times larger in the hollow
specimens than in the solid ones and correlated well with the difference in experimental data.
Other researchers have found up to a factor of two reduction in fatigue life for hollow compared
to solid bar specimens, with possible contributions to the difference being attributed to either
geometry and/or internal pressure. Twite et al. [42] reported data which showed a clear
difference between strain-controlled fatigue behavior of hollow and solid specimens in a PWR
environment, with the lives of hollow specimens being a factor of approximately 1.8 lower.
Possible reasons for the observed differences were discussed and it was reported verbally at the
2016 ASME PVP meeting that FEA results performed subsequent to publishing of the paper
suggested that most of the difference is due to the internal pressure in hollow specimens from
PWR coolant, but with some contribution from plastic strain peaks in the taper regions at the
ends of the specimen gauge length. Further finite element analysis by Gill et al. was
subsequently published which indicated that the calculated von Mises strain was about 20%
higher due to this effect of internal pressure and a strain amplitude correction factor could be
applied to normalize the hollow and solid specimen data [83]. These conclusions show some
similarity to the FEA work of Spatig et al. [54; 81], in which the analysis of load-controlled tests
found the strain amplitude of the hollow specimens to be greater at 200 bar than at ambient
pressure. Furthermore, when considering the impact that pressurization had on the strain
amplitude in tension and compression, it was found that the effect was greater in tension. This
would give a tensile strain amplitude of ~0.31 % and a compressive amplitude of 0.25 % (an
average of 0.28 %) for tests conducted under load control at 165 MPa. This is compared to an
unpressurised value of ~0.26 % for the same loading. However, in contrast to the work by Twite
et al. [42] and Gill et al. [83], Spatig et.al [54; 81] concluded that there would be minimal impact
on the fatigue lives of load-controlled specimens due to the mean strain imposed on the specimen
as a result of pressurization counteracting the increase in effective strain amplitude.
Summary
Work in the past five years addressing mean stress effects has been predominantly on stainless
steel in air, with the main focus being on test method development and creating baseline data in
air. Limited data have been generated in LWR conditions, and since 2012, reported tests
addressing this issue have been performed only on hollow specimens. These data need to be
assessed for compatibility with solid specimen data before significant conclusions are drawn on
the effect of mean stress on fatigue behavior in LWR conditions.
One of the aspects of the INCEFA+ test program [12] focuses on mean stress and strain effects
in air, and LWR conditions and may produce significant progress towards resolving KG 11.
Currently three options are being considered within the program which will attempt to produce
meaningful data: using mean strain with strain control, mean stress in strain control or mean
stress with stress control. The initial testing phase of INCEFA+ will narrow down which of the

3-25
13321451
LWR Environmental Fatigue

three methods is reasonable to use in the test program with the subsequent phase producing
quality data to establish the impact of mean stress or strain on fatigue life in LWR environments.
In the original KG report, this issue was assessed as low priority. More generally, there is a need
to evaluate the possibility that differences in the fatigue lives of hollow specimens and solid
fatigue specimens may have biased the existing database supporting NUREG/CR 6909 and, if
appropriate, to quantify any effect. Until further data are available, KG 11 should remain open.

3.3.4 Knowledge Gap 14 (Previously Classified as Category A, Medium Priority)


Knowledge Gap Definition
The strain amplitude and the associated number of cycles for which the consideration of
environmental effects on fatigue is not required are based on zero mean stress test data only. The
situation may be different for positive or negative mean stress. The lack of non-zero mean stress
test data prevents this analysis from being undertaken.
Research & Development Need
The need for non-zero mean stress testing was noted in Gap 8. Further test data are required to
identify the strain amplitude threshold for non-zero mean stress, both tensile and compressive.
Ancillary Information
Chapter Section Subsection(s) Page(s) Applicable
Materials
9 Critical Review of Review of NUREG/CR-6909 Fen Factor 9-6 All
Design Code Approach\ Limitation on the Number of
Developments Cycles N

Background
The Fen models described in NUREG/CR-6909 Rev 0 [8] are recommended for fatigue life
predictions below 1x106 cycles. NUREG/CR-6909 did not specifically explain the origin of this
limit; however, the KG Analysis report [6] contains the observation that the trends in data for
both ferritic and stainless steels suggest the environmental effects are small above 1x106 cycles.
A corollary to this was the implication that environmentally enhanced fatigue damage would not
occur for high frequency cycling, since the strain rates would exceed the limits of the Fen
equation. The limit on cycles for the calculation of Fen is consistent with the proposed strain
amplitude threshold for EAF as it roughly corresponds to 1x106 cycles. If this is the case, the
interpretation is purely based on data collected under zero mean stress conditions and
considering that plant components will be subjected to mean stresses it should also be
demonstrated using data produced under mean stress conditions.
Review and Summary of Activities Relevant to KG 14
No work has been identified that specifically addresses this KG. However, it should be noted that
this KG is essentially an extension of KG 11 to the high cycle fatigue / fatigue threshold region.
The significance of the HCF region of the S-N curve may increase in significance now that the
strain rate threshold has been increased to 7 % s-1 NUREG/CR-6909 Rev 1 [3] (see KGs 10 and
13). This KG should therefore be merged with KG 11.

3-26
13321451
LWR Environmental Fatigue

3.4 Non-Isothermal Loading


The majority of fatigue damage done during LWR plant operation is thermal in origin where the
applicable transients can have variable strain rates, temperatures and stresses. NUREG/CR 6909
[3; 8] accommodates varying parameters by using averages or maximum values of temperature
to calculate the temperature parameter in the Fen expression. However, the data supporting such
approaches to this aspect of complex waveforms is very limited and, considering the impact that
the averaging methods have on the cumulative usage factor, represents a significant uncertainty.

3.4.1 Knowledge Gaps 15 And 16


Knowledge Gap Definitions
Knowledge Gap 15 (Previously Classified as Category A, High Priority)
Limited data are available on the influence of variable temperature and variable strain rate within
test cycles and of the influence of out-of-phase variations of temperature and strain rate.
Knowledge Gap 16 (Previously Classified as Category C, High Priority)
Test data supporting averaging procedures for complex non-isothermal transients are very sparse,
and this represents a significant uncertainty. Therefore, the averaging procedures are based
largely on assumptions. Mechanistic understanding is required as a basis for identifying those
parts of the cycle for which a water environment is damaging. This understanding can then be
used as the basis for developing averaging procedures, which should then be validated with test
data involving cyclically variable parameters.
Research & Development Needs
Knowledge Gap 15
Testing with mixed thermal/mechanical loading both in and out-of-phase. Resolution of this
issue is likely to prove difficult, but there is the potential to justify significant benefit.
Knowledge Gap 16
Further tests similar to the Bettis testing with welded pipe fittings and thermal cycling would be
helpful in simulating plant conditions. Smaller scale testing with controlled temperature cycling
would support further mechanistic understanding.
Ancillary Information
Chapter Section Subsection(s) Page(s) Applicable Materials
9 Critical Review of Cyclically Variable 9-7 All
Design Code Parameters
Developments

Background
Many of the transients in operating LWRs are thermal in origin and the change in temperature
results in a change in stress. For a typical component containing pressurized water, a sudden
drop in temperature results in a rapid increase in stress. The temperature will vary throughout the
event with the minimum and maximum temperatures associated with the transient corresponding
approximately to the maximum tensile stress and compressive surface loading, respectively. In

3-27
13321451
LWR Environmental Fatigue

addition, the strain rate for thermal transients varies in a non-uniform manner throughout the
cycle. This raises difficulties for assessments of environmental fatigue, since it is necessary to
decide on an appropriate strain rate for the Fen calculation and also to determine whether the
fatigue damage incurred is dependent on the position in the loading or temperature cycle. An
additional issue in the case of thick-walled components is that the strain (and hence stress)
induced by a thermal transient decreases relatively rapidly through-wall. The influence of such a
strain gradient may mean that the calculation of component life using a conventional fatigue
assessment based on membrane-stressed small test specimens may be very conservative.
These three aspects of transients that induce thermo-mechanical loading are covered by four
existing KGs. KG 18 (discussed in Section 3.3.2) addresses complex cyclic loading under
isothermal conditions where two main issues arise: calculation of strain rate for complex cycles
and assignment of damage to different portions of the transient cycle. KG 15 focuses on the
influence of temperature on the degree of environmental enhancement of fatigue as stress and
temperature change in- or out-of-phase during a cycle. The focus of KG 16 is predominantly on
the influence of strain gradient for thick section components. Finally, KG 33 (Section 3.5.1)
addresses the further effects of more complex geometrical features, such as elbows, bends and
changes of bore, which may occur in real plant components.
When considering thermal transients, it is possible to have transients with identical Fen factors
but that contribute very differently to the total amount of corrosion fatigue damage done to the
material. This was highlighted in isothermal testing carried out by Le Duff et. al. [16], in which a
representative loading signal was manipulated by juxtaposing different portions of the original
waveform to give three additional waveforms each with the same calculated Fen factor. It was
then found that, although the calculated Fen factors were identical, the experimentally measured
environmental effects were significantly different. The level of data available to determine the
correct method for dealing with thermal transients was very limited at the time of the 2011 KG
review, and considering the complexity of this issue, it was judged that a considerable amount of
additional data would be required to define an appropriate assessment methodology for such
transient loading conditions. KGs 15 and 16 tackle different aspects of the same problem which
is how to properly account for damage done by thermal transients in LWR environments. KG 15
deals with both the impact of varying strain rates throughout a cycle on fatigue life as well as the
effect of varying the temperature in-phase and out-of-phase with the stress. KG 16 advocates a
combination of features testing such as the thermally-cycled stepped pipe test combined with
smaller scale testing under thermal cycling; a key part of this testing would involve investigation
into the impact of thermally-induced, through-wall, strain gradients.
Review of Activities Relevant to KG 15 and 16
Knowledge Gap 15
Some experimentation has been performed in air on ferritic reactor steel and austenitic cladding
relevant to VVER-440 reactors in Hungary [20]. In this work cylindrical specimens were
subjected to isothermal tests at 260 °C and in-phase, non-isothermal testing between 270 °C and
150 °C using strain amplitudes between 0.3 % and 2.0 %. The thermo-mechanical (TMF) tests
achieved a heating and cooling rate of 20 °C s-1 by spraying cool air onto the gauge length of the
specimen. The results showed that, for the strain amplitudes investigated, the TMF testing
resulted in shorter lives than their isothermal counterparts. This is a somewhat unexpected result
given that in air, up to 400 °C, temperature appears to have little impact on fatigue life of

3-28
13321451
LWR Environmental Fatigue

stainless steels and only a small negative impact (a factor of 1.5) on ferritic steels [8]. One
possibility is that the use of a single nozzle to rapidly cool the specimen to attain the desired
thermal cycle could have caused the specimen to bend. This bending could have had a significant
impact on the fatigue lives of the specimens tested.
Some TMF testing was conducted in air on austenitic stainless steel (X6CrNiNb18-10) by Utz et
al. [84]. This testing was done on ‘calotte’ specimens that are a square cross section (30 mm
width by 21 mm thick) with a concave spherical shape machined into the center of each face
such that the thickness of the material in the center of the specimen was 1 mm. The specimens
were then inductively heated up to temperature (either 300 °C or 270 °C) with the temperature
being monitored by two pyrometers and the surface strains measured using a non-contacting
materials-independent method known as ARAMIS [85]. Temperature cycles were performed
between 300 °C and 160 °C or 270 °C and 210 °C by cooling the specimen with pressurized air.
This method resulted in cooling rates of 75 Ks-1 and heating rates of 90 Ks-1 (300 °C to 160 °C)
or 60 Ks-1 (270 °C to 210 °C). Finite Element simulations using the Chaboche Material Model
[86] were used to analyze the results of the TMF testing. The required materials properties were
experimentally derived from standard uniaxial testing on this material. From this model an
equivalent strain range was determined for each temperature range which corresponded to strain
amplitudes of 0.38 % for the 300 °C to 160 °C cycle and 0.135 % for the 210 °C to 270 °C cycle.
The number of cycles to failure for these tests were found to lie within the data scatter of the
reference uniaxial fatigue tests. This indicated that TMF cycles in air have little impact on the
fatigue life of stainless steel specimens in contrast to the work described above [20].
Utz et al. also performed fractography on the uniaxial fatigue and TMF test specimens using
SEM and XRD measurements. Deformation-induced martensite was observed on the surface for
specimens tested under mechanical loading at room temperature. This form of martensite is soft
and ductile as opposed to the type formed by quenching stainless steels which is very hard. The
amount of the soft martensite is dependent on the strain amplitude applied where greater
amplitudes result in higher fractions of martensite. However, the phase transition from austenite
to martensite was only observed to occur at strain amplitudes of 0.3 % or higher. Utz et al. found
that the formation of deformation-induced martensite was a prelude to crack initiation, and crack
initiation occurred in the phase boundaries between the martensite and austenite or in the purely
martensitic regions of the specimen. The initial fractography done on the TMF test specimens
showed that, in air, fatigue cracks in these specimens initiate in the same fashion as standard
cylindrical specimens. They observed shear band formation and intrusions/extrusions on the
surface where the cracks initiated from the surface intrusions.
Several organizations have developed test facilities capable of performing TMF testing to
investigate the effect of variable temperatures within a cycle on EAF using Types 304L and
316L stainless steels [52; 67; 87; 88]. Furthermore, it is expected that these experiments will
provide clarity on the temperature to use in an analysis on variable temperature waveforms.
These facilities use hollow specimens which contain the test environment and can be subjected to
externally-applied mechanical loading, together with thermal cycling by alternating hot and cold
flows of water through the specimens. The thermal and mechanical loading can be varied
between in- and out-of-phase simulating different possible transient conditions. Specimen wall
thicknesses are typically of the order of 3 mm where a through-wall crack should roughly
correspond to a 25 % load drop in a standard solid LCF specimen. Although several facilities
exist, only a limited amount of data have been produced to date. Typically, these experiments are

3-29
13321451
LWR Environmental Fatigue

conducted above 0.3 % strain amplitude due to the excessive experimental timescales involved in
testing at lower strain amplitudes. At strain amplitudes greater than 0.3 % it is expected that
crack growth will dominate the life of the specimen. Therefore, the current test programs will
only shed limited light on the impact of TMF testing on crack nucleation. Nevertheless, the data
from these facilities are consistent in demonstrating that the fatigue life of the test specimens
follows the order: Max T (shortest life) < In Phase< Out-of-phase<Min T (longest life).
Recently Platts et al. [67] have reported further TMF testing in support of a new Fen
methodology, described by Currie et al. [66] and termed the weighted Fen procedure (covered in
more detail in KG 43). This work confirms that the current recommendation in NUREG/CR-
6909 [3] to use an average temperature for the analysis of non-isothermal transients may lead to
over conservatism. However, the Fen expression in NUREG/CR 6909 also takes into account
strain rate and, to take full benefit from TMF testing, it is necessary to gain a better
understanding of the complex waveform shapes associated with thermal transients and their
impact on fatigue life. Furthermore, there is a definite need to understand the impact of TMF
testing on crack nucleation, which would require testing closer to the high cycle regime.
However, the significant testing times involved in EAF experiments in this regime could be a
significant obstacle.
Knowledge Gap 16
There have been some facilities commissioned in the past five years that directly investigate the
implications of thermal shocking of stainless steel. Two of the facilities use hollow type
specimens [88; 89] and the other uses a mix of notched ring specimens and cruciform specimens
[90]. Each facility has a slightly different aim not just with respect to crack growth or initiation
but also what aspect of plant-realistic loading is incorporated into the tests. As a result, the
specimens used in these tests as well as the facilities themselves, are very different, and trying to
compare results is limited to qualitative indications.
The facility described by Platts et al. [88] focusses on assessing the impact of through-wall strain
gradients on EAF using fatigue initiation testing. The specimens used in these tests are thick
walled hollow specimens (30 mm diameter, 6 mm bore, 60 mm length) in which a simulated
LWR coolant is continuously flowed through the inner bore. The test facility produces a thermal
shock by varying the coolant temperature between, typically, 300 °C and 60 °C. The total cycle
time, allowing for the specimen to thermally stabilize at the target temperature, was 10 minutes
in testing reported to date. FEA was used to optimize the specimen geometry and determine both
the magnitude and position of highest strain within the specimen according to the intended
thermal shock profile. For the defined thermal shock and specimen design a strain amplitude of
0.37 % was calculated using the R5 methodology and 0.7 % when using ASME (Ke)
methodology. Based on data presented by Stairmand et al. [15], the life of the specimen was
predicted to be 5000 cycles and 1400 cycles for the strain amplitude calculated by the R5 and the
ASME methods, respectively. At the time of publication, the experiment had been run for 3000
cycles without any detectible cracking. It was concluded that this observation demonstrated the
conservatism inherent in the ASME methodology. Further testing would be required before
conclusions could be made on the benefit that could be obtained by accounting for strain
gradients and revised effective temperatures.

3-30
13321451
LWR Environmental Fatigue

Niffenegger et al. [90] used blunt notch ring and cruciform specimens to investigate the
implications of thermal shocking on crack growth and initiation respectively. The notched ring
specimens (70 mm diameter, 25 mm bore, 5 mm thick) were thermally shocked by passing hot or
cold oil through the inner bore of the ring specimen. Two different sequences with different
loading patterns were used: a unilateral shock (hot followed by cold) and a bilateral shock (cold
– hot – cold). The time period of the shocks was 10 seconds. A set of standard fatigue
experiments was also done to establish the cyclical behavior of the material for use in FEA
calculations. The crack length was determined during the experiments by imaging the specimens
using scanning electron microscopy after block loading of a number of cycles. For the two strain
amplitudes studied (0.0316 % and 0.0175 %) the calculated number of cycles to initiation was
significantly less than the number of cycles observed experimentally. This indicates that the
assessment methods used in the calculation of crack initiation in these specimens were very
conservative.
The cruciform specimens, used in the work of Niffenegger et al. [90], were optimized by FEA to
give a well-defined stress state in the center of the specimen. The specimens could be loaded in
the rig at the same time as being thermally cycled, in order to better simulate plant conditions.
The loading on the specimen, if used, could be controlled based on force/displacement
measurements provided by strain gauges, piezo transducers and inductive distance sensors. The
authors stated that most of the tests were performed using load control to account for thermal
expansion. The thermal shocks were provided by spraying water vapor (460 °C) followed by
cold water onto the surface of the specimen with a cycle period of 1 second. This procedure
produced a ΔT of 200 °C and, after FEA was performed, it was found to correspond to a strain
amplitude of 0.12 %. Initial experiments on polished specimens found no cracking within the
timescale of the experiment and there was no observable network of cracking observed at
200,000 cycles. These results are not surprising considering that the fatigue limit of stainless
steels provided in NUREG/CR-6909 [3] is very close to the strain amplitudes used in these
experiments and some heats of stainless steel have been reported to have fatigue limits
significantly higher than this. Although crack initiation was not expected, or observed, for
smooth specimens, specimens finished by sand blasting developed cracking after 200,000 cycles.
The sand blasted specimen would have had a significantly rougher finish which is known to
shorten fatigue initiation lifetimes in air.
KG 16 also refers to the need for thermal cycling testing to produce an enhanced understanding
of the mechanisms behind EAF and ultimately a method of validating and improving assessment
methodologies. There are multiple examples of test facilities attempting to address this issue, the
most well-known being the Bettis stepped pipe experiment. These experiments can be extremely
complicated with strains that are hard to define and often very difficult to analyze accurately.
Testing incorporating plant-representative features is dealt with further under KG 33 (Section
3.5.1).
Summary
KGs 15 and 16 both relate to aspects of how thermal transients impact the fatigue life of
components and how best to replicate this in a laboratory setting. KG 15 deals with the effect of
varying the parameters of strain rate and temperature through a cycle on fatigue life. Work on
this gap has focused on thermo-mechanical testing, which attempts to better define the correct
temperature to use in the analysis methodologies. Recent work highlights the conservatisms in

3-31
13321451
LWR Environmental Fatigue

the methods proposed by ASME and NUREG/CR-6909, demonstrating that out-of-phase


thermo-mechanical testing produces fatigue lives closer to isothermal tests conducted at the
minimum temperature than the maximum or mean temperatures. In terms of variable strain rate,
isothermal testing simulating stresses induced by thermal cycling clearly shows the conservatism
in the NUREG/CR-6909 methodology even without taking into account the fact that maximum
strain often occurs in practice close to minimum temperature, which would further reduce the
environmental effect on fatigue. Damage models based on plastic strain have been proposed to
incorporate the observation that fatigue damage is not constant throughout a complex load cycle.
Following on from this work, models (including the S-N weighted model) have been proposed
which can account for both TMF and variable strain rate testing more accurately than the
methodology proposed by NUREG/CR-6909 (see KG 43 for further discussion). Nevertheless,
the models have not been fully validated, and further test data under plant-representative loading
conditions are required. No similar testing has been reported under BWR conditions nor for other
materials such as carbon or low alloy steels or nickel-based alloys.
KG 16 deals specifically with features and strain gradient testing. Although significant activity
has occurred towards addressing this KG, most of the facilities commissioned to address this
issue are in the very early stages of data generation. As a result, the work reported in support of
this gap typically describes the test facility and outlines preliminary data produced. Nevertheless,
the available data identify further conservatisms in the existing assessment methodologies and
indicate the potential benefits that could be gained by further investigation towards
understanding thermal strain gradients.
Since the identification of both these KGs relating to thermal transient testing, a significant effort
has been undertaken to commission new test facilities and generate new data under non-
isothermal conditions, as well as variable strain rate data using more conventional isothermal
testing but with complex waveforms. The data available to date all suggest that addressing these
gaps could result in significant benefits to designers and in plant life extension. These benefits
would be in terms of optimizing conservatism in assessments of fatigue in components exposed
to LWR environments under plant transient loading conditions. The testing work currently
available for both gaps is insufficient to resolve the issues completely. In light of the work done
towards KGs 15 and 16, it is clear that further work is required to close them and provide the
basis for development of alternate assessment methodologies (see KGs 39, 41 and 42 in Section
5.3.2). Data on this subject for other alloys such as ferritic steels and nickel based alloys are
much more limited than for stainless steel. Therefore, KG15 and KG16 should remain open.

3.5 Prototypical and Component Testing

3.5.1 Knowledge Gap 33 (Previously Classified as Category A, High Priority)


Knowledge Gap Definition
More data using component-like features with plant-representative loading conditions are
required to develop and validate methods for considering corrosion fatigue in LWR
environments.

3-32
13321451
LWR Environmental Fatigue

Research & Development Need


Analysis to identify conditions needed for testing to simulate plant conditions. Focused testing
on geometries representative of components under plant-relevant transient conditions. Testing of
standard specimens under plant-relevant loading noted above also supports this need. Testing
should accurately simulate plant water chemistry, including transient chemistry where
appropriate.
Ancillary Information
Chapter Section Subsection(s) Page(s) Applicable Materials
9 Critical Review of Similitude with Feature 9-13 to 9-14 All
Design Code Tests and Plant
Developments Operating Experience

Background
The discrepancies between plant experience and laboratory data are a key driver for a
considerable number of KGs and EAF research, but very few tests have been performed directly
investigating plant-relevant features and geometries. Examples of these tests include the Bettis
thermally cycled stepped pipe [91] and Areva’s mechanically-loaded U-bend pipe [92]. The
results of these studies were discussed in the KG Analysis report [6], and it was suggested that
significant benefits could be gained from additional testing of this type.
Review of Activities Relevant to KG 33
Experiments on full-size stainless steel pipe in air have been conducted by Arora et al. [93] with
the aim of understanding the fatigue crack growth behavior using loading cycles which included
overloads, underloads, as well as variable-amplitude loading. In this work, Type 304LN stainless
steel straight pipe (OD 169 mm, 14.4 mm thick) was tested where the cyclical loads were applied
using a 4-point bend arrangement. To fully understand the fatigue crack growth behavior, the
Paris law constants were determined using compact tension type specimens in standard fatigue
crack growth tests. It was observed that the application of an overload or underload caused a
short term crack growth acceleration. Furthermore, crack growth retardation was found to occur
after the pipe was subject to an overload but not in the case of underloads.
Selvam [94] conducted tests investigating the thermal fatigue caused by hot and cold flow
mixing at a tee-junction. Thermal fatigue failures at mixing tees are often unexpected due to the
fact that a description of the thermal loads cannot be adequately built up from thermocouples.
The aim of the work by Selvam was to assess the suitability of large eddy simulations as a means
of predicting the thermal fatigue that can occur in this location due to the turbulent mixing of two
hot and cold streams of coolant that produced temperature fluctuations with frequencies in the
0.1 to 10 Hz range. The test facility consisted of a main (diameter 71.8 mm) and branch
(diameter 38.9 mm) loop connected to a water supply and pressurized to 75 bar. The main loop
was heated using ceramic pads so that a ΔT of 260 °C could be achieved when the coolant from
the main loop and branch were mixed. To monitor the temperature, thermocouples were arranged
around the diameter of the test section at 45° intervals. The experimental work was used to
validate the large eddy simulation of the thermal fluctuations resulting from the turbulent mixing
of the two coolants. Several simulations were run to determine the effect of flow rate on the
mixing behavior in the tee. These simulations demonstrated that the mean temperature in the

3-33
13321451
LWR Environmental Fatigue

mixing regions increased with increasing flow rates, near-wall temperature fluctuation
amplitudes are highest near the stratification layer and that there was no single dominant
temperature fluctuation in the 0.1 to 10 Hz range. Ultimately, the large eddy simulations were
found to be a good match to the experimental data.
Schuler et al. [95] performed thermo-mechanical testing on a full scale welded pipe in a high-
temperature water environment. The welded pipe was of a dissimilar metal weld between ferritic
and austenitic steel with a nickel-based alloy weld filler. The objective of the work was to
generate further understanding of the fatigue properties of the nickel-based alloy weld filler in a
LWR environment. This was done in an oxygenated BWR chemistry environment at pressures of
up to 75 bar. The thermal cycling, up to 100 cycles, was performed by injecting room
temperature water into the 280 °C water in the pipe segment before the welded pipe. This was
done by the use of valves which could inject 200 gs-1 of cold water into the hot branch that had a
flow of 560 gs-1 and resulted in combination of thermal shocks and temperature fluctuations in
the welded pipe. Schuler et al. found that the largest temperature fluctuations were between
150°C and 170°C in magnitude and occurred at 270 °C. These temperature fluctuations were
monitored using sets of thermocouples that were placed both in the fluid and within the metal
pipe. The thermocouples in the fluid were placed 1 mm from the inner surface of the pipe, and
the thermocouples inside the metal pipe were inserted between 1-1.5 mm from the inner surface.
Numerical simulations were carried out to determine which regions of the pipe were subject to
the greatest loading and the likely position for the largest cracks. The simulations indicated that
the largest cracks were expected to be in the transition region between the nickel-based alloy
weld material and the austenitic steel. Using this information, Schuler et al. focused their
inspection of the welded pipe specimen on this region and found that the fluctuations created in
the pipe by the injection of the cold water created microcracks up to 30 µm in length after the
100 cycles were completed. The numerical simulations and supporting uniaxial fatigue testing
suggested that any cracks found in the pipe were not expected to have reached an engineering
size after 100 cycles.
A reanalysis of the stepped pipe experiment [91] was performed by Mann et al. [57] and was
based on a total life approach. The methodology was developed from a striation counting study
that determined the crack growth behavior of fatigue endurance specimens in air and LWR
environments beyond 250 µm crack growth and up to 3 mm. It used this work to define a best-fit
curve representing initiation (including crack growth to 250 µm in this case) such that whole
fatigue lives could be predicted. This work found that the stepped pipe would be expected to
develop a 3 mm crack, the ASME definition of initiation, after 862 cycles. This is considerably
higher than the values for initiation (3 mm crack) calculated for this same experiment by Gurdal
et al. [96] who predicted lives of 30 and 39 cycles to failure using the NUREG/CR-6909
procedures based on Fen factors of 4.23 and 4.88, respectively. The total life approach used by
Mann et al. to reanalyze the stepped pipe tests indicates that the current Fen factor approach to
analysis may be conservative. The work suggests that using methods that take advantage of a
better understanding of the underlying mechanisms, FEA analysis methods and a more focused
treatment of each stage of crack growth would be beneficial.
Steininger et al. have identified a need for a full-scale component test to investigate the apparent
conservatism in current assessment methodologies predicting fatigue life in LWR environments
and have outlined the contents of an EPRI Request for Proposal (RFP) for such work [97]. The
basis of the requirement is that, in contrast to the predictions of current assessment

3-34
13321451
LWR Environmental Fatigue

methodologies, there is no operational experience of an ongoing EAF problem in either U.S.


domestic LWR plants or in international plants. Therefore, it is considered that the current
requirements are not optimized for reasonable, safety operation of nuclear power plant
components. The stated aims of the proposed work are to better quantify margins in current
fatigue assessment methods and to provide benchmark data for potential new assessment
approaches. It is planned that the test results would be compared with the results of predictions
using a range of different assessment methods by means of an international round-robin exercise.
It is observed that whilst current small scale specimen tests, even using complex and non-
isothermal loading, only study one effect or the interactions of a small number of effects, plant
components experience a wider range of conditions over plant life. The need for scaling of
observations to larger size components, inclusion of constraint features and thermal shock
loading representative of plant components was also highlighted. The proposal outlines some
examples of component design and test conditions. For example, a simulated valve body
(potentially with internal features such as steps or notches) could be subjected to thermal shocks
or blocks of temperature cycling to investigate variable-amplitude loading. Another option
presented was a pipe section welded to a nozzle that could be thermally shocked or have external
bending loads applied to it. The authors highlighted the need for the test to be as relevant to plant
conditions as possible to avoid uncertainty in the applicability of the results. However, a smaller
scale test could be acceptable if its applicability to plant conditions could be adequately
demonstrated. It was stated that the RFP would not contain constraining details which could limit
the ability of potential bidders to produce innovative design options for the test. Initially the
focus of the proposed testing would be on austenitic stainless steel in PWR primary coolant with
a possibility of subsequent extension to BWR conditions.
A further requirement was that the contractor should provide analytical results to support the
proposed experimental designs and that adequate monitoring to determine the stresses/strains
applicable to the test piece and the length of any cracks was an important consideration. Both
online and offline techniques were suggested in the paper. However, the periodic shutdowns
required for offline monitoring techniques were highlighted as a concern. Furthermore, the paper
stated the necessity of post-test examination of the specimen using SEM or other methods to
characterize any cracks that had initiated. Finally, Steininger et al. recognized that this type of
work was expensive and would require international backing if it was to be widely accepted and,
therefore, stated that EPRI was actively looking for co-partners to join them in co-funding this
activity.
Summary
There has been only limited additional work involving large-scale testing aimed towards
understanding the discrepancies between laboratory data and plant experience since the KG
Analysis was issued [6]. Work by Arora et al. has provided evidence that underloads and
overloads can cause similar behavior in large scale test pieces to that observed in CT specimens.
Sevlam and Schuler et al. have produced work that has provided confidence in theoretical
predictions of fatigue behavior of large-scale tests under thermal loading. Analytical work has
been done to develop new methodologies to reanalyze existing data generated from the stepped
pipe experiment. Further data are required to validate assessment procedures and understand the

3-35
13321451
LWR Environmental Fatigue

difference between plant experience and laboratory data. Considering the limited amount of data
currently available for geometries and pipe sizes representative of plant components under plant-
relevant transient conditions KG 33 remains open.

3.5.2 Knowledge Gap 36 (Previously Classified as Category C, High Priority)


Knowledge Gap Definition
The basis for the selection of effective stress parameters for biaxial stress conditions is not
established. Test data are required under conditions of biaxial loading for the treatment of plant
thermal transients and non-proportional loading for combined thermal and mechanical transients.
The most appropriate parameter may be different for the crack nucleation and subsequent
propagation of microstructurally small cracks.
Research & Development Need
Test data are required to identify the appropriate multiaxial stress parameter for the treatment of
biaxial condition in corrosion fatigue assessments.
Ancillary Information
Chapter Section Subsection(s) Page(s) Applicable Materials
9 Critical Review of The Effective Stress 9-11 All
Design Code Parameter –
Developments Multiaxial Loading

Background
Thermal transients represent a significant proportion of fatigue usage for nuclear plants and,
when coupled with complex geometries, can induce multiaxial and non-proportional loads onto
components that may affect their fatigue life. Currently, both ASME III and NUREG/CR-6909
use the Tresca yield criterion, which is more conservative than the von Mises equivalent, to
define an effective stress for biaxial loading. However, the evidence supporting this position for
EAF is lacking. The most appropriate method for defining the effective stress for EAF will be
dependent upon the mechanisms involved as well as between initiation and crack growth and,
therefore, could potentially be different for each material and environment.
Review of Activities Relevant to KG 36
Currently there is a lack of data on fatigue with respect to multiaxial structural loading. CEA has
commissioned a set of test rigs capable of generating fatigue strength data on austenitic stainless
steel under biaxial loading conditions which are free of the influence of mean stress; results are
reported by le Roux et al. in air [98] and in PWR conditions [99]. Results for Types 304L and
316 material have revealed that, when the load ratio is -1 and the results are rationalized as a von
Mises equivalent strain, biaxial loading has minimal impact on fatigue life relative to
conventional uniaxial fatigue testing. This work in air was further supported by Kamaya et al.
[100] who showed that crack growth under biaxial conditions, when considered in terms of
equivalent von Mises strain, was close to that found for uniaxial test specimens.
Work on carbon steel regarding non-proportional loadings was recently performed by Arora et
al. [101], who discuss the impact of torsion on fatigue life using in-phase and out-of-phase axial
loading. This work was done on solid specimens for axial and hollow specimens for the shear

3-36
13321451
LWR Environmental Fatigue

and non-proportional loading tests. There were also a limited number of pure axial loading tests
done on hollow specimens. Tests done in air using pure axial strain (on both solid and hollow
tubular specimens) showed no significant difference in hardening for both specimen types, and
the fatigue data were within the 95 % scatter band of the ASME mean air curve for carbon steel.
One interesting observation was that the tubular specimens tended to produce fatigue lives at the
lower edge of the data scatter for the solid specimens, which possibly indicated a geometric
impact on fatigue life. The use of hollow specimens in the shear and non-proportional loading
tests, as opposed to solid specimens, was to minimize the shear stress gradient effect in the
thickness direction. Pure torsion (shear) tests showed a slightly lower level of hardening in an
equivalent strain (von Mises) vs. equivalent stress plot and consistently longer fatigue lives when
compared to pure axial strain tests. Combined torsion and axial strain tests using triangular,
sinusoidal or trapezoidal waveforms resulted in higher levels of hardening and fatigue lives that
were over a factor of 2 shorter than the pure axial tests. The tests using sinusoidal and trapezoidal
waveforms gave slightly greater hardening and shorter lives at the same stress amplitude as the
triangular waveforms.
The conclusion drawn from these tests was that the ASME procedure may over predict the
fatigue lives for non-proportional axial-torsion loading conditions. This work proposed a
“Modified Chu” model which was based on the Biaxial-Smith-Watson-Topper (BSWT) model
[102] and used a fatigue damage parameter to describe the fatigue life of carbon steel specimens
under non proportional loading. However, unlike the BSWT model, the Modified Chu model has
a calibration factor (k) to weight the relative contributions from shear and axial stresses during
non-proportional loading. Furthermore, it includes a method for evaluating shear stress/strain
amplitudes which, under non proportional loading, can have varying magnitudes and directions.
The model was used to predict the lives of specimens subjected to axial, pure shear and non-
proportional tests. Using the axial and shear tests, the value of k was determined to be 0.4 and
was subsequently used in the analysis of the non-proportional test data. Predictions of fatigue
lives using the model on carbon steel specimens subjected to non-proportional loading were
within a factor of 2 of the observed lives which is considered good.
Arora [101] cited work by Liu [103] which studied specimens from non-proportional tests on
aluminum using TEM. This study investigated the dislocation substructures generated during
non-proportional loading. This work found that the activation of multiple slip systems caused
gliding dislocations on different slip planes to interact forming cross dislocation bands that
further restricted dislocation movements compared to axial testing. This mechanism was
proposed to explain the additional hardening and reduction in fatigue life in the axial-torsion
tests described above when compared to standard uniaxial tests. This mechanism fits in well with
the pure torsion test results which produced similar hardening results but slightly longer fatigue
lives than the pure axial tests. Since axial-torsion tests affect the cyclical hardening and softening
of a material that occurs early in fatigue life, non-proportional loading could have a measurable
impact on fatigue initiation.
Summary
Work on multiaxial fatigue is ongoing with activities investigating biaxial testing in air and LWR
conditions in addition to non-proportional and torsional testing. Biaxial loading, when
rationalized in terms of Von Mises equivalent strain, has been shown by multiple authors to have
little additional impact on the fatigue life of stainless steels. Some work has also been done on

3-37
13321451
LWR Environmental Fatigue

carbon steels to investigate the impact of torsional and non-proportional loading on fatigue life.
It was found that the ASME procedure over-predicts the life of specimens subjected to non-
proportional loads. However, a modified approach was found to more accurately predict the
impact of non-proportional loads. Although work has been ongoing towards generating an
understanding of multiaxial loading there is still a considerable amount of work left to do to
address this gap. For example, further work on both stainless and ferritic steels is required to
understand the response of each material to non-proportional and biaxial loading. For these
reasons KG 36 remains open.

3.6 Summary of LWR Environmental Fatigue


The KGs described in this section relate to five of the seven hypotheses presented in the
Roadmap report [7] and cover high, medium and low priorities. The hypotheses relevant to the
Gaps contained in this section are 1, 2, 4, 5 and 6. According to option 2 of the Roadmap
document recommendations, hypotheses 1 and 6 are the most significant of the five relevant to
the Gaps in this section. This section will summarize the advancements made towards the high-
priority gaps as defined in the Roadmap report.
KGs 15 and 33 are classed as high-priority Gaps for hypothesis 1. KG 15 focuses on non-
isothermal loading and waveforms with variable strain rates whilst KG 33 is concerned with
component-like features and plant-relevant testing. Of the two Gaps it is unsurprising to find
that, given the complexity and cost involved in testing component-like features, KG 15 has
received much more development than KG 33. The Budapest University of Technology and
Economics has built experimental apparatus for testing the TMF properties of ferritic and
austenitic steels in air by cyclically-loading a specimen and applying an in-phase thermal cycle.
These tests initially produced fatigue lives that were less than the comparable isothermal tests
done in the work. However, the use of a single nozzle to spray cold air onto the specimen gauge
length could produce a bending moment and an uneven thermal profile across the specimen. The
presence of an uneven temperature profile and potential bending during a fatigue test could
account for the lower lives recorded for the in-phase TMF tests. PSI, MHI and Amec Foster
Wheeler have all developed TMF facilities based on hollow specimens that are at varying levels
of maturity and can test in LWR conditions. These facilities have produced data that show the
conservatisms of current predictions for out-of-phase TMF loading. These data have been used to
create new assessment methodologies that can better predict the fatigue lives of test specimens
than the procedures in NUREG/CR-6909. However, more data—including variable strain rate
waveforms with thermal cycling—are required to fully validate the new methods. There is a
significant benefit to be gained from revising current assessment methodologies through proper
understanding of the impact of TMF testing on the fatigue lives of small-scale specimens.
There has been some work on component-like features testing by Arora, Sevlam and Schuler,
which has investigated the impact of variable-amplitude loading (overloads and underloads),
temperature fluctuations in a mixing tee and thermal cycles on a section of welded pipe. The
studies by these researchers have typically been used to validate numerical methods or to get
further understanding of existing theories. Due to the complexity of component-like features
testing the studies relating to KG 33 are not directly comparable to each other. This is likely to
be the case for future features testing, which will be driven by a specific industry need and could
potentially be unique, and could hinder the progress of KG 33.

3-38
13321451
LWR Environmental Fatigue

KGs 7, 16, 17 and 36 are considered high-priority Gaps for hypothesis 6 of the Roadmap
document. KG 7 relates to current mechanistic understanding indicating that Fen should be
dependent on strain amplitude; KG 16 focuses on complex non-isothermal transients and current
averaging procedures; KG 17 revolves around the data required to substantiate the modified rate
approach in NUREG/CR-6909; and KG 36 highlights the lack of understanding of what the
impact of biaxial/ non-proportional loading has on fatigue life. All of these Gaps relate to either
aspect of calculating the effect of environment on fatigue life, aspects of loading that are relevant
to plant that are not fully considered within current methods or both.
Some progress has been made towards KG 7 and understanding the relationship between
environmental effects and strain amplitude. Most of the work done in relation to this Gap
indicates that Fen is dependent on strain rate, although some of the studies disagree. It appears
that an inconsistency in the definition of when the transition between stage 1 and stage 2 crack
growth occurs may be responsible for different authors reaching conflicting conclusions.
Biaxial and non-proportional testing (KG 36) have had some investment and are currently being
developed. New test facilities are being made by CEA to investigate biaxial loading under PWR
conditions that has shown that, when biaxial loading is considered using an effective von Mises
strain, there is no real impact on fatigue life. Arora has begun testing on biaxial and non-
proportional loading in air to develop a new model (called the “Modified Chu” model that is
based on the Biaxial Smith-Watson-Topper model) which showed good agreement between the
predicted and experimental lives of the specimens.
KGs 16 and 17 both deal with the lack of data supporting the way in which current
methodologies handle plant-realistic thermal transients and variable strain rates. The issues
identified by KG 16 have received a lot of effort recently with several organizations
commissioning test facilities capable of delivering thermal shocks to various unique specimen
designs. Due to these facilities producing data relatively recently there has only been a small
amount of data on which to make conclusions. The initial test data from these organizations
typically highlight the inherent conservatisms in current methodologies that do not specifically
consider stress/strain gradients or other aspects of complex thermal cycles. The initial results
show that putting further effort into this area is likely to produce a good return. KG 17 has
received little work and, as a result, is still extant as it is generally understood that the modified
rate approach will be very conservative when considering complex waveforms or transients due
to fatigue damage being unevenly distributed throughout the cycle. However, this Gap may be
better addressed within KG 15 or 18, which deal with complex transient loading data as, during
the development of new methods of handling complex transients, comparisons will be made with
respect to the modified rate approach. During this process, evidence and justification for and
against the modified strain rate approach will be presented.
Hypotheses 4 and 5 each have one high-priority Gap discussed within this section. Although
these hypotheses were considered not as significant as 1 and 6 in the Roadmap report, KG 18
(which relates to hypothesis 4) has received considerable attention. On the face of it, KG 18
deals with non-contiguous cycles and its research and development need states that complex
transient loading data are required and to be compared with current assessment methods. This
means that it is arguable that variable strain rate testing falls within the remit of KG 18 rather
than KG 15 (which also passively mentions it). Therefore, KG 18—as discussed in this section—
contains the data on variable strain rate testing and hold times. In terms of variable strain rate

3-39
13321451
LWR Environmental Fatigue

testing, considerable effort has been put into investigating the impact of complex waveforms on
fatigue life and comparing that information to predicted Fens. This has shown that current
assessment methods do not adequately handle variable strain rates. Testing involving hold times
is an attempt to investigate the impact of steady-state operation of a plant and, to some extent,
the idea that the peak stresses which are chosen to perform fatigue analysis could be separated by
a significant length of time. The data from tests using holds at high-temperature indicate that
there may be a beneficial effect of holds on fatigue life at strain amplitudes of 0.4% s-1 and
below. This work has resulted in some progress towards a mechanistic understanding of this
effect and appears to show that a thermally activated process is responsible for the life extension.
KG 13 is the high-priority Gap considered within this section for hypothesis 5. In this section,
KGs 10 and 13 are dealt with together as they correspond to the definition of the Fen equations
and their thresholds. After the publication of the KG Analysis in 2011, NUREG/CR-6909 Rev 1
was issued, initially as a draft for comment in 2014, and very recently in final form [3]. This new
revision altered the Fen equations and their thresholds based on an expanded database of fatigue
data and in response to questions from the nuclear industry. Further work on stainless steel, by
Areva appears to cast some doubt on the method used in NUREG/CR-6909 for calculating Fen
which multiplies the strain rate, dissolved oxygen and temperature parameters together,
suggesting instead that a more complex procedure would be preferred. However, as evidenced in
NUREG/CR-6909 Rev 1 [3], the amount of data available to develop and validate a more
complex procedure which could account for interdependence between the subfactors is
insufficient.
Interestingly, KGs 11 and 19, which were rated as low and medium priority, respectively, were
given as much or more attention than some of the high-priority Gaps. KG 11 is a hypothesis 2
Gap that looks at the effect of mean stress on EAF. Much of the work here attempts to address
whether or not mean stress has a positive or negative impact on fatigue life in air and high-
temperature water. The findings of the data are somewhat unclear due to the complexity of
testing mean stress and the use of different specimen geometries. This represents an emerging
issue in EAF testing where there is mounting evidence that hollow specimens are not directly
comparable to standard solid specimens. This observation could have significant implications for
test programs which need to compare LWR data to a mean air data curve generated using
standard solid specimens. Several contributing causes have been discussed but the addition of a
mean stress due to an internal pressure load in hollow specimens appears dominant and methods
of correcting for the differences have been proposed. The database used by NUREG/CR-6909
contains a large amount of hollow specimen data that could potentially be adding conservatism
to the methodology by biasing the mean LWR line to lower numbers of cycles to failure.
NUREG/CR 6909 Rev 1 [3] states in Appendix F that, although the fatigue lives of hollow
specimens were found to be less than those of solid specimens, they did not significantly alter the
end result. Based on this, the method described in NUREG/CR-6909 Rev 1 is founded on both
hollow and solid data.
KG 19 is a hypothesis 2 Gap which deals with the theory that the surface finish transference
factor used by ASME Section III in the creation of a design line does not have the same
magnitude in air and water. This Gap has been viewed as “low hanging fruit” where there
appears to be a significant benefit that can be gained through an obvious testing route. As such
this Gap has seen significant effort from the British, French and Japanese national programs.
Despite this effort and the publication of several new methods taking advantage of these

3-40
13321451
LWR Environmental Fatigue

observations, there still remains much work to do before this Gap can be closed out. This work
needs to resolve the observed reported differences in the magnitude of the reduction in fatigue
life due to surface finish. The situation is complicated by the different methods of obtaining and
characterizing surface finish in different laboratories.
Overall, the KGs within this section have been given significant thought and effort by the
international community. On the whole, the effort has generally been focused on the high-
priority Gaps, as identified in the Roadmap document, but has not specifically focused on
hypotheses that were given the most significance by the EAF expert panel. Additionally, some
significant effort has been given to lower priority Gaps due to emergent or easily testable issues
that could have significant value to designers.

3-41
13321451
13321451
4
ENVIRONMENTAL FATIGUE CRACK GROWTH

4.1 Introduction
This section is intended to address all KGs that relate specifically to the generation and collation
of fatigue crack growth data on a range of LWR structural materials (stainless steels, ferritic
steels and nickel-based alloys). Data in air and also in PWR and BWR environments are
included. Studies which address specific aspects of crack growth behavior, for example, stress
intensity factor (ΔK) thresholds (in air and water), conditions leading to retardation of enhanced
growth and crack growth under compressive load (negative R ratio) are also considered. This
section will also discuss the development of crack growth reference curves, such as those in
ASME Section XI and related Code Cases.
Fatigue crack growth data are most commonly generated using notched and pre-cracked compact
tension (CT) specimens. If data are required under compressive load (negative R),
double-edge-notched or square-corner-cracked specimens have been used. The specimens may
be tested as notched or pre-cracked (with a mechanically large crack) to study either initiation
and short crack growth or long crack growth, respectively. Tests are typically conducted in load
control where, due to the high stress and strain intensification from the notch or pre-crack, a
plastic zone forms at the crack tip. Under these conditions, the bulk of the specimen is in the
elastic state and, therefore, the plastic zone is controlled by the elastic response of the specimen.
Crack growth data are typically presented as a curve showing crack growth per cycle vs. the
linear elastic stress intensity factor. By using a log-log plot the Paris law can be used for
intermediate values of the stress intensity factor. An important aspect of fatigue crack growth
data plotted in this manner is that, prior to establishing Paris law crack growth, there is a
threshold for small stress intensity factors below which crack growth is retarded. At this
threshold stress intensity factor, the crack growth rate is taken to be zero.
To fully understand the implications of corrosion fatigue crack growth data it is typical for data
collected in a particular environment to be related back to crack growth data generated in air.
This can be done using either the fatigue crack growth curves contained within Appendix A of
ASME Section XI, which deals with the acceptability of flaws detected by in-service inspection,
or by producing materials-specific reference data. ASME Section XI contains crack growth
curves for ferritic steels, austenitic steels (wrought, cast and weld metals) and nickel-based Alloy
600 in air. It also gives reference curves for wetted flaws in ferritic steels and nickel-based alloys
in BWR and PWR environments. It does not, however, contain reference curves for austenitic
stainless steel wetted flaws, and the curves for ferritic steels are based on very old data which do
not adequately reflect the frequency dependence of crack growth. ASME Code Case N-809 [5]
was developed to address the deficiency for stainless steel in PWR environments, whilst ASME
Code Case N-643-2 provides more appropriate curves for ferritic steels, again in PWR
environments only. Relevant reference curves for BWR conditions are unavailable.

4-1
13321451
Environmental Fatigue Crack Growth

The ASME Section XI reference curves and Code Cases for high-temperature water
environments are based on data obtained using continuous cyclic loading (triangular, sawtooth
and, occasionally, sine wave) which do not reflect plant transient loading. This was highlighted
as a significant KG in the EPRI review and a significant proportion of recent work has aimed at
addressing this omission.

4.2 Conventional Cyclic Loading Fatigue Testing

4.2.1 Knowledge Gap 22 (Previously Classified as Category C, Medium Priority)


Knowledge Gap Definition
The lack of relevant environmental crack growth data for some materials (e.g., Alloy 690 and its
weld metals* 1) or grades of material (e.g., Types 316L(N) or 347 stainless steel) represents a
KG. Heat-to-heat variability also appears to be important, especially the influence of sulfur for
stainless steel, but is not adequately understood. There is also a lack of threshold ΔK data for
many materials.
Research & Development Need
Further corrosion fatigue crack growth data are required, especially under near threshold
conditions.
Ancillary Information
Chapter Section Subsection(s) Page(s) Applicable Materials
9 Critical Review of Proposed Revisions of ASME 9-17 to All*
Design Code Boiler and Pressure Vessel 9-19
Developments Design Codes/Section XI

Background
The KG Analysis report [6] highlighted a lack of environmental crack growth data supporting the
fatigue crack growth curves and ΔK thresholds for some grades of austenitic stainless steel (and
also stabilized grades such as Types 347 and 321). ASME Section XI and Code Case N-809 [5]
provide fatigue crack growth curves for dry and wetted flaws respectively; although, the latter
refers only to PWR environments. Code Case N-809 gives separate curves for low carbon (Types
304L, 316L) and high carbon (Types 304, 316) grades of stainless steel (and their associated
weld metals). The crack growth rates calculated by Code Case N-809 are a power law function
of rise time, with an exponent of 0.3 and are also influenced by temperature and R ratio [104].
The ΔK threshold value included in the Code Case is just 1.1 MPa√m which leads to
environmental enhancement factors of 80x the ASME air rates for very long (>20 h) rise times.
The report recommended generation of further environmental crack growth data, especially
under near threshold conditions, for grades of stainless steel not included in the N-809 database.

1
Data for other materials for which data are lacking, such as nickel-based alloys, was also included in the original
version of this KG, but are covered specifically by KGs 26 and 27, so they are discussed within those gaps.

4-2
13321451
Environmental Fatigue Crack Growth

Review of Activities Relevant to KG 22


Long Crack Growth
A number of laboratories have continued to carry out corrosion fatigue crack growth testing on
austenitic stainless steels in PWR environments. In conventional triangular or sawtooth loading
tests, there are only relatively small differences in behavior between different heats of stainless
steel, except when retarded rates occur as the loading frequency is progressively reduced (see
KGs 28 and 31). To gain insight into the impact of LWR conditions on crack growth, the data
collected in environmental tests are compared to a base line which defines the behavior of
stainless steel in air. The air fatigue crack growth laws presented in ASME Section XI, JSME or
material-specific reference curves are typically used for this purpose [105; 106].
In contrast to PWR conditions, comparatively fewer studies have been conducted under BWR
conditions. The work of Seifert et al. [106; 107] compared the effect of simulated PWR and
BWR (both NWC and HWC) conditions on fatigue crack growth for a range of stainless steels
including Types 304, 304L, 316L and the stabilized grades, Types 321 and 347. This work
established that, in all three environments, significant environmental acceleration occurred
between loading frequencies of 0.1 Hz and 3 x 10-6 Hz, with the degree of environmental
enhancement relative to air increasing with reducing loading frequency. In general, Seifert found
that there were no significant differences in growth rates between the grades of 300 series
stainless steel. However, in low electrode potential environments (both PWR and BWR HWC),
the high sulfur material (0.025 %) studied in this work showed rates that were up to factor of two
smaller than their low sulfur variants [75]. This behavior was shown to increase with decreasing
frequencies or small ΔK and high load ratios. The crack growth retardation could be reversed in
the high sulfur containing materials by a temporary increase in loading frequency or ΔK after a
sufficiently large amount of growth was achieved. Furthermore, Seifert noted that a Type 347
material with a sulfur content of 0.006 % also experienced retardation. In this case retardation
was observed at low loading frequencies, small ΔK and high R ratios, and for tests involving
hold periods. This observation of retarded crack growth in high sulfur steels is consistent with
the work done by Tice et al. [53] and discussed further in KGs 25 and 31. For a given material,
growth rates were somewhat higher in the oxidizing BWR NWC environment than in HWC or
PWR conditions.

ΔK Threshold
Work on defining the ΔK threshold (ΔKTH) and the impact of load history on the threshold itself
for stainless steel in 250 °C air and PWR conditions was performed by West et al. [108]. These
experiments estimated the ΔKTH of Types 304/304L stainless steels as 3 MPa√m (air) and
2 MPa√m (LWR). This results in the crack growth rates in the low ΔK regime predicted by the
ASME N-809 [5] and Mills models [109] (which both use a threshold value of 1.1MPa√m) being
larger than those experimentally observed. Further experimentation using overloads and
underloads showed that, following an underload within the region approaching the threshold, the
crack growth rate immediately after the event accelerated up to a rate consistent with the Mills
model. In contrast to this, tensile overloads suppressed the subsequent crack growth rate as a
result of plastic wake closure effects that raised the apparent ΔKTH. It was argued that, under
plant conditions in which both underloads and overloads occur, these two effects would balance
one another, especially since it was stated that overloads were the more common event. Work
performed by Mills in 2013 [110] also found that in crack growth tests in 250 °C air on Type 304

4-3
13321451
Environmental Fatigue Crack Growth

materials, a ΔKTH of 5 MPa√m was observed which is considerably higher than proposed in
Code Case N-809. Together, these ΔKTH data appear to demonstrate that the degree of
conservatism in the ASME crack growth and Mills model may be excessive in the near threshold
regime for stainless steel. Discussion of threshold effects on crack growth rate predictions under
plant transient conditions appears under KG 23 in the assessment section.
Summary
The main focus of this gap is on extension of the crack growth database for stainless steel to
other grades, such as stabilized materials, and on threshold ΔK data for stainless steel.
Considerable work has been undertaken towards this gap for stainless steel, especially in PWR
conditions, reinforcing the idea that a single fatigue crack growth curve is reasonable for the
majority of heats. Overall, it is not clear that the difference drawn between standard and L grades
of Types 304 and 316 in Code Case N-809 is consistent across all the new data. The behavior of
stainless steel heats appears to be less predictable when considering high sulfur steels under low
electrode potential conditions (PWR and BWR HWC) which show retardation of enhanced crack
growth as the loading frequency is reduced; this issue is covered specifically by KG 25. Work
done in the low ΔK regime demonstrates that the 1.1 MPa√m ΔKTH in Code Case N-809 is likely
to be too low and results in significant conservatism.
Further work is required towards this gap to confirm ΔKTH values for stainless steel over a wider
range of materials and to confirm the extent to which these values may be influenced by loading
history. There is also a need to further explore the differences between high and low electrode
potential environments (see KG 24).
Therefore, KG 22 remains open and should focus on ∆K threshold conditions and
underload/overload effects for austenitic stainless steels in PWR environments. Nickel-based
alloys are covered by KGs 26, 27 and 32 and ferritic steels by KGs 29 and 30. Threshold
behavior for stainless steels in BWR environments could be addressed as part of KG 24 below.

4.2.2 Knowledge Gap 24 (Previously Classified as Category C, Medium Priority)


Knowledge Gap Definition
ASME XI does not include a fatigue crack growth law for wetted flaws. Fewer relevant data are
available for BWR environments although recent data suggest environmental effects are
somewhat greater in BWR NWC than HWC or PWR (this is in contrast to S-N data).
Research & Development Need
Need to evaluate significance of recent BWR NWC and HWC data and ascertain whether further
work is necessary to develop an assessment methodology for BWRs.
Ancillary Information
Chapter Section Subsection(s) Page(s) Applicable Materials
10 Discussion of Laboratory Data/Crack 10-6 to 10-7 Austenitic Stainless Steel
Knowledge Gaps Growth Data

4-4
13321451
Environmental Fatigue Crack Growth

Background
ASME Code Case N-809 [5], which has been published since the previous KG review, provides
crack growth reference curves that reasonably describe the fully enhanced crack growth rate data
produced in a PWR environment. Recent developments and outstanding issues relating to this
Code Case are discussed in KG 23 in the assessment section. However, Code Case N-809 does
not cover BWR environments. The amount of data that has been collected in BWR HWC or
NWC environments is much more limited than in PWR, but HWC data appears to be reasonably
consistent with PWR data. The NWC data show slightly higher rates than in PWR, but seem to
follow the same trends. This is in contrast to what is observed for fatigue endurance data which
indicates that environmental effects are less for stainless steels in oxygenated conditions such as
BWR NWC [8] than for PWR. This underlines the need to evaluate any new BWR data to
determine if a separate assessment methodology is required for this environment.
Review of Activities Relevant to KG 24
When comparing results between BWR NWC and PWR/BWR HWC for an array of 300 series
stainless steels Seifert et al. [106] found the PWR and HWC environments generated comparable
rates to each other and that the NWC environment produced rates which were a factor of 1.5 – 5
higher. Additionally, Seifert stated that the slip-dissolution model reasonably predicted the crack
growth rate under NWC conditions but severely underestimated the crack growth rate observed
in HWC and PWR environments. This suggests that the mechanism of corrosion fatigue may be
different between the two types of environment (high electrode potential BWR NWC and low
electrode potential BWR HWC and PWR). Seifert advocated the use of a single crack growth
curve for PWR and BWR HWC and a separate curve for BWR NWC conditions. The Japanese
approach, which consists of separate curves for BWR (based on NWC BWR data) and PWR
environments, bounds the data produced by Seifert.
Summary
Work towards evaluating the need to use separate curves for PWR and BWR environments has
indicated that PWR and BWR HWC environments give similar crack growth rates and can be
described using a single curve. Tests in BWR NWC environments appear to result in crack
growth rates up to 5 times faster than the corresponding BWR HWC tests. Seifert argues that the
environments could be divided by high and low electrode potential environments and that the
N-809 PWR curves are likely to be suitable to BWR HWC conditions; whereas, separate curves
should be developed for BWR HWC. Further work exploring the differences between the high
and low electrode potential environments should be conducted to identify the underlying
mechanisms to fully support the use of separate curves. Therefore, KG 24 should remain open.

4.2.3 Knowledge Gaps 25 and 31

Knowledge Gap Definitions


Knowledge Gap 25 (Previously Classified as Category C, Low Priority)
The effects of parameters influencing when retardation of enhanced crack growth occurs [in
austenitic stainless steels in a PWR environment] are not adequately understood. The possibility
of crack growth rate retardation is not evident in the available data for BWR environments.

4-5
13321451
Environmental Fatigue Crack Growth

Knowledge Gap 31 (Previously Classified as Category D, High Priority)


The mechanism of environmental enhancement is not well understood.
1. Why does enhancement occur, and why does crack growth rate sometimes retard?
2. Several possible mechanisms for crack growth retardation have been proposed, but it is
unclear which are operative under specific conditions.
3. There is a lack of understanding regarding the reasons for effect of sulfur content on crack
growth. Does this also affect S-N behavior?
4. Effects of flow rate appear to differ between S-N and crack growth testing.
5. Reasons for the different influences of DO/corrosion potential on S-N and crack growth
behavior are not known.
Note: Many observations relate specifically to PWR data but relevance to BWR requires
evaluation.
Research & Development Needs for KG 25 and 31
Knowledge Gap 25
Further tests are required to identify the circumstances that may lead to crack growth rate
retardation, and to quantify any predictable effect. This may prove difficult because of the
number of influencing variables.
Knowledge Gap 31
Mechanistic understanding is necessary to explain the apparent discrepancy between field and
laboratory data. Specific areas to be addressed could include:
• Mechanisms of retardation of enhanced crack growth.
• Is hydrogen generated by corrosion implicated in enhancement?
• Reasons for influences of material composition, e.g., sulfur content on crack growth
enhancement/retardation.
• Relevance of low and high flow rate laboratory data to plant conditions.
Ancillary Information
Chapter Section Subsection(s) Page(s) Applicable Materials
10 Discussion Crack Growth Data 10-6 to 10-7 Austenitic Stainless Steel

Background
ASME Section XI provides crack growth curves for austenitic stainless steel in inert/air
environments, but not for high-temperature water environments. At the time of publication of the
KG Analysis report [6], a Code Case was in preparation for PWR environments only. This has
now been published as ASME Code Case N-809 [5], which is based on so-called fully enhanced
data. However, under some circumstances, retardation of the enhanced rates towards the air rates
can occur. KG25, therefore, identifies a need for improved understanding of the conditions under
which retardation behavior can occur, and states that this would be required if benefit were to be
claimed in assessment. KG31 focusses more specifically on the mechanistic understanding of

4-6
13321451
Environmental Fatigue Crack Growth

retardation behavior. The available data indicate that variations in R ratio, temperature, coolant
flow rate and composition could impact crack growth retardation. Tests done on high sulfur
stainless steels have shown that this particular type of steel is much more prone to retardation
than low sulfur steel. Although the presence of retarded crack growth in some materials has been
shown, it is not adequately understood. Additionally, given the lack of any evidence showing this
behavior in an oxidizing BWR NWC environment, a need was also identified for testing to
determine the impact of different water chemistries. It was recognized that providing a
generalized substantiation of retardation based on environmentally specific compositional effects
may be difficult.
Review of Activities Relevant to KG 25 and 31
Knowledge Gap 25
The KG Analysis report [6] states that R ratio, coolant flow rate and material composition are
important factors influencing the onset of retardation. There has been significant ongoing
research, over the past 5 years, focusing on the impact that material composition has on
retardation effects since the 2011 Gap Analysis report. Studies by Tice et al. [53; 111], Mills
[110], Paraventi et al. [112; 113] and West et al. [114] have provided substantial insight into this
topic and shown that the amount of sulfur in a material has a strong influence over retardation. It
has now been clearly demonstrated that sulfur levels as low as about 0.006 % in Types 304/316
stainless steels (and their L variants) are sufficient to reduce the degree of environmental
enhancement at low frequencies/long rise times and that increasing the sulfur content closer to
the specification maximum (0.030 %) produces retarded rates at somewhat higher loading
frequency [53].
Although significant work has been done to understand the role the composition of the material,
there has been only limited further work examining the effects of coolant flow. Work by Platts
et al. [115] demonstrates the complex relationship between flow rate and retardation of enhanced
crack growth. This work indicates that at high flow rates, retardation is possible even for steels of
low sulfur content. Platts et al. observed retardation during low flow rate tests using stress ratios
greater than 0.7 at 300 °C. Tests using lower R ratios, at 300 °C did not show any retardation of
crack growth even when long (~2 days) rise times where used. Conversely, tests at load ratios
above 0.5 with high flow rates, as opposed to low flow rates, were found to produce retardation
in stainless steels when the coolant temperature was 250 °C. This work highlights the complex
nature of retardation of enhanced crack growth.
Knowledge Gap 31
Recent literature investigating the mechanisms of environmental enhancement of crack growth
has focused on the cause of enhancement or retardation of crack growth in stainless steels tested
in LWR environments [53; 110; 112; 114]. Mills [110] proposed two possible contributory
processes to enhanced crack growth: enhancement of planar slip by hydrogen, or inhibition of
slip reversal due to formation of an oxide film at the crack tip. Mills examined the fracture
surfaces of the specimens tested and found two distinct surface morphologies. The first, observed
during low ΔK (<13 MPa√m), consisted of cleavage-like facets with planar slip steps, which is
consistent with a planar slip mechanism. The second fracture morphology was observed at ΔK
values greater than 13 MPa√m, where the degree of enhancement was lower. In this case,
striation bands were also observed along with the cleavage-like facets indicating that both wavy

4-7
13321451
Environmental Fatigue Crack Growth

and planar slip mechanisms were active. These fracture surfaces were consistent with the
observations of Hanninen et al. [116] and have been linked to hydrogen embrittlement. Mills also
postulated that the oxide formed in the crack during the experiments could also contribute
towards the accelerated crack growth due to the oxide inhibiting slip reversal leading to
additional accumulation of plastic damage. However, he stated that the accelerated crack growth
rates were more consistent with those expected from a hydrogen embrittlement mechanism than
a purely oxide-driven mode. Therefore, Mills concluded that the enhancement to crack growth
from the hydrogen embrittlement mechanism was much more significant than the acceleration
from oxide formation. A recent study by Mukahiwa et al. [117] showed that tertiary cracking
was also visible in regions where the crack growth was enhanced which was considered to be
associated with embrittlement due to cathodically-produced hydrogen and strain
incompatibilities that develop at the crack tip during cracking.
Panteli et al. [118] conducted tests on low sulfur stainless steel CT specimens in air, argon
(containing 20 vpm of O2) and argon/hydrogen environments to produce a baseline in support of
their PWR tests investigating crack growth retardation. The crack growth rates for the tests done
in air and argon (containing 20 vpm of O2) were found to be similar. The argon/hydrogen gas
mixture was used to further reduce the levels of oxygen to obtain an “inert” environment, and
specimens tested in this environment showed a lower crack growth rate than that of the air or
argon environments. Post-test examination of the fracture surfaces of these specimens showed
significant oxidation for the specimens tested in argon and air environments, but very little in the
tests done in the argon/hydrogen mix. Furthermore, the planarity of the fracture surfaces of the
specimens tested in an argon/hydrogen mix was less than specimens tested in the other
environments. This could indicate that oxidation of the crack tip, rather than a hydrogen effect, is
inhibiting slip reversal resulting in the enhancement of planar slip for tests done in the argon
(20 vpm) or air environments. It was suggested that a similar effect may also contribute in high
temperature water, but with planar slip, and hence crack growth, further enhanced by hydrogen.
Retardation of enhanced crack growth in stainless steels has been investigated by multiple
authors [53; 110; 112; 114]. This is in part due to the potential benefits that could be claimed
from being able to use the retarded rates as opposed to the fully enhanced rates. However, to take
full advantage of retardation a better understanding of the mechanisms involved is required to
determine how applicable these crack growth rates are across material grades. Insights have been
made into the mechanisms for the effect of sulfur on crack growth, with a clear demonstration
that the dissolution of MnS inclusions from the steel creates an aggressive environment within
the crack enclave that enhances corrosion and oxidation [119]. Tests conducted by Panteli et al.
[118] and Platts et al. [119] in which sulfide (Na2S) was injected into the crack of a low sulfur
stainless steel specimen caused retardation similar to that observed in high sulfur materials,
therefore clearly showing that sulfide in the crack can retard crack growth.
Characterization of crack tips of high and low sulfur specimens have been performed using
transmission electron microscopy [111; 114; 118]. Examination of the crack tip regions revealed
that high sulfur steels exposed to cyclic fatigue tests typically contained a non-continuous
sulfur-rich layer between the metal and oxide in the crack tip region. In a number of cases, the
sulfur enrichment was coincident with locally elevated nickel levels. In contrast, no sulfur
enhancement was detected in the crack tips of the low sulfur material and the outer oxide
coverage was substantially reduced, containing discrete Fe-containing crystallites. The fracture
surfaces also appeared substantially different with discrete planar faceting for low sulfur steels

4-8
13321451
Environmental Fatigue Crack Growth

and a more diffuse appearance with high oxide crystallite coverage for high sulfur materials.
Fracture surfaces of the low sulfur specimens with sulfide injection appeared visually similar to
the high sulfur samples.
Despite the clear evidence of the role of sulfides from MnS dissolution, there is still a debate
surrounding the exact mechanism involved in retardation with hypotheses including
oxide-induced crack closure (most likely at low R), corrosive blunting and injected vacancy
enhanced creep (IVEC) [53; 110]. Oxidee-induced crack closure has been postulated to cause
retardation by the reduction of the crack driving force due to the crack becoming filled with
corrosion products. IVEC and corrosive blunting are proposed to be most significant at high R
ratios with low ΔKs. IVEC is suggested to occur when corrosion of the crack tip creates
hydrogen-stabilized vacancies in the base metal surface at the tip of the crack. This may have the
effect of reducing crack tip stresses, blunting and breaking up planar slip bands, which would
result in a reduced crack growth rate. Corrosive blunting, on the other hand, is simply the
blunting of the crack tip by corrosion in the local sulfide-rich crack tip environment. Either of the
latter mechanisms may be responsible for retardation of enhanced crack growth observed during
high R tests on high sulfur materials. However, confirmation of the responsible mechanism is
difficult and still under investigation.
Since retardation of enhanced crack growth rates appears to be dependent on the sulfur content
of the material, there is a concern that depletion of the sulfur-rich environment in the crack may
reverse the beneficial effect of sulfur. This situation may happen under plant loading conditions
with long hold times making it difficult to claim a benefit from retardation in high sulfur
materials in practical situations. Some evidence for such an effect has been seen in tests on a
high sulfur steel with 2-week hold times interspersed between periods of relatively fast cyclic
loading [53] in which the retarded crack growth rate increased following the later hold periods.
Similarly, testing for long periods at very long rise times has sometimes produced a restoration
of enhanced crack growth following retardation at slightly higher rise times [118]. A similar
effect has also been observed using high flow rates which have resulted in increased growth rates
due to depletion of the sulfur-rich local environment [53]. These recent results suggest that it
may prove difficult to take benefit from retardation in higher sulfur steels in practical
circumstances where long periods of operation under relatively steady load may lead to depletion
of the sulfur rich environment.
Summary
Although there have been significant developments in this area since the Gap Analysis report,
especially with regard to understanding of crack growth retardation behavior in austenitic
stainless steels in PWR environments, there are still gaps in mechanistic understanding. KGs 25
and 31 recognize that retardation effects in austenitic stainless steel in PWR environments are
not well understood and that more data are needed if a benefit is to be included in fatigue crack
growth assessment curves for a PWR environment (KG23). It is also noted that there is no
available evidence for a similar effect in BWR NWC; although data are more limited. KG25
highlights the need for improved understanding of the factors influencing retardation such as
steel sulfur content and water flow rate; whereas, KG 31 emphasizes the need to better
understand the fundamental mechanisms driving retardation of enhanced crack growth to
ascertain if a benefit can be claimed for crack growth retardation in some materials. Despite
these outstanding issues concerning retardation, recent evidence suggest that it may be difficult

4-9
13321451
Environmental Fatigue Crack Growth

to guarantee retardation under plant conditions due to loss of sulfur from the crack enclave
during long hold periods close to steady load. Therefore, it is considered that KGs 25 and 31
should therefore remain open, but with reduced level of priority compared to more plant specific
issues such as complex and non-isothermal loading.

4.2.4 Knowledge Gaps 29 and 30


Knowledge Gap Definition
Knowledge Gap 29 (Previously Classified as Category C, Medium Priority)
PWR: Reference crack growth curves [for carbon and low alloy steels] are available covering
both BWR and PWR but they do not explicitly represent the influence of all significant factors
on the degree of enhancement such as DO concentration and transient rise time.
ASME CC N-643-2 provides alternative crack growth curves for PWR environments according
to sulfur content and rise time.
BWR: Sufficient data are available, but there is a requirement to develop an improved
assessment code for BWR.
For BWR HWC, ASME XI reference curves may be excessively conservative, but may be non-
conservative for some BWR NWC conditions.
Knowledge Gap 30 (Previously Classified as Category D, Medium Priority)
Mechanistic understanding for carbon and low alloy steel is better than for austenitic stainless
steel, but some uncertainties remain, e.g., influence of time-dependent material deformation
behavior, e.g., dynamic strain aging.
Research & Development Need
Knowledge Gap 29
For BWR HWC it may be prudent to develop an approach based on ASME CC N-643-2, with
different criteria for EAC/non-EAC.
Knowledge Gap 30
Data relating to EAF in carbon and low alloy steels need to be incorporated in crack growth
codes, especially for BWR.
Ancillary Information
Chapter Section Subsection(s) Page(s) Applicable Materials
10 Discussion of Laboratory Data/Crack 10-6 Carbon/Low Alloy Steel
Knowledge Gaps Growth Data

Background
The ASME Section XI crack growth curves for carbon and low alloy steels are based on a data
set that does not recognize differences between BWR and PWR environments or the impact of
rise time on crack growth rate. A more recent data set, developed in the 1980s and 1990s, shows
a higher degree of environmental enhancement in BWR environments than PWR and a strong

4-10
13321451
Environmental Fatigue Crack Growth

dependence of crack growth rates on the rise time of the waveform. This results in the Section XI
curves being excessively conservative for PWR environments under some conditions, especially
for steels of low sulfur content, for which the degree of environmental enhancement in PWR
chemistry is small. It is important to recognize that this dependence of crack growth rate on steel
sulfur content for ferritic steels is the opposite of that for stainless steel, ASME Code Case
N-643-2 gives a more reasonable alternative for PWR environments, and also recognizes an
influence of both rise time and steel sulfur content on the degree of environmental enhancement
of crack growth. However, as yet, no suitable alternative exists for BWR conditions which takes
into account these effects, nor the difference between oxygenated (NWC) and hydrogenated
(HWC) water chemistries. Therefore, there is a need for more realistic assessment methodologies
that take into account the above factors whose influence is demonstrated by current data for
crack growth in BWR environments.
Review and Summary of Activities Relevant to KG 29 and 30
Only limited work directly addressing these KGs has been identified since the publication of the
KG analysis report. In that report, work by Seifert from 2008 and 2009 was cited and indicated
that the current ASME Section XI code assessment curves may either over-predict or under-
predict crack growth rates of ferritic steels in BWRs depending on the environmental and loading
conditions. A recent paper [107] states clearly that high crack growth rates are the result of
simultaneously maintaining a critical high sulfur anion concentration in the crack tip electrolyte
and a critical range of strain rate. In NWC, the sulfur ion concentration can be maintained over a
wide range of strain rates; whereas, at lower potentials, the enhanced growth rate falls to a value
just above the air rate for sufficiently low loading frequencies. The paper also describes work
addressing the influence of chloride contamination that was found to influence stress corrosion
crack initiation under dynamic straining conditions at very low strain rates. In contrast, an effect
on corrosion fatigue (at a level of 100 ppb) was found only at intermediate corrosion potentials
and very low loading frequency. As very little work has directly contributed to KG 30 it remains
open.

4.2.5 Knowledge Gap 26 (Previously Classified as Category C, Medium Priority)


Knowledge Gap Definition
Some data are available for Alloy 600 and its weld metals, which has enabled an assessment
curve to be incorporated in ASME XI. Alternative curves have also been published that appear
more conservative.
Research & Development Need
Need to understand basis of alternative assessment curves.
Ancillary Information
Chapter Section Subsection(s) Page(s) Applicable Materials
Discussion of Laboratory Data/Crack
10 10-10 Nickel-Based Alloy
Knowledge Gaps Growth Data

4-11
13321451
Environmental Fatigue Crack Growth

Background
Currently ASME Section XI gives nickel-based alloy crack growth curves based on Alloy 600
for BWR and PWR conditions. At high-temperatures and low loading frequencies the
acceptability of flaws in nickel-based alloys such as Alloy 600 and associated weld metals
(Alloys 182, 82 and 132) is usually considered to be governed by the stress corrosion crack
growth rate rather than the rates generated by fatigue. The Alloy 600 crack growth rates are
substantially lower than in stainless steels. Additionally, it was noted in the KG Analysis report
[6] that the Japanese had produced crack growth curves for Alloy 600 that were more
conservative than those used in ASME. This inconsistency from the international community
with respect to crack growth of Alloy 600 requires resolution.
Review and Summary of Activities Relevant to KG 26
Ogawa et al. [120] investigated fatigue crack growth of Alloy 600 HAZ and weld metal in
ambient and high-temperature air. The purpose of this work was to establish a fatigue crack
growth curve for both the HAZ and weld materials in support of the air curves for nickel-based
alloys adopted by the JSME for use in BWR reactors. This study demonstrated that the crack
growth behavior of the weld metal bound that of the HAZ with a moderate degree of
conservatism. Additionally, the ΔKTH value was determined from the air data of HAZ material
and found to be 3 MPa√m. This ΔKTH value is similar to that found by West et al. [108] for
Types 304/304L stainless steels.
Subsequently, Ogawa et al. [121; 122] proposed a reference curve for nickel-based alloys in
BWR (NWC) environments which included power law dependencies on ΔK, rise time and
(1-R), of the form:
𝒅𝒅𝒅𝒅
= 𝑪𝑪 ∙ 𝒕𝒕𝒎𝒎 𝒏𝒏
𝒓𝒓 ∙ ∆𝐊𝐊 ∙ (𝟏𝟏 − 𝑹𝑹)
−𝒑𝒑
Eq. 4-1
𝒅𝒅𝒅𝒅

where the constants and exponents differ between base metal (Alloy 600) and weld metal (Alloy
182). This approach was supported by experimental data with rise times down to 1000 s, and it
was found that fatigue crack growth saturated at rise times between 30 s and 1000 s. However, a
later paper [122] noted that, at longer rise times, SCC may be dominant over corrosion fatigue.
Additional crack growth data in BWR (NWC) environments is provided and a slightly modified
relationship, but with a similar form to Ogawa’s equations, is proposed which takes account of
the SCC contribution down to longer rise times. The revised equation is claimed to be
appropriate down to rise times of 30,000 s; although, it is stated that an alternative approach is to
use a superposition model of inert fatigue and SCC.
Although some work has been done towards supporting the Alloy 600 crack growth curves in air
and BWR environments, no work in PWR environments within the last five years has been
identified. Therefore, KG 26 remains open.

4.2.6 Knowledge Gap 27 (Previously Classified as Category C, Medium Priority)


Knowledge Gap Definition
Very limited data [on fatigue crack growth in LWR environments] are available for Alloy 690
(and [weld metals] 52, 152, and variants).

4-12
13321451
Environmental Fatigue Crack Growth

Research & Development Need


More fatigue crack growth data are needed for Alloy 690.
Ancillary Information
Chapter Section Subsection(s) Page(s) Applicable Materials
10 Discussion of Laboratory Data/Crack 10-7 Nickel-Based Alloy
Knowledge Gaps Growth Data

Background
As noted above, ASME Section XI gives nickel-based alloy crack growth curves based on Alloy
600 for BWR and PWR conditions. The acceptability of flaws in nickel-based alloys, such as
Alloy 600, is usually governed by the much faster SCC crack growth rate rather than the rates
generated by fatigue. However, the higher chromium content of Alloy 690 means that it is much
less susceptible to SCC and, therefore, the fatigue behavior of this material may be more
significant. Additionally, it was noted in the KG Analysis report [6] that the Japanese had
produced crack growth curves for Alloy 600 that were more conservative that those used in
ASME. This inconsistency from the international community with respect to crack growth of
Alloy 600 and the lack of Alloy 690 data means that there is a need for a limited amount of data
to support analysis of crack growth in Alloy 690 and its weld metals.
Review of Activities Relevant to KG 27
Xiao et al. performed crack growth tests on Alloy 690 CT specimens in a PWR environment to
investigate the effect of crack tip plastic zone size (Xiao, et al., Effects of Crack Tip Plastic Zone
on Corrosion Fatigue Cracking of Alloy 690(TT) in Pressurized Water Reactor Environments,
2015) and dissolved oxygen (Xiao, et al., Effects of Dissolved Oxygen on Corrosion Fatigue
Cracking of Alloy 690(TT) in Pressurized Water Reactor Environments, 2015) on the EAF
behavior. The acceleration in crack growth in PWR water vs. the air environment was used to
calculate Fen values for the test results. Xiao et al. reported Fen values as large as 6 for low DO
PWR environments and found that Fen increased with decreasing monotonic plastic zone size. On
examination of the crack morphology they found that all cracks were transgranular in nature and
in some cases there was significant branching and crack deflection. Laser microscopy revealed
that crack branching and deflection became more pronounced as the cyclic plastic zone
decreased. They stated that there was no clear dependence of Fen on the cyclic plastic zone size
and the relationship between Fen and ΔK was also uncertain. Xiao et al. postulated that small
amounts of crack deflection or branching could result in significant decreases in the crack growth
rates (Xiao, et al., Effects of Crack Tip Plastic Zone on Corrosion Fatigue Cracking of Alloy
690(TT) in Pressurized Water Reactor Environments, 2015) and, therefore, fracture mechanics
would be unlikely to accurately predict corrosion fatigue cracking under small scale yielding in
these specimens.
When investigating the impact of DO on the EAF crack growth behavior of Alloy 690, Xaio
et al. found that the largest values of Fen were found for low DO conditions (<5 ppb). The tests
were done at two different DO levels of <5 ppb and 9 ppb and specimens loaded under constant
K and R conditions with frequencies of 2, 0.5, 0.25 and 0.1 Hz. By using laser microscopy to
characterize the fracture faces and by comparing these results with the monotonic plastic zone
size it was found that that for the higher DO level the monotonic plastic zone size was smaller

4-13
13321451
Environmental Fatigue Crack Growth

and there was more out-of-plane cracking than was found for the low DO environment.
Furthermore, this work showed that the Fen factor was larger for the low DO than the high-DO
environment. This is consistent with the approach of NUREG/CR-6909 Rev 1 (Chopra &
Stevens, Effect of LWR Water Environments on the Fatigue Life of Reactor Materials Final,
May, 2018), which advocated the use of the stainless steel Fen equations for nickel-based alloys.
Summary
There has been a very limited study of the crack growth behavior of Alloy 690 which has
focused on characterizing the detrimental effect of PWR environments in high and low DO
environments. This work has shown a link between monotonic plastic zone size and Fen for this
material in addition to confirming that low DO environments enhance crack growth to a greater
extent than high-DO environments. This was rationalized by investigating the fracture face
morphology and revealed that higher levels of crack branching and deflection occurred in high-
DO environments. In general, it was observed that the greater the level of crack branching and
deflection the lower the crack growth rate during the test.

4.2.7 Knowledge Gap 32 (Previously Classified as Category C, Low Priority)


Knowledge Gap Definition
Do existing uncertainties concerning SCC propagation behavior in Alloy 690 have any
implications for EAF, for which the available database is very limited?
Research & Development Need
There is an initial requirement for additional data to expand the available database.
Ancillary Information
Chapter Section Subsection(s) Page(s) Applicable Materials
Fatigue Crack
4 Cr-Ni-Fe Alloys 4-13 to 4-17 Nickel-Based Alloys
Growth

Background
KG 27 relates to the limited corrosion fatigue crack growth data for Alloy 690 in LWR
environments. The PWR environment is the main focus because this is where Alloy 600 is most
commonly employed. This is due to Alloy 690 being a replacement for Alloy 600 which has a
relatively high susceptibility to SCC in PWR primary coolant. Despite the good service
experience of Alloy 690, it has been shown to be exhibit relatively high SCC growth rates if in a
moderate to highly cold worked condition (typically >15-20 % although there is significant
variability in the literature). Extensive reviews of this topic have been performed by Hickling
and the latest update is reported in EPRI MRP-237 Rev. 2) [125]. KG 32 was identified to
address the possibility that interactions between corrosion fatigue and SCC might be significant
for Alloy 690 and its associated weld metals.
Review of Activities Relevant to KG 32
The majority of studies on SCC in Alloy 690 have involved introducing cold work as a means of
enhancing SCC, initially with the aim of comparing crack growth rates with similarly cold
worked Alloy 600, although this approach has received some criticism [125]. Although most

4-14
13321451
Environmental Fatigue Crack Growth

studies use a series of pre-cracking stages using cyclic and trapezoidal loading, the resulting data
are not directly relevant to fatigue in non-cold worked Alloy 690. However, the presence of cold
work of up to 15% has been identified in the heat-affected zones of Alloy 690 weldments, so that
data obtained on actual or mock-up weldments might be of some plant relevance.
Perosanz et al. [126] tested crack growth specimens manufactured from a weld mock-up in PWR
conditions, as well as control tests on 20 % cold worked Alloy 690 CRDM and plate materials.
For both weld metal (Alloy 52 and 152) and Alloy 690 HAZ, crack growth rates during the
cyclic loading phases of the test at frequencies down to 10-3 Hz were below the air rates for
Alloy 182 described in NUREG/CR-6721 [127]. Crack growth rates with respect to time
increased when hold periods of 2.6 or 24 hours were imposed, indicating a contribution from
SCC for both HAZ and weld metal. Crack growth was more variable under constant load
conditions, with the crack growth rate obtained in the HAZ material from the mock-up (which
had 11 % effective cold work) being slightly lower than the crack growth rate observed in the 20
% cold worked material.
Significant testing of weld metal and HAZ using combinations of cyclic and trapezoidal loading
has also been carried out in a PWR environment by Alexandreanu et al. Crack growth rates for
Alloy 152 weld metal under cyclic loading were generally close to or slightly below those for
Alloy 182 from NUREG/CR-6721 [127] with the crack morphology changing from transgranular
to intergranular when hold periods were imposed [128]. A test on Alloy 690 HAZ [129] showed
some intergranular features during the cyclic phase of the test, but this had only a marginal
influence on crack growth rates compared to the parent material. Additional tests on Alloy 152
weld butter near the low alloy steel interface, where dilution of the chromium content was
expected to have occurred, showed that the susceptibility of the material to SCC potentially
increased. The range of crack growth rates obtained from this testing highlighted susceptible
regions of the weld but did not correlate well with the surrounding chromium content [130].
Further tests were carried out by Alexandreanu et al. on an Alloy 52M overlay on Alloy 182
[131; 132]. In this case, cracking was initiated in Alloy 182 under cyclic loading, after which the
loading was changed to trapezoidal and then constant load, with the growth rate slowing as the
crack approached the interface. A so-called “gentle cycling” regime (1000 s hold with periodic
unload every 4 h) was then imposed. When the crack approached the interface, it diverted along
the chromium-depleted interface zone with intergranular morphology and a propagation rate
described as high. Shallow intergranular penetrations into the 52M weld metal were also
observed. Further testing on Alloy 52M weld overlay material was done in the ST direction
where the SCC crack growth rates were measured as a function of orientation and distance from
the weld interface. Tests done in the ST direction found that the interface between the two metals
and first layer of Alloy 52M material was highly susceptible to SCC and correlated with the local
concentration of chromium, unlike the behavior observed for Alloy 690 HAZ.

Summary
Some work on Alloy 690 and its weld metals has been performed with a view to assessing the
extent to which SCC and environmental fatigue interact. These data indicate that SCC can occur
in Alloy 690 subjected to bulk cold work of 10-15 % or greater, and crack growth rates are
usually higher under trapezoidal loading than at constant load, except at the highest levels of cold
work. Only limited evidence for SCC has been reported in weld HAZ material or in weld metal,
but when it does it appears to be enhanced by a degree of cyclic loading. Regarding corrosion

4-15
13321451
Environmental Fatigue Crack Growth

fatigue, the available data indicate that crack growth rates for Alloy 152 and 52 are only slightly
lower than for Alloy 182, but nevertheless are significantly lower than for stainless steel.
However, there is still large scatter in the available data and this is an ongoing area of research.
Despite the contributions made so far to this KG, there still remain uncertainties over the
interactions between SCC and fatigue crack growth for Alloy 690 and its weld metals. Due to
this and the need for further data, KG 32 remains open.

4.3 Complex Waveforms and Non-Isothermal Loading

4.3.1 Knowledge Gap 28 (Previously Classified as Category A, High Priority)


Knowledge Gap Definition
Very few [crack growth] data are available under plant-representative loading conditions and the
influence of complex loading conditions (including hold times and spectrum loading) waveforms
and combined loading are not well quantified. Crack growth data are obtained under isothermal
conditions whereas many facility plant transients involve simultaneous temperature and load
cycling (either in-phase or out-of-phase).
Research & Development Need
Further tests are required to investigate the influence of different loading waveforms and better
represent the temperature and loading conditions experienced in real plants.
Ancillary Information
Chapter Section Subsection(s) Page(s) Applicable Materials
10 Discussion of Laboratory 10-8 All
Knowledge Gaps Data/Component Testing

Background
Plant transients often involve simultaneously occurring thermal and load cycling which produce
complex waveform shapes, potentially containing both tensile and compressive portions. These
waveforms differ substantially from the continuous triangular or sine wave cyclic loading
waveforms, usually at positive stress ratio, which are typically applied during laboratory testing
under isothermal loading. Many nuclear plant components also experience long periods at
relatively constant load interspersed by relatively infrequent transients. The KG Analysis report
[6] suggested that, although some data indicated that there may be some benefit in terms of crack
growth from some aspects of non-continuous loading, for example hold times, there appeared to
be inconsistencies between the data generated in different laboratories. Therefore, a definite need
was identified for additional data to clarify the impact of complex loading conditions, which
include hold times, spectrum loading, non-isothermal loading, waveform shape and negative R
loading.
Review of Activities Relevant to KG 28
Hold Times
Nuclear plants spend significant time at a nearly constant stress while operating. This means that
plant is subject to sets of transients broken up by periods at nearly static loading. It is noted that
steady load operation is usually at considerably lower load than the peak load which arises

4-16
13321451
Environmental Fatigue Crack Growth

during some transients, such as thermal shocks. The effect of periodic hold times on the fatigue
life of metals is not currently well understood, but it has been suggested that it could
significantly influence crack growth.
PWR Environments
Platts et al. [133] performed a limited study of the effect of plant-relevant waveforms on fatigue
crack growth of austenitic stainless steel. The work indicated that under some circumstances
(loading regime, temperature, material composition) hold periods at maximum load (i.e.,
trapezoidal loading) can lead to crack growth rate retardation. More recent data, from the same
laboratory [134] concluded that the retardation observed for trapezoidal loading is largely
determined by the sulfur content of the steel and is only observed for steels with sulfur content
greater than about 0.006%. The degree of retardation for trapezoidal waveforms was similar to,
but not greater than, the predicted fatigue crack growth rate for sawtooth loading with similar
cycle durations.
Tice et al. also performed testing using hold periods for steels of varying sulfur content [111].
Again, a substantial influence of sulfur content was observed, with one heat containing 0.004 %
sulfur showing retardation using trapezoidal loading, in contrast to a 0.001 % sulfur steel. Tests
were also performed using packets of cycling followed by long (14 day) hold periods. In this
case, the degree of retardation was substantially reduced compared to trapezoidal loading for the
same average hold time per cycle, suggesting that the benefit of holds saturated for extended
hold times. Moreover, holds at mean load were much less effective than holds at maximum load.
A further study indicated that retarded crack growth rates can be expected at low loading
frequencies (or long rise times) for steels with a sulfur level greater than ~0.006 %; whereas, for
slightly lower sulfur levels (~0.004 %) significant variability between the behavior of different
steel heats was observed [53]. It was also found that the benefit of elevated sulfur levels may be
lost where sulfur species are able to escape from the crack enclave, which can occur after
extended cycling at long rise times, during long hold periods or at high flow rates.
A study on the impact of hold times between 60 and 3000 seconds on the crack growth rate of a
limited number of 304 austenitic stainless steel heats was performed by Mills et al. [110]. Mills
found that the crack growth rates observed decreased with increasing hold times up to 1200 s
after which point the effect saturated. The crack growth rates associated with hold times of
>1200 s was found to be within a factor of 2 of the air rates. Mills observed a significant
variation in the onset of crack growth retardation between different heats of the same material
grade with one giving 20 % reduction in crack growth rate for 60 s holds (severe retardation) and
another heat giving the same level at 600 s (moderate retardation). Inspection of the fracture
surfaces for the moderately and severely retarded specimens revealed significant coverage by
corrosion products. It was also observed that switching between sawtooth waveforms and those
with holds produced an abrupt change from clear faceting of the fracture face to one covered in
corrosion product. This observation was repeated on the return from holds to the original
waveform.

4-17
13321451
Environmental Fatigue Crack Growth

Mills suggested that, based on the comparison between the two heats tested, the defining factor
was the sulfur level in the material. The work demonstrated that the higher sulfur material clearly
had more oxidation products on its fracture face and the highest level of retardation. Mills argued
that the presence of higher levels of sulfur in the crack results in increased corrosion enabling an
IVEC mechanism to relax the stress at the crack tip giving the observed crack growth retardation.
One possible consequence of the results reported by Platts et al. [111; 133] and Mills et al. [110]
is that in the crack growth dominated low cycle fatigue regime, there is a possibility of a
significant decrease in the level of environmental reduction of fatigue life for intermediate and
high sulfur steels. However, this has not been observed in practice and has not yet properly
studied with fatigue life experiments (see KG 20 Issue 4).
BWR Environments
Seifert et al. have performed tests on fatigue pre-cracked stainless specimens in BWR NWC and
HWC environments. [106] They tested ten stainless steel materials (Types 304/304L, 316L, 321
and 347 grades) with varying sulfur contents ranging from 0.003 % to 0.025 %. The test
procedure used periods of cyclic loading with rise times from 5-5,000 s interrupted by hold times
at maximum load. The hold times lasted from 30 hours up to a maximum of 300 hours for HWC
or 744 hours for NWC. For the HWC environment, no significant retardation of fatigue crack
growth occurred as a result of static hold periods at maximum load for most of the austenitic
stainless steels studied, except for one Type 347 heat at very long rise times. The same was
generally true for stabilized or annealed materials in NWC conditions. Evidence for intergranular
SCC was found in tests on sensitized stainless steel in NWC, and enhanced crack growth for
hold times of 100 hours or above. In contrast, Roth and Devrient [135] had previously reported
beneficial effects for hold times of 3 and 30 days for both stainless steel and low alloy steel in
oxygenated BWR conditions (400 ppb DO) at 240C. Although the steel sulfur content is not
reported in that paper, subsequent discussion with the author indicates that the tests were on a
high sulfur material, 0.018 wt.%. Comparison of the stainless steel with the results of Seifert
suggests that the benefits of hold times in retarding crack growth seem to occur only for
relatively high steel sulfur contents. Although this is similar to PWR environments, it is likely
that the threshold sulfur content for the benefit to be observed may differ between the two
environments, as suggested by Platts et al. [134]. These data reinforce the observation stated in
the KG Analysis report relating to the paucity of data in BWR environments addressing
retardation effects on crack growth.
In contrast to stainless steel, there have been very few studies of hold time effects on fatigue
crack growth in ferritic steels. The study by Roth and Devrient [135], referred to in the Gap
Analysis report, was conducted on both austenitic stainless steel and low alloy steel in
oxygenated BWR NWC conditions. Similar to stainless steel, retardation of crack growth was
observed as a result of long holds for both materials, with the degree of retardation increasing
with increasing hold time from 3 to 30 days.
Plant-Specific Waveforms
Work investigating the impact of exponential and sawtooth waveforms has been reported by
Platts et al. [136]. The basis for this work was the idea that a plant will experience thermally
driven transients and that the stresses associated with the down-shocks would rise exponentially
to its maximum. In this work, the exponential waveform was approximated into three rising

4-18
13321451
Environmental Fatigue Crack Growth

segments and one falling segment. When comparing the crack growth rates for stainless steel in
PWR environments between the two waveforms it was found that the rates resulting from the
sawtooth waveforms bound and agreed well with those of the exponential. The influence of
single overloads between packets of smaller amplitude cycling was also investigated and shown
to retard crack growth under some conditions. Further work by the same laboratory [133]
investigated the influence of a range of spectrum loading and complex waveform shape on
fatigue crack growth in PWR water. It was found that, in many cases, total crack growth for a
complex waveform with variable rise times was significantly less than would be predicted using
the ASME Code Case equations combined with either the average rise time over the transient or
that calculated using the modified strain rate approach. Instead, it was found that a better
prediction of observed behavior could be obtained by assuming that greater contribution to
fatigue crack growth occurred if the slower part of the rising load was in the upper portion of the
transient. It was demonstrated that the damage done to a material is weighted towards the
maximum load of the waveform and that the damage appeared to be proportional to the localized
plastic strain at the crack tip. Therefore, each section of the waveform’s contribution to the
effective rise time influencing environmentally enhanced fatigue crack growth is dependent upon
its position within the waveform itself. The data generated in this study shows that for many
waveforms, which are more representative of plant than simple triangular or sawtooth loading,
the simplistic application of Code Case N-809 based on total rise time can be overly
conservative. A modelling approach for fatigue crack growth under complex loading has been
developed by Emslie et al. based on the test data and is described further in Section 5.4.2.
Negative R Testing
The great majority of laboratory fatigue crack growth rate testing, including all the data in LWR
environments which form the basis of ASME Code Case N-809, were obtained under positive
stress ratio (R>0). In contrast many plant transients involve changes between tensile and
compressive stressing, i.e., are under negative stress ratio. Data under these conditions in LWR
are very limited. It is usually assumed that only the positive-going (rising load) portion of the
load cycle contributes to fatigue damage and, in LWR conditions, the crack growth rate increases
as the rise time in the tensile portion of the loading cycle increases. However, opinions differ as
to whether or what proportion of fatigue damage occurs during the positive going part of the
cycle that is in the compressive loading region. For example, the UK R5 code assumes that 50 %
of the compressive part of the rising load portion of the cycle contributes to crack growth;
whereas, ASTM E647 assumes that there is no contribution during compressive loading; the
basis of these assumptions is not well documented. Watson et al. [137] described a study aimed
at quantifying the contributions to fatigue crack growth under negative R loading. It was
concluded that it was appropriate to assume that 33 % of the compressive portion of rising load
contributed to fatigue damage under elastic conditions (i.e., at low strain amplitudes), but 50 % is
more appropriate when plasticity is present (high strain amplitudes). The test data underlying this
study was in air at 260 C, but it was argued that the results were likely to be conservative for
PWR environments because crack closure was expected to be greater in that case. Platts et al.
[138] reported preliminary results from a fatigue crack growth study using corner-cracked
specimens under elastic-plastic loading. Testing in both air and PWR water was carried out and
the preliminary conclusion was that good agreement between crack growth rates under positive
and negative R could be achieved if the compressively loaded portion was ignored entirely.

4-19
13321451
Environmental Fatigue Crack Growth

Non-Isothermal Testing
The above testing was aimed at simulating the mechanical component of the complex transients
experienced by operating plant. In reality, since many transients are thermally induced, both
temperature and loading will vary in a complex manner during the cycle. Only very limited data
are available from thermo-mechanical testing of austenitic stainless steels on which to base an
estimate of what is the appropriate temperature to use when assessing thermal transients using
isothermal data. Platts et al. [136] have conducted thermo-mechanical crack growth rate testing
which compared in and out-of-phase thermo-mechanical loading. This indicated that for in-phase
loading the crack growth rate was close to that measured in isothermal tests conducted at the
maximum temperature of the cycle, whilst for out-of-phase loading, the crack growth rate was
closer to that at the lowest temperature. The out-of-phase condition is relatively similar, but not
identical, to that which occurs for most thermal transients. Further results are reported in a more
recent paper [67] that extended the range of temperatures and loading rates compared to the
earlier work. The weighted-K rate (WKR) method (discussed further in section 5.4.2) had
previously been shown to effectively predict the effective rise time for a range of isothermal
complex waveforms [139]. Platts et al. [67] applied the WKR method to the data collected from
non-isothermal crack growth tests and found that it adequately predicted the effective rise times
for in-phase loading for two heats of 304 stainless steel. However, the WKR predictions for the
out-of-phase loading data showed considerable conservatism when compared to the measured
rates. Two postulated reasons for the over conservatism in the out-of-phase loading tests were a
possible deficiency in the Code Case N-809 equations at low temperatures or an effect of the
thermal cycle on how crack tip plasticity develops during a loading cycle. It was considered that
there would be less crack tip plasticity in an out-of-phase loading regime compared to in-phase,
resulting in lower than predicted crack growth rates for the out-of-phase loading tests.
Nevertheless, when comparing the predictions for out-of-phase loading with those of Code Case
N-809 and average rise time through the cycle, it was shown that the WKR method better
predicted crack growth rates than the N-809 methodology. Therefore, the WKR method shows
promise for assessing fatigue crack growth for isothermal transients, although further validation
data are required.
Summary
The impact of plant-realistic waveforms on fatigue crack growth in LWR environments has seen
significant progress with recent data obtained from spectrum and complex waveform (albeit for
isothermal) testing. The demonstration that the damage contribution from part of a complex
waveform to the total amount of fatigue damage is dependent on its position in the waveform is
enabling alternative methodologies for calculating rise time to be developed. The first of such
approaches to take this into account is the WKR Method which takes cognizance of the idea that
the amount of damage done in a cycle is weighted towards the maximum load. This method then
demonstrated that it could account much more accurately for the reduced crack growth rates
generated by complex waveform loading than the use of Code Case N-809 in which the rise time
is assumed to be the sum of all the rising load portions of the complex waveforms (except very
slow ramps which correspond to stress changes below 1.92 kPa·s-1 [139]. Further verification of
this method is required to fully take advantage of it in assessment procedures.

4-20
13321451
Environmental Fatigue Crack Growth

Only limited published data are available on crack growth under non-isothermal conditions. The
work generally supports the observations found in non-isothermal testing on S-N type specimens
in that the appropriate temperature to use in calculating damage is closer to maximum
temperature when temperature and load cycling are in-phase and to minimum temperature when
out-of-phase. Although the WKR method also shows promise for application on non-isothermal
transients, refinement of the model may be required to improve the dependency of temperature
on crack growth in Code Case N-809. Considerably more data will be required to support the
development of analysis procedures that can utilize any advantages that this work may bring to
light.
Investigations into hold times and their effects on crack initiation and growth in stainless steels
have shown a positive influence of hold times in BWR HWC conditions, under which conditions
initiation is delayed. In BWR NWC only sensitized material experienced a reduction in the
number of cycles to fatigue initiation. For crack growth, retardation was linked to sulfur content
with only medium and high sulfur materials experiencing significant retardation in LWR
environments as a result of hold periods. Also, in PWR environments at least, the effect of hold
periods appeared to saturate relatively quickly so that long (two week) holds between periods of
continuous cyclic loading had a relatively small effect. However, it was observed that there was
a tendency for crack growth rates to return to fully enhanced rates once the sulfur in the crack
mouth had depleted. It is possible that differences in electrode potential between BWR NWC and
PWR or BWR HWC coolant chemistries may impact on rates of sulfur loss from the
environment and, hence, on the sustainability of sulfur-related retardation processes. Therefore,
the value of pursuing retardation due to hold times to benefit from reduced crack growth rates in
plant assessments may need to be questioned.
Overall, it is judged that sufficient data are available on the effect of hold times with respect to
crack growth retardation to conclude that no benefit can be consistently claimed under plant
loading conditions. In terms of plant-realistic loading and non-isothermal testing there is a
significant need to produce more data to validate and underpin emerging assessment
methodologies and to support mechanistic understanding of these topics. In the case of negative
R data, it is worthwhile to consider further work in this area to justify or improve on the
relationships for effective ∆K which have been proposed. Considering the need for more data to
support what appear to be areas in which significant benefits can be found to reduce
over-conservatisms, KG 28 should be kept open.

4.4 Summary of Crack Growth Data


The KG Analysis Roadmap report [7] sets out seven hypotheses which may contribute to the
apparent discrepancy between the predictions of fatigue assessment models developed from
laboratory data on small specimens and the in-service plant experience. Hypothesis 1, which is
defined as cyclically variable parameters in a thermally-induced stress cycle reduce or negate the
environmental influence on fatigue, highlights KG 28 as being a high-priority. Similarly,
Hypothesis 7 was defined as improved mechanistic understanding which would identify
circumstances where the application of the Fen factor approach is not required – indicates that
KG 31 is a high-priority. The recommendations of the Roadmap document state that Hypothesis
1 is the more significant of the two. These gaps center on complex and plant-relevant loading

4-21
13321451
Environmental Fatigue Crack Growth

(KG 28), and increasing the mechanistic understanding of environmentally-assisted crack growth
(KG 31). The Roadmap document, therefore, recommends that KGs 28 and 31 are given the
most attention and effort.
The majority of work done towards the KGs concerning crack growth data has focused on the
high-priority gaps identified by the Roadmap. In particular, KG 28 which deals with plant-
relevant loading has received a considerable amount of effort. This is broadly consistent with the
S-N data-based KGs that have received attention. Currently, the plant-realistic loading and
conditions are being widely investigated by the international community as it is believed that
large benefits can be claimed by increasing the understanding of the impact that thermal
transients have on fatigue life. Considerable advancements, from a crack growth point-of-view,
resulting from research into plant-realistic loading have been made. Potentially, the most
significant of these is the development of the WKR approach for calculating the crack growth
rate of complex transients as an alternative to the simpler methods currently used. Supporting
this new methodology was a large amount of crack growth data collected using complex loading
which demonstrated that the damage done to the specimen during the cycle was weighted
towards the maximum load of the cycle. It was inferred that the damage would be proportional to
the localized plastic strain at the crack tip.
KGs 25 and 31 are considered together for the purpose of this work as they deal very closely
with enhanced crack growth and retardation of stainless steels. KG 25 deals specifically with the
collection of crack growth data studying retardation and feeds into KG 31 which seeks to
determine the mechanisms of enhancement and retardation. In terms of data collection, a
substantial amount of work has been performed which investigates the role the sulfur content of
the material plays in retardation. Furthermore, it was noted that the roles of temperature and flow
rate have yet to be fully established. When investigating the mechanisms of enhancement, the
international consensus is that a hydrogen embrittlement mechanism is responsible for the
enhanced crack growth observed in LWR environments. This mechanism could produce the
faceted fracture faces observed on the specimens after testing. For retardation of enhanced crack
growth, there appear to be several mechanisms which could be responsible. Under low R ratios
and low ΔK, oxide-induced closure of the crack appears to be the most likely mechanism. At
high R ratios and low ΔK, there appears to be more uncertainty over the mechanism responsible,
but either IVEC or corrosive blunting of the crack tip appears to be the most likely. Although
mechanisms have been proposed that could explain retardation more data would be required
before they could be positively determined.
Significant advancement has been made towards the high-priority gaps dealing directly with
crack growth data and the understanding of environmental fatigue obtained from it. However, the
work done to date is not sufficient to close out any of the crack growth gaps. Further work
supporting the high-priority gaps is still needed and, in the cases of carbon/low alloy steels and
nickel-based alloys, the limited amount of work that has been done relating to plant relevant
loading requires significant expansion.

4-22
13321451
5
ASSESSMENT METHODS

5.1 Introduction
There is significant prevailing uncertainty concerning treatment of the effects of LWR
environments when designing for fatigue in nuclear plant components. Assessment methods have
been developed to account for environmental effects on both the inferred fatigue endurance life
and fatigue crack growth, for example those in NUREG/CR-6909 [3; 8] and ASME Code Cases.
However, these assessment methods can result in high fatigue usage factors and substantially
higher wetted crack growth rates for some PWR and BWR components. These predictions are
not consistent with accumulated field experience of no reported LCF failures of components in
existing plants, so they challenge continued ability to justify safe, long-term nuclear plant
operation. There is also a potential impact on plant management activities, such as in-service
inspection intervals or component replacement policies. Consequently, there is a need to resolve
current uncertainties concerning the understanding and treatment of EAF for LWR components
to understand the level of risk to the plant and minimize unnecessary economic burdens on plant
owners. This section will, therefore, review the current EAF assessment methods together with
the knowledge gaps identified in Reference [6], in order to provide a basis for the revised listing
of KGs in Section 7 which have been developed to focus both on EAF assessment methods and
supporting test work. This should support working towards developing assessment methods that
are less onerous and prohibitive in the demonstration of safe, long-term operation.
The general structure of this section is as follows:
• Summary of the mechanistic understanding of fatigue and EAF
• Review of key concepts
• Review of fatigue initiation assessment
• Review of fatigue crack growth assessment
• Summary of current assessment rules, codes and standards
• Updates on status of KGs (Fatigue initiation, Mean stress, Fatigue Crack Growth, Welds,
other KGs).

5.1.1 Summary of Mechanistic Understanding of Fatigue and EAF


The formation of fatigue cracks in a component comprises the stages of crack “nucleation”, the
formation of “microstructurally” small cracks, the formation of “mechanically” small cracks and
the formation of “engineering” cracks, typically considered to be of the order of 3 mm deep in
laboratory test specimens. In ASME Section III terminology, this is referred to as fatigue crack
initiation. The design curves methodology is a simple screening approach developed from fully-
reversed testing on membrane-loaded, polished small specimens where failure was defined either
as 25% load drop in solid specimens or failure in hollow specimens. The design curves in ASME

5-1
13321451
Assessment Methods

Section III are intended to prevent “fatigue failures”, but may yield overly conservative results
when these screening procedures are used to predict fatigue lifetime estimates for plant
components. An alternative methodology which considers each of the phases of fatigue lifetime
separately, may become a more attractive prospect to plant operators when traditional fatigue
design curves do not adequately estimate fatigue lives of actual components. This is especially
the case for thermal transients in thick-walled components where the crack growth phase occurs
in a through-thickness reducing stress gradient. Experimental investigations compiled in [29]
(Sections 2.2.2 and 3.3.2) have been formulated into such a methodology, and ASME Code Case
work is proceeding using such methods [140], but it is stressed that further testing is required to
verify the assessment concept.
The Stage 1 fatigue initiation phase is predominantly driven by the level of plastic strain that
correlates to the applied strain. In an inert environment, cracks of between 10 µm and 20 µm
long, inclined to the surface, form early in life at surface irregularities or discontinuities, grain
boundaries or second phase particles. The second phase of fatigue initiation involves the growth
of “microstructurally” small (Stage I) cracks up to about 250 µm long and are again often
inclined to the free surface. The growth of these cracks is very sensitive to microstructure.
Following this micro-crack formation, the subsequent crack growth (Stage II) results in a crack
of engineering dimensions, usually defined to be 3 mm based on standard S-N tests. The
proportioning of cycles for the formation of Stage 1 shear and Stage II tensile cracks to a
combined depth of 3 mm is different for the LCF and the HCF domains. For LCF, the cycles for
Stage I predominate whereas, for HCF, the cycles for Stage II predominate.
The EdF Energy R5 procedure [25] provides a numerical calculation for partitioning smooth test
specimen fatigue data into the Stage I and Stage II phases. This procedure is discussed in more
detail in the R5 Assessment Methodology Section.
EAF, also known as corrosion fatigue, is a process that, under certain environmental
circumstances (e.g., in LWR water), can both reduce the period over which fatigue initiation
occurs and enhance crack growth rates. This is a chemical process whereby the high-temperature
water interacts with the metal surface under cyclic stress. The mechanisms by which LWR
environments influence fatigue life and crack growth appear to differ between materials (e.g.,
ferritic and austenitic steels) and also between reducing (PWR and BWR HWC) and oxidizing
(BWR NWC) environments. Environmental influences also appear to differ between stages of
cracking, but this will also be influenced by the stressing conditions, e.g., in the absence or
presence of a significant strain gradient.
Other damage mechanisms due to environment, such as irradiation and stress corrosion cracking
(SCC), have the potential to interact with EAF. These are not considered in detail in this report,
although the potential for SCC interacting with EAF is something that is discussed in relation to
nickel-based alloys (see Section 4.2).

5.1.2 Review of Key Concepts


A fatigue assessment generally involves a comparison of the cyclic loads experienced by a
component or structure in its design life against its resistance to fatigue. The resistance must be
sufficient to withstand the loads without exceeding an allowed value. Two distinct methods are
employed in assessments and depend on whether the structure is defect free or it is assumed to
contain a crack-like defect. The first approach is called the fatigue initiation life assessment (also

5-2
13321451
Assessment Methods

known as fatigue endurance or S-N method), and is predominantly used in design calculations to
demonstrate the structure is expected to remain defect free over its intended operating life. This
approach relates nominal applied cyclic stress (or strain) ranges with the number of allowed
cycles (Nf) from a suitable fatigue design curve. The data forming the basis of most design
fatigue (S-N) curves are usually generated from small-scale, cylindrical uniaxial tensile
specimens between 6 and 12 mm diameter, and the “failure” is defined to be either complete
separation of the specimen or a specified load drop (25% is commonly used). The other approach
(often called flaw tolerance assessment) requires the use of fracture mechanics and the
postulation of a crack in the structure - usually based on detection capability limits for pre- or in-
service inspection. The flaw tolerance assessment procedure requires calculation of the crack
driving force, which is then used to calculate the crack growth rate using material-specific data.
The amount of allowable fatigue crack growth (FCG) is assessed based on an allowed crack size
determined from a fracture assessment. One of the major advantages of the flaw tolerance
approach is that calculation of cumulative usage factors and Fen multipliers (i.e., the fatigue
initiation assessment approach) is not needed.

5.1.3 Fatigue Initiation Assessment


The intent of the fatigue initiation (or fatigue life) assessment approach is to demonstrate that the
structure will remain defect free over its anticipated lifetime. The cyclic loading is analyzed to
derive a stress (or strain) range for each loading cycle anticipated over the component design life.
The incremental fatigue damage is then calculated for each loading cycle, where identical cycles
are grouped together. Damage, in this context, is defined to be the inverse of the predicted life
from a fatigue endurance (S-N) design curve. The partial usage factor is then calculated by
multiplying the applied number of loading cycles with the incremental fatigue damage. The sum
of all partial usage factors for all applied loadings are added together to generate the Cumulative
Usage Factor (CUF). Generally, a CUF of less than unity is taken to demonstrate fatigue
assessment acceptability. The summation methodology is commonly referred to as “Miner’s
rule” because it was popularized by M. A. Miner in 1945 [141]. The most common method of
accounting for environmental effects involves the use of an environmental enhancement factor
(Fen). Fen values are determined for each partial usage factors and the two are multiplied to obtain
the EAF CUF. Since Fen values can exceed 15 under some circumstances, this may result in very
high EAF CUFs for some component locations.
To distinguish between transients in a potentially very complex loading sequence, there are a
variety of cycle counting methods available to break up loading spectra into individual applied
loading cycles. One such methodology commonly used for real-time loading histories is called
“rainflow-counting” which reduces a complex loading spectrum into a set of simple stress cycles
that then allows the use of Miner’s rule.
Transient Parameters Influencing Environmental Enhancement
The environmental enhancement factor is a function of environment (oxygen level, metal
temperature 2) and loading conditions (strain rate). For non-isothermal loading transients, there is
a need to understand what effective temperature should be used in evaluating EAF. For transients
with varying strain rates throughout the cycle, there is a need to capture either the change in Fen

2
And steel sulfur content in the case of carbon and low alloy steels

5-3
13321451
Assessment Methods

throughout the cycle or an effective overall strain rate for the cycle. Furthermore, the strain
amplitude is known to influence Fen due to the positive correlation between mechanical damage
and environmental enhancement. Therefore, a more precise definition of the transient, through
material properties definition and detailed stress analysis, is often required to underpin current
EAF assessment methods. This provides confidence in the more simplified assessment methods,
without being unduly conservative. However, a linkage of some of the detailed assessment
methods for temperature averaging, strain rate determination, etc. to supporting tests that verify
their validity is lacking.
Transient Definition
In design and assessment space, a transient is defined to be a change in one plant state to another
over a certain period of time, such that the loading for that particular transient can be considered
in isolation. The transient is then analyzed separately and the necessary parameters for an
assessment are derived. For fatigue initiation, this consists of the partial usage factor that is
calculated from the S-N curve using cycle counting methods and Miner’s rule, as described
above. For FCG, the stress intensity factor (K) for the postulated flaw is derived and used to
obtain the crack growth rate. This crack growth rate is then used with cycle counting methods to
obtain the amount of crack growth over the component life (or, until the next required
inspection). Generally, this is used to define an appropriate period and frequency for inspection
of the component to verify the absence of actual cracks. Such an approach is usually taken if the
fatigue initiation assessment produces a CUF that exceeds unity.

5.1.4 Fatigue Crack Growth Assessment


Fatigue crack growth is based on the assumption of a postulated crack present in a component.
The applied loading is combined with the assumed crack geometry to generate a stress intensity
factor. The range of the K through the cycle is then used in a crack growth law which relates the
increment of crack growth per cycle, da/dN, with the K range, ΔK. The formula to calculate this
is called Paris’ law, as it was introduced by P. C. Paris in 1961 [142].

5.1.5 Summary of Current Fatigue Assessment Methods


Based on laboratory testing, one of the most significant factors influencing fatigue behavior in
LWRs is the influence of primary coolant chemistry on fatigue lifetimes and fatigue crack
growth rates. Considerable reductions in the number of cycles to failure have been observed in
fatigue endurance tests in LWR environments on small tensile specimens when compared to
experiments conducted in air. Similarly, enhancement of fatigue crack growth rates is observed
using pre-cracked laboratory specimens in simulated plant water environments. The degree of
enhancement of fatigue damage or CGR is a function of several factors, including material and
composition, coolant environment and temperature, and of mechanical loading, including
loading rate. These observations from laboratory tests have prompted a significant international
effort to improve the fatigue life predictions for components subjected to LWR plant conditions.
The method that has been widely adopted involves the use of life reduction factors that account
for environmental effects. These environmental correction factors, or Fen values, reduce the
lifetimes predicted by the S-N curve. Similarly, for FCG, the growth rates are increased

5-4
13321451
Assessment Methods

according to the level of environmental enhancement by the use of modified crack growth
reference curves. This section will summarize the latest developments in environmental fatigue
assessment methods since the EPRI KG report was published in 2012.
R5 Assessment Methodology (2014 Revision)
The R5 assessment procedure is used in the United Kingdom to assess both defect-free
structures, and structures with defects, at high-temperature. The aims of the procedure are to
estimate, by a simplified approach based on elastic stress analysis, the steady state cyclic stresses
and strains to determine crack initiation in the structure. R5, Volume 2/3, gives a procedure for
assessing defect free structures and, although it is not intended to provide an estimate of the
number of cycles to failure of a component, it does give the crack initiation endurance which is a
lower bound to failure. In its current form, the procedure was developed primarily for high-
temperature air or gaseous environments where fatigue and creep interactions can occur, and it is
not directly applicable to LWRs where creep effects are negligible but the effect of primary
coolant may be significant. Furthermore, the procedure does not specify the long-term damage
which may result from irradiation, erosion or severe corrosion.
Section 10 of R5, Volume 4/5, is dedicated to fatigue crack growth where Appendix A3.3
provides the inputs to the calculations. For small crack sizes, i.e., where the crack depth is
smaller than the cyclic plastic zone size, high strain laws are provided.
Other factors affecting fatigue endurance are considered in R5, and are accounted for by the
application of a Fatigue Strength Reduction Factor (FSRF). The FSRF is derived from the ratio
of strain ranges which give the same endurance for the different materials or conditions.
Reduction factors are also provided for weldments and these are discussed under KG 38.
R5 is considered a total life prediction methodology, in that it can account for the nucleation and
growth phases, where the nucleation is defined as the creation of a defect 0.02 mm deep. The
number of cycles to produce an initiated defect, a0, may be accounted for by the procedure
described in R5, Volume 2/3, Section 8.2 and Appendix A10.1. One advantage of using a total
life approach is the ability to take advantage of strain gradient effects in plant lifetime
assessments. This can be done by calculating fatigue life to a smaller crack size (e.g., 250 µm)
than the 3 mm crack which is assumed to relate to failure in conventional membrane-loaded
tensile S-N specimens, and then performing a crack growth calculation to a critical crack size or
failure. Application of such an approach has been described by Mann et al. [57], and is discussed
in Section 3.2.4 in the context of KG 7. Other methodologies are available, but all share the same
objective, which is to provide a strain amplitude to be used in conjunction with the S-N curves,
or a cyclic stress intensity factor to be used with crack growth rate curves. Environmental effects
are generally considered as an additional factor to be applied once the stress analysis calculations
have been performed. However, there is a growing consensus that environmental effects should
be considered from an early stage of the assessment due to the close connection of environmental
enhancement with mechanical strain.
Codes and Standards
Although there exist best estimate approaches such as R5 to assess components and structures,
codes and standards are most commonly used throughout the world to design nuclear plants. A
brief overview of some of the more widely used codes and standards is provided in this section.

5-5
13321451
Assessment Methods

ASME Section III, Subsection NB provides rules for the design of Class I Components.
Evaluation rules for fatigue are given in NB-3200, “Design by Analysis,” and NB3600, “Piping
Design” [1]. The rules provide a methodology for calculating strains and use pre-defined fatigue
endurance curves. Design fatigue curves are provided for specific classes of materials to
establish the suitability of a component for cyclic service and define the allowable number of
cycles as a function of applied alternating stress amplitude.
The ASME Section III fatigue design curves do not explicitly account for the effects of LWR
environments. Laboratory test data indicate that such effects could be substantial. This led the
NRC to fund research performed by Argonne National Laboratory that developed expressions for
the impact of PWR and BWR coolant environments on fatigue life, which were documented in
several reports, including NUREG/CR-6909 [8]. These expressions utilize an environmental
fatigue correction factor (Fen), which is a function of temperature, strain rate and water chemistry
(and, for ferritic steels, sulfur content). In 2007, the NRC issued guidance on the use of this
assessment methodology for application to new plants in Regulatory Guide 1.207. A revision to
NUREG/CR-6909 Rev 1 [3] was issued in May 2018 and includes a number of changes to the
expressions used for calculating Fen (see Section 3.2.2). Regulatory Guide Rev 1 [2] was also
issued in May 2018 that incorporates the changes in NUREG/CR-6909 Rev 1.
Based on the issue of NUREG/CR-6909, ASME Code Cases have now been published for
austenitic stainless steel to incorporate environmental effects. Code Case N-792-1 uses an
environmental correction factor (Fen) based on the methodology in NUREG/CR-6909 Rev. 0. A
second Code Case, N-761, provides EAF fatigue (S-N) curves which are intended to describe S-
N data in high-temperature water. These Code Cases may be subject to revision based on the
recent publication of NUREG/CR-6909 Rev 1 [3].
The French Code RCC-M [143] contains similar rules on fatigue analysis to ASME Section III
and the latest editions incorporate methods to evaluate environment effects; details are discussed
later in this report. The British Standard PD5500, Annex C [144], or the equivalent European
Standard BS EN13445 [21], provide design rules for fatigue assessments in air. There are
separate design curves for weldments but no explicit guidance is given on how to treat
environmental effects. All these Codes and standards propose simple rules to evaluate the strain
amplitude using simplified elastic approaches and simplified plastic strain correction factors
(e.g., Ke (ASME Section III, NB-3200) and Kv (Section B 3234.6 of RCC-M)). These factors are
applied to the elastically calculated stresses, the results of which are then used in the fatigue
design curves. KTA is the German Code for assessing primary components [17] and has
so-called “attention thresholds” to modify the CUF for environmental effects. The plasticity
correction methods in the KTA Code are the same as those in ASME Section III. The JSME
Code [145] alternative to the Ke factor is complex and is calculated from a series of
mathematical formulae describing curve fits. A good comparison of these various plasticity
correction models was developed by Emslie et al. [146]. The number of allowed cycles are then
determined from the transient combination rules (“Rainflow”) and damage cumulative procedure
(Miner’s rule). It is generally recognized that the stress methods employed by most codes and
standards procedures are conservative for application to actual plant components. This
contributes to some of the disparities observed between fatigue life predictions vs. field
observations independent of the application of EAF effects.

5-6
13321451
Assessment Methods

Applicability of Codes and Fen to Plant Components


Field experience suggests that the application of predicted environmental effects using the
NUREG/CR-6909 (Rev. 0 and Rev. 1) or ASME Code Case assessment procedures for fatigue
life are overly conservative for application to actual plant components and transients. One
potential reason for this over-conservatism is thought to be that, although the majority of plant
transients result from variations in thermal loading, most available data are derived from
isothermal testing. Up to now, the majority of models used to assess environmental effects have
been empirical, and are highly dependent on the testing conditions that were used to derive the
models. This limits the assessment to consider changes between plant states in a simplistic
manner, with no account taken of peak/minimum temperatures occurring in or out-of-phase with
the peak/minimum strain peaks. The modified rate approach outlined in NUREG/CR-6909
recommends integrating Fen along the actual plant transient strain profile, if this is known, and
this can be very complex to perform, even with modern FEA tools. For simple triangular
temperature waveforms where temperature varies both in and out-of-phase with strain
waveforms, there is no difference between the predicted Fen values using the modified rate
approach. This result is incorrect because temperature peaks occurring with strain minima are
less damaging than temperature peaks occurring with strain maxima. Therefore, a significant
amount of recent work in the assessment community has been focused on addressing this issue.
For example, the work performed by Currie, et al. [66] weights the calculation of Fen based on
the amount of damage at each point in the cycle. This has been applied to thermo-mechanical
fatigue (TMF) data and has been found to yield good predictions of fatigue lives. In contrast, the
modified rate approach gives good predictions for in-phase TMF loading, but is overly
conservative for out-of-phase loading.
Fatigue Crack Growth
Reference fatigue crack growth curves for carbon and low alloy (ferritic), austenitic stainless,
and nickel alloy steels are provided in Section XI of the ASME Boiler and Pressure Vessel Code.
These curves are applicable for both air and water environments, with the exception of austenitic
stainless steels where only air FCG rates are provided. For ferritic steels, the reference curves
provided in Appendix A of Section XI for PWR and BWR environments are based on data prior
to 1980. Code Case N-643-2 provides updated FCG curves for PWR environments and
recognizes the effects of loading frequency and steel sulfur content on the degree of
environmental enhancement. Fatigue crack growth rate curves for austenitic stainless steels in
PWR environments are also available in ASME Code-Case N-809 [5], which include parameters
such as rise time, stress intensity factor range, load ratio and temperature. Efforts to update N-
809 for BWR environments is getting underway in 2018. Reference fatigue crack growth curves
for PWR and BWR environments are also available in Appendix C for Alloy 600 and related
weld metals (Alloys 182, 132 and 82).
Alternative reference crack growth curves are available or under development for incorporation
into The Rules on Fitness-for-Service for Nuclear Power Plants of the Japan Society of
Mechanical Engineers (JSME).
Although the above reference crack growth curves can be used to provide a conservative
prediction of crack growth from existing or postulated flaws for plant components, they may
prove excessively conservative for complex or thermal transients. An alternative method, which
involves weighting environmental effects to the more damaging portions of the applied loading,

5-7
13321451
Assessment Methods

was developed by Emslie, et al. [139], which considers the rate of change of stress intensity
factor in the calculation of crack growth rates in water. This method has been shown to provide a
more realistic prediction of measured crack growth in laboratory tests using complex isothermal
waveforms or non-isothermal testing. This is discussed further in Section 5.4.2.
Summary
Generally, there is a need for more refined assessment procedures to reduce conservatism and
more accurately represent actual component stresses and predicted fatigue lifetimes. Historically,
fatigue assessments have used the results from laboratory testing in simplified format, i.e., a
crack growth curve, or an S-N curve. Therefore, if the assessed component is exposed to an inert
environment, and the stress analysis is conservative, then the fatigue assessment will also be
conservative. Similarly, the effects of water environments are assessed conservatively. These
conservatisms make continued justification of safe, longer term component acceptability and
reliability difficult.
Parameters that estimate the effects of water chemistry, dissolved oxygen, sulfur content,
temperature, strain rate, and rise time are required for the calculation of crack growth or fatigue
initiation. Plasticity corrections, as documented in ASME Section III [1] and R5 [25], provide
conservative estimates of stress and strain, based on simplified elastic analysis and material-
specific stress-strain data. This methodology significantly over-predicts the strain amplitude such
that, when it is used in the S-N curve, the fatigue lifetime may be substantially under-predicted.
However, when these same data are used to calculate Fen factors, the use of a maximum strain
amplitude does not result in a minimized strain rate, so calculated Fen values may be non-
conservative. Therefore, the assessment of a component in a water environment requires
consideration of upper and lower bound stress/strain to ensure it is appropriately conservative.
These effects were documented in studies by Stevens and Davis [147].
Several of the KGs highlight the issue of knowing detailed transient history and how the
environmental factor should be calculated. Detailed stress/time history is more readily available
now that FEA is routinely used in fatigue assessments or on-line fatigue monitoring is available.
However, the FEA stress/time history reflects the fluid temperature and pressure used as input;
therefore, best estimate fluid temperature and pressure, along with accurate FEA, provide the
most comprehensive inputs to fatigue initiation and fatigue crack growth calculations. A tradeoff
is that the fatigue initiation and crack growth models must accommodate this detailed
information to ensure the assessment is not unduly conservative whilst at the same time not
being overly complex.
The current state of affairs with respect to EAF methods is that there are mechanistic models
available, but they require data to underpin their validity. These models, if sufficiently validated
with extensive test data, will provide closure to a large number of assessment related KGs (4, 7,
8, 15, 39, 41 and 42 for fatigue initiation life; 23 and 28 for crack growth). Nevertheless, despite
showing some degree of promise of being closed out with the latest results and further
understanding of environmental fatigue test results, these KGs still need to remain open due to
their lack of refinement based on lack of supporting test data. In this sense, the models can be
considered semi-empirical, in that they have some physical basis, but they are also dependent on
fitting the models to data with appropriate fitting parameters.

5-8
13321451
Assessment Methods

The rest of this chapter summarizes the status of these KGs in more detail. Certain KGs, even
though they are very relevant to environmental fatigue assessments, describe more general
assessment issues in their own right. Examples include KG 38, which considers the different
strain state at a feature such as a weld, or KG 40, which discusses stress indices to account for
stress modifications to pipe bends and elbows. Both are issues in their own right that are
independent of EAF effects, but they also significantly influence the results of environmental
fatigue calculations.

5.2 Fatigue Initiation

5.2.1 Knowledge Gap 2 (Previously Not Classified)


Knowledge Gap Definition
There is a lack of correlation between expectations from laboratory test data and plant operating
experience, which does not give confidence in the methods being developed for the treatment of
corrosion fatigue in LWR environments.
Research & Development Need
A wide ranging investigation into the basis of the Fen factor, or similar approaches, and their
application to plant transient analysis is required.
Ancillary Information from the 2011 KG Report
Chapter Section Subsection(s) Page(s) Applicable Materials
Critical Review of Design Compounding Individual Factors
9 9-3 All
Code Developments to Define the Fen Factor

Background
Currently, design and assessment procedures assume that observations of the EAF phenomena
are transferable between components of different sizes and shapes. This is said to apply under the
assumption that any differences between influences such as loading, boundary conditions,
material properties and environment are accounted for. This principle is also assumed when
using test data produced using small-scale laboratory specimens to predict the fatigue life of
actual plant components. However, as identified by the KG Analysis document, although a
number of plant component fatigue failures have occurred, the mechanism of low cycle corrosion
fatigue was not applicable [6]. Rather, these failures were caused by additional loading cycles
that were not explicitly anticipated or accounted for during the design phase, or were the result of
high cycle fatigue caused by thermal mixing or vibrational fatigue. This lack of correlation
between predictions using methods based on laboratory test data and plant operating field
experience undermines the confidence in the current methodologies used to assess corrosion
fatigue.
Review of Activities Relevant to KG 2
Alternate Fen Methodologies
In an effort to address the interactions of the various factors influencing transference and
environmental factors used to assess EAF, EdF proposed a modification to the Fen methodology
[44] presented in NUREG/CR-6909. The idea for this new method of calculating Fen, developed

5-9
13321451
Assessment Methods

specifically for austenitic stainless steels in PWR environments, is to account for the portion of
environmental effects that are considered to be already accounted for by the fatigue design curve.
EdF’s proposal moderates the environment factor through the use of a Fen-integrated value in the
situations where the calculated Fen exceeds Fen-integrated. In this method, Fen-integrated is defined as a
measure of the environmental effects which are already covered in the fatigue design curves. For
complex transients, a detailed Fen evaluation is proposed wherein only positive strain rate
sections of the transient are considered. The positive strain rate sections are divided up into
increments and the Fen factor is calculated using the equation given in NUREG/CR-6909. EdF
advises on the use of the maximum temperature of the transient to calculate Fen to ensure
conservatism, which is consistent with recommendations given in NUREG/CR-6909 Rev 1 [3].
This aspect of complex transients is discussed further in KGs 8 and 43.
The basis for Fen-integrated was discussed by Le Duff et al. [47] and was developed further by
Métais et al. [31; 44]. It was initially proposed that Fen-integrated can take two discrete values of 3.0
for sawtooth waveforms and 5.0 for thermal shocks; the higher value was intended to account for
strain gradient effects in thick-walled components. If the calculated Fen, using either
NUREG/CR-6909 or the more detailed calculation outlined in the previous paragraph, is less
than the appropriate Fen-integrated value, the original CUF calculated by fatigue analysis is
validated. Otherwise, the partial CUFs are modified by the ratio between Fen and Fen-integrated to
calculate the environmentally corrected CUF. Considering that, prior to the introduction of
NUREG/CR-6909, it was assumed that environmental factors were inherently accounted for
within the ASME Section III design curves, it is not surprising that some institutes have tried to
quantify this further and use it to lower calculated Fen factors.
Solin et al. [71; 72] proposed an alternate method to improve transferability of laboratory test
data to fatigue assessment of plant components. In this method, the effects of material, water
environment, temperature and service loading patterns on fatigue life are discussed. The authors
found that sequences of cycles and hold periods extended the fatigue lives of stainless steel
specimens in comparison to the results of standard cyclic experiments. The authors proposed a
Fen factor that is composed of three parts consisting of temperature, water, and transferability.
The inclusion of the temperature factor was to ensure that the effect of temperature on fatigue
life was not accounted for twice. The purpose of the transferability factor was to account for the
mapping from laboratory data to plant conditions. For example, plant-realistic effects such as
hold times are accounted for in a fatigue life transferability factor, Ftransferability. The authors
acknowledge that further mechanistic understanding is required to underpin this approach and its
potential benefits, and they present a method that could be applied to plant assessments by
defining the Fen in terms of temperature (including the effects of dynamic strain ageing), the
water environment, and the transferability to plant components. The authors emphasize that
further re-evaluation of existing data, as well as refinement of the methodology, are required
before it can be successfully applied to plant assessments.
Thermal Gradients
Although some components may be subjected to significantly varying membrane loads, the
majority of loading transients on power and propulsion reactor systems are thermal in origin with
rapidly decreasing strain gradients through the thickness of the component. Conventionally, the
ASME Section III procedure requires the fatigue life and, therefore, the CUF for a component,

5-10
13321451
Assessment Methods

be calculated based on the surface strain or equivalent stress in the component. This is likely to
impart a significant, potentially excessive, degree of conservatism in the plant assessment since
the design fatigue curves are based on membrane-loaded test specimens.
Very limited plant component testing, such as the Bettis stepped pipe tests [91], attempted to
justify alternative assessment methods to ASME Section III / NUREG/CR-6909, or to justify
continued operation in specific cases when CUFs exceeded unity. More recent examples of such
work consist of Total Life approaches [148; 149; 57] and comparisons between different
assessment methodologies [150; 151; 152]. Both Kamaya [58] and Harrison [150] use plant-
relevant testing to highlight the over-conservatism in current analysis methods, especially for
thermal transients that contribute to the majority of calculated CUF.
The approach taken by Gosselin et al. [140] attempts to benefit from two apparent discrepancies
in translating laboratory test data into information relevant to actual plant components. The first
discrepancy deals with the difference in the wall thicknesses of actual plant components
compared to small-scale laboratory test specimens. The second discrepancy deals with the
difference caused by through-wall strain variations in actual plant components compared to the
pure membrane loading applied to laboratory specimens. This latter discrepancy means that, for
thermal transients, the driving force for crack propagation decreases into the wall thickness, thus
leading to a significantly longer fatigue life in actual plant components compared to membrane-
loaded test specimens.
Gosselin’s approach introduces a life factor, LF, and a gradient factor, GF, to deal with enhanced
life due to thickness effects and through-wall strain gradients, respectively. Both of these factors
are assessed based on the cycles for Stage 1 initiation and the cycles for Stage 2 crack growth.
They are intended to readily fit into existing fatigue calculations so that they are easy and cost-
effective to implement. The life factor is a ratio between the number of cycles to a 3 mm deep
crack in a cylindrical test specimen to the number of cycles associated with a 25% load drop in a
cylindrical specimen with a thickness equal to the component thickness. The gradient factor is
defined by the equation below:
𝑵𝑵𝑰𝑰 𝑴𝑴𝑴𝑴𝑴𝑴𝑴𝑴𝑴𝑴𝑴𝑴𝑴𝑴𝑴𝑴 + 𝑵𝑵𝑰𝑰𝑰𝑰 𝑴𝑴𝑴𝑴𝑴𝑴𝑴𝑴𝑴𝑴𝑴𝑴𝑴𝑴𝑴𝑴
𝑮𝑮𝑮𝑮 = Eq. 5-1
𝑵𝑵𝑰𝑰 𝑴𝑴𝑴𝑴𝑴𝑴𝑴𝑴𝑴𝑴𝑴𝑴𝑴𝑴𝑴𝑴 + 𝑵𝑵𝑰𝑰𝑰𝑰 𝑮𝑮𝑮𝑮𝑮𝑮𝑮𝑮𝑮𝑮𝑮𝑮𝑮𝑮𝑮𝑮

The GF is the ratio between the number of cycles under uniform membrane loading to the
number of cycles under gradient loading. Gosselin assumed that, because stage 1 growth occurs
over the first 200 µm, membrane loading could be assumed for both laboratory specimen and
component loading conditions. Therefore, he proposed that the main difference in life occurs as a
result of the difference in crack growth under membrane versus gradient loading under stage 2.
In an example calculation for a carbon steel reducing elbow in a BWR low pressure core spray
system, Gosselin highlighted the advantages of the approach by calculating environmental CUFs
at 60 and 80 years that were 28% lower. Crucially, the use of the life and gradient factors
resulted in environmental CUFs of less than unity for 80 years of operation for the carbon steel
reducing elbow evaluated, whereas the CUF exceeded the allowable value prior to application of
the life and gradient factors.

5-11
13321451
Assessment Methods

Summary
Considerable effort is now being invested into investigating the impact of different aspects of
thermal transients on the predicted fatigue lives of actual components compared to laboratory
specimens. Experimentation using complex waveforms, hold times, thermo-mechanical fatigue,
and thermal shock and multiaxial loadings are some of the examples of work topics that are
currently active which will underpin future closing of this KG; these aspects are covered by other
more specific KGs discussed in Sections 3.3, 3.4 and 3.5 of this report. Although significant
progress has been made against KG 2, the work performed to-date is not sufficiently mature to
close this gap. In terms of investigating the impact of plant-relevant transients on Fen and
development of a total life approach, the work producing data to support this is in an early phase.
The strain gradient factor methodology, although promising, again is very new and requires
further validation and testing. There is also the potential to develop some of the methods for
alloys other than austenitic stainless steel, and for BWR environments. KG 2, therefore, remains
open.

5.2.2 Knowledge Gap 4 (Previously Classified as Category B, High Priority)


Knowledge Gap Definition
The reasons for the apparent discrepancy between laboratory data and plant experience regarding
the effects of environment on fatigue are not fully understood. Excessive conservatism in the
current rules for design and/or the influence of complex loading may, at least in part, provide an
explanation.
Research & Development Need
The reasons for this apparent discrepancy require further understanding. Many of the research
needs identified in this report are ultimately aimed at resolution of this issue.
Ancillary Information from the 2011 KG Report
Applicable
Chapter Section Subsection(s) Page(s)
Materials
2 Background Summary: Residual Conservatism 2-14 to 2-15 All
and Prevailing Uncertainty

Background
Laboratory data from testing standard specimens under membrane loading in LWR environments
demonstrates that a reduction in fatigue life occurs compared to testing in air. However, the
operational experience reported by the nuclear industry does not align with predicted lives based
on this premise. This discrepancy between actual plant and laboratory experience undermines the
validity of using the current ASME fatigue design curves and environmental enhancement
factors, and highlights the need to reconcile laboratory data with operational experience.
Review of Activities Relevant to KG 4 and Summary
This KG is one of the key overarching gaps identified by the KG Analysis report [6] and is very
broad in nature. Answering this gap would require multiple other gaps to first be closed. A major
contributor to this gap is the data and analysis methods being developed in response to KG 2 that
are aimed at accounting for the difference between laboratory and plant data. For example, all of

5-12
13321451
Assessment Methods

the KGs contained within this section contribute to the data, and a large number of the analysis
KGs in Section 5 will directly support any conclusions eventually made under this KG. The
statement that “Excessive conservatism in the current rules for design and/or the influence of
complex loading may, at least in part, provide an explanation” appears to be supported by recent
work addressing complex and non-isothermal loading, by observed differences in design
transference factors between air and LWR environments, and by investigations of the
conservative nature of stress and CUF calculation methods in current codes and standards.
Closing out this KG will require a deep understanding of the impact of EAF on plant components
versus laboratory specimens, supported from both data and analytical points of view. Although
there have been significant recent developments for austenitic stainless steels in PWR
environments, considerably more work will be required before sufficient understanding can be
claimed to close this Gap. There is also a need to address BWR environments and alloys other
than austenitic stainless steels. As such, KG 4 should be considered open and represents an
overall statement for the need for ongoing studies on EAF.

5.2.3 Knowledge Gap 1 (Previously Classified as Category C, High Priority)


Knowledge Gap Definition
There is a disparity between the lower bound values of Fen derived by NUREG/CR-6909 and [the
Japanese Environmental Fatigue Evaluation method], EFEM.
Research & Development Need
Further analysis of available test data is required.
Ancillary Information from the 2011 KG Report
Chapter Section Subsection(s) Page(s) Applicable Materials
9 Critical Review of Design Minimum Fen 9-4 All
Code Developments Factor

Background
The Fen expressions for stainless and ferritic steels described in NUREG/CR-6909 Rev. 0 [8]
generate a minimum value of approximately 2, regardless of the number of transformed
parameters which do not meet their threshold values. Other methodologies, such as the Japanese
approach to Fen, have lower bound values close to unity if any parameter does not meet its
bounding values. The approach in NUREG/CR-6909 Rev. 0 is inconsistent with international
opinion and with its own logic, because a Fen value greater than unity in instances where
environmental conditions are not present is not technically correct.
Review of Activities Relevant to KG 1
NUREG/CR-6909 Rev 1 [3] has revised this position so that when one or more of the
transformed parameters do not meet their threshold values, the value Fen is unity. Concern has
been raised over the new limits applicable to the Fen expression in NUREG/CR-6909 Rev 1 [3],
in particular the maximum threshold of 7 % s-1 relating to strain rate (this was 0.4 % s-1 in
NUREG/CR-6909 Rev 0). This is described in more detail in Section 3.2.2. Having this limit set
so high means that the vast majority of plant-relevant transients and laboratory testing will still
be subject to a Fen of 2. Although, on the face of it, NUREG/CR-6909 Rev 1 appears to bring the

5-13
13321451
Assessment Methods

Fen expression into line with international expectations it does so in a superficial manner. This is
especially true considering the dramatic change in the upper limit on strain rate with very little
change in the data between the two Revisions of NUREG/CR-6909 to justify it.
Summary
Although formally addressed, the way in which the Fen expression has been made to give values
of unity still leaves the underlying problem extant. The remaining issue is that for the vast
majority of waveforms Fen will always apply, even in those with very high strain rates.
Considering that the main intent of this KG has been addressed and that its extant aspect (the
lack of data underpinning the selection of a 7 % s-1 threshold for strain rate) is directly
considered by KGs 10 and 13, KG 1 should be considered closed.

5.2.4 Knowledge Gap 5 (Previously Classified as Category C, High Priority)


Knowledge Gap Definition
The proposed new fatigue life curves for carbon and low alloy steels in water environments in
ASME Code Case N-761 cover only high strength materials, whereas the current ASME curves
also cover lower strength materials.
Research Development Need
Additional assessment curves for low strength ferritic steels in water environments need to be
included in ASME Code Case N-761.
Ancillary Information from the 2001 KG Report
Chapter Section Subsection(s) Page(s) Applicable Materials
9 Critical Review of Proposed Revisions of ASME 9-14 to 9-15 Carbon and Low Alloy
Design Code Boiler and Pressure Vessel Steels
Developments Design Code

Background
ASME Code Case N-761 [153] is an alternative approach to account for EAF effects compared
to calculating Fen. This Code Case provides S-N curves that directly account for EAF effects.
The intent of the case was to provide an easy, straightforward way for designers to address EAF
effects in fatigue assessments without the complication of calculating Fen values. A series of
curves are provided for different values of strain rate, together with a lower bounding curve
which can be used if the strain rate is unknown. This Code Case differs from the Fen approach
documented in NUREG/CR-6909 [3] and ASME Code Case N-792-1 [154] in that the fatigue
curves in the case are not based on the use of Fen values. As such, the curves in the case cannot
be duplicated from a simple application of the Code Case N-792-1 Fen values to the fatigue
design curves in ASME Section III. In addition, the background work to Code Case N-761
contained within it a set of design curves for stainless steel and another for low alloy/carbon
steels. The latter design curves covered both high and low strength materials. However, when
Code Case N-761 was published, it contained only the high-strength fatigue design curves for
carbon/low alloy steels.

5-14
13321451
Assessment Methods

Review of Activities Relevant to KG 5


Code Case N-761 has not been subsequently revised to include low-strength design curves. In
light of the above discussion, the KG 5 primary issue still remains unresolved. It is also noted
that there is no general acceptance of Code Case N-761 by the NRC or other members of the
nuclear industry and technical community [155]. The main reason cited is it does do not give
results consistent with Code Case N-792-1 [154] which uses a Fen approach similar to the
original version of NUREG/CR-6909 [3].
Summary
As the low strength design curves are still absent from Code Case N-761, KG 5, therefore,
remains open.

5.3 Fatigue Initiation: Complex Waveform Analysis

5.3.1 Knowledge Gap 8 (Previously Classified as Category A, Medium Priority)


Knowledge Gap Definition
For non-isothermal cycles, the issue of temperature selection appropriate to the full calculation
procedure of thermal analysis, elastic stress analysis, strain analysis, and Fen factor calculation
requires further consideration.
Research and Development Need
Further analysis of available test data is required to assess the most appropriate temperature to
use for analysis for non-isothermal cycles.
Ancillary Information from 2011 KG Report
Chapter Section Subsection(s) Page(s) Applicable Materials
Critical Review of Design The Use of Maximum or
9 9-5 to 9-6 All
Code Developments Minimum Temperature

Background
Modern analysis techniques allow non-isothermal stress analysis through the use of thermo-
mechanical finite element modelling. This analysis typically involves time-dependent
temperature input to a mechanical model and the resulting stress analysis will produce a
strain-temperature history. This can then be used as input to an appropriate Fen calculation.
Temperature-dependent material properties will increase the accuracy of such analyses.
However, if the assessment is through a more simplified route, then an appropriately
conservative temperature should be chosen, i.e., one that maximizes stress for an air fatigue
assessment, and one that maximizes the Fen for a water fatigue assessment. On the contrary, most
of the test data used to develop the Fen methodology were obtained from testing performed at a
range of constant temperatures, the results of which were then used to establish the temperature
variation in the Fen equations. This is believed to be quite different (and conservative) compared
to the case of varying temperature during a test, which is more reflective of actual plant
component operation.

5-15
13321451
Assessment Methods

The calculation of Fen for cycles with temperature variations is required to consider various
assumptions about the appropriate temperature to select for simplified analysis. The full
calculation concerning stress analysis, strain analysis and Fen factor calculation should be
considered by setting material properties for the cycle minimum temperature or cycle maximum
temperature. The difference in Fen factor for these two assumptions may be significant. Other
assumptions concerning forms of temperature averaging should be considered. Guidance should
be formulated on the appropriate single temperature to select for a simplified analysis where the
cycle is a thermal up-shock or thermal down-shock.
Review of Activities Relevant to KG 8
Thermal shock profiles occur when a change in temperature results in a change in stress. For a
typical cylindrical component containing a hot and pressurized fluid, a sudden drop in
temperature results in a rapid increase in stress. This means that the maximum stress occurs at a
temperature significantly less than the maximum temperature of the transient in the near surface
region of the component. Therefore, the use of maximum transient temperature for Fen
calculation is very conservative, but it does not reflect the temperature when the maximum stress
occurs; so, a lower effective temperature is generally sought by analysts. Effective temperature is
a key aspect of Fen calculation, and requires a modified rate approach according to NUREG/CR-
6909. Further to this, there may be situations where the strains that come about due to a thermal
shock are predominantly occurring when the temperature is below the threshold value. This links
in with KG 43, which suggests that environmental effects can be reduced if parameters are below
threshold values during the cycle.
For transients of a thermal origin, the temperature will vary throughout the event where the
minimum and maximum temperatures associated with the transient correspond to the maximum
tensile stress and compressive surface loading, respectively. For ASME Section III fatigue
assessments, there is a lack of clarity surrounding whether the maximum, minimum or mean
temperature should be used as the effective temperature (Teff) for such thermal transients. ASME
Section III stress-based fatigue design curves, used during fatigue assessments, are set using a
reference elastic modulus - at room temperature - and so a correction must be applied to the
design curves to account for the assumed temperature in the assessment. This temperature
correction produces a small change in the fatigue assessment’s elastically calculated stress range,
which will vary depending on the Teff used.
The effect of temperature selection on fatigue analysis is substantially amplified when
considering Fen which, according to NUREG/CR-6909 [8] (and available test data), is strongly
dependent on the strain rate and Teff. For stainless steels, the literature considered in Reference
[6] regarding this KG suggested that the average temperature should be used for fatigue analysis.
This was due to the minimum temperature being concurrent with maximum crack opening [156].
However, where there is uncertainty in the definition of the Teff, it is common practice to use the
upper bound value as a conservative estimate. Due to limited examples of data in the literature at
the time of publication, and the potential benefits with respect to reducing Fen, Reference [7]
suggested that more effort should be put into data analysis to clarify the issue surrounding
temperature selection. The basis for KG 8, therefore, is the need to determine the appropriate
point in a thermal transient to calculate a value for Teff. It should be noted that data generated to
address this KG may also be relevant to KG 15 which deals with variable temperature, strain rate
and stress data.

5-16
13321451
Assessment Methods

ASME Code Case N-792-1 [154] recommends using the average of the highest and lowest metal
temperatures of the surface in contact with the fluid in the transients constituting the stress cycle
or load set pair. To obtain the most accurate Teff, tests need to be carried out with different load
and temperature combinations and compared to an isothermal case. This will provide more
evidence to determine whether the modified strain rate approach is a valid method of predicting
fatigue initiation for non-isothermal transients or whether alternative methods provide a more
realistic, but adequately conservative estimate of fatigue life (see Section 5.3.1).
Summary
The thermo-mechanical data collected so far suggests that the use of an average temperature for
non-isothermal transients may lead to over-conservatism in fatigue assessments. Therefore, the
use of a lower temperature—which will produce a lower Fen—may be more suitable, but to claim
this benefit, a better characterization of the existing transients may be required.
In light of the above discussion and work done towards KG 8, there still remain outstanding
uncertainties over temperature selection in fatigue analysis. However, new lifetime prediction
models are being developed which account for varying transient temperature, and the choice of a
specific temperature to use for EAF calculation becomes less crucial. In addition, there are
several non-isothermal fatigue tests underway that may provide further insight for this KG.
Therefore, KG8 should remain open.

5.3.2 Knowledge Gaps 39, 41, 42 and 43


Knowledge Gap Definition
Knowledge Gap 39 (Previously Classified as Category C, High Priority)
The calculation of strain rate is required to evaluate EAF initiation life. While methods for the
determination of cycle effective strain rate can be proposed for conformance to ASME Code
analysis, there are very few experimental data or plant data that can be used to validate the
methods for use in corrosion fatigue assessments. Methods need to be consistent with
mechanisms that operate under plant conditions.
Knowledge Gap 41 (Previously Classified as Category C, High Priority)
Interpreting a plant transient with variable strain rate in terms of the single strain rate curves is
problematic, and no relevant guidance is given.
Knowledge Gap 42 (Previously Classified as Category A, Medium Priority)
Procedures for determining a cycle specific Fen factor are very important since they underpin the
application of test data to plant assessment. Test data supporting averaging procedures for
treatment of cyclically varying temperature, strain rate, and stress (tension or compression) are
sparse, and this represents a significant uncertainty, so there is a need to understand real behavior
under temperature variable conditions.

5-17
13321451
Assessment Methods

Knowledge Gap 43 (Previously Classified as Category A, Medium Priority)


NUREG/CR-6909 lacks guidance on the procedure to be followed for stainless steel when both
strain rate and temperature vary during a cycle. Thus the NUREG/CR-6909 model does not
consider the possibility that one or more of the influencing parameters may be outside its
threshold value at all times during a cycle so that the combined influence may be reduced or
negated.
Research and Development Need
Knowledge Gap 39
Further work is required to establish a clear approach for calculation of strain rate. This is likely
to include testing under complex loading conditions combined with analytical work to assess the
suitability of different evaluation approaches for effective strain rate. Work to develop improved
mechanistic understanding may also be required.
Knowledge Gap 41
Further consideration is required to develop a methodology for dealing with variable strain rate
that is not unduly conservative.
Knowledge Gap 42
Analysis is required to understand how the necessary simple representation of complex plant
transients results in conservatism.
Knowledge Gap 43
The concept of the “Modified Rate Approach” (MRA) could reasonably be applied as the
weighted average of the combined Fen factor around the cycle. However, the MRA has never
been successfully benchmarked against test results. Work is required to develop and validate a
suitable method for treatment of variations in both temperature and strain rate, and also to
provide a realistic treatment of variable temperature transients.
Ancillary Information from 2011 KG Report
Knowledge Applicable
Chapter Section Subsection(s) Page(s)
Gap Materials
Review of NUREG/CR- Calculation of
39 9 9-9 to 9-11 All
6909 Fen factor approach. strain rate
Proposed Revisions of
41 9 ASME Boiler and Pressure Section III 9-14 to 9-17 All
Vessel Design Code
Review of NUREG/CR- Cyclically Variable
42 9 9-7 All
6909 Fen factor approach Parameters
Review of NUREG/CR-
43 9 Multiple 9-4 and 9-7 All
6909 Fen factor approach

Background
Fen calculation is dependent on temperature, strain rate and strain amplitude, and there is a
minimum threshold temperature below which environmental effects become insignificant.
Assessing a transient using the maximum temperature can be unduly conservative as some or all

5-18
13321451
Assessment Methods

of the rising portion may be well below the threshold temperature. Methodologies are under
development which can account for the actual transient temperatures and, therefore, provide
more realistic predictions of life [154]. The strain also becomes significant once its value
exceeds a threshold, or fatigue limit. This is due to the increased damage that occurs when the
strains are no longer purely elastic. A Fen calculation methodology must accurately account for
the temperature and strain in the cycle to capture any regions where effects are nullified due to
threshold limits.
Review of Activities Relevant to the KGs
The MRA can be applied to transients where the strain rate and temperature vary through the
cycle [157]. This integrated approach gives a Fen which can then be used to calculate an effective
temperature. This is shown to generally give a lower Fen as the temperature is less than the
maximum temperature for a considerable portion of the cycle. Testing in [88] considers
triangular strain waveforms, in- and out-of-phase with triangular temperature waveforms. When
the temperature is in-phase with the strain, i.e., the maximum temperature occurs at the
maximum strain, and the minimum strain occurs at the minimum temperature, Fen calculation
using the modified rate approach could be considered an appropriate methodology. However,
this approach calculates the same Fen for out-of-phase conditions, and it has been shown that
there is up to a factor of two difference in these lifetimes [67]. This occurs because only the
strain rate is used in the modified rate approach, which for triangular waveforms is constant on
the up-ramp portion of the cycle. However, mechanistic understanding of the environmental
effect, supported by experimental data (see Section 3.3), indicates that more damage accumulates
at higher strain amplitudes, which is not captured by any of the Fen approaches. Therefore, a new
approach that weights the Fen according to the position in the cycle is required. Currie et al. [66]
evaluated several models which recognize the variation of damage accumulation with strain
amplitude using the modified rate approach. The Tsutsumi Binary Weighted (TBW) model
[158], “Plastic Strain Weighted” (PSW) and “Strain-Life Weighted” (SNW) models were all
applied to complex waveform isothermal and thermo-mechanical test data in a PWR
environment reported by Platts et al. [67]. Using the material-specific water curves for the
material heats considered, the agreement between the predicted Fen and the experimental Fen was
relatively good for all three strain amplitude weighted models, but there is some over-prediction
of lifetimes for out-of-phase loading. When the NUREG/CR-6909 Rev 1 [3] mean water curve is
used, the agreement between predicted and measured lifetimes was still better than using the
modified rate approach, but the predictions were somewhat more conservative compared to using
the material-specific data.
The SNW method is currently considered to provide the best combination of accuracy and
practicality for assessment purposes, and its derivation is briefly summarized below:
SNW
The weighting used by the SNW method is based on the observed phenomenon that
environmental damage increases with increasing plastic strain, which is a similar concept to that
adopted in the WKR method. In order to obtain plastic strain for use in the model, it is necessary
to have the complete hysteresis data for the cycle. However, this requirement is circumvented by
utilizing the S-N curve and the way damage is related to strain amplitude. The weighting curve is

5-19
13321451
Assessment Methods

then derived as the ratio of the instantaneous fatigue life at the current strain within the rising
load cycle, to that at the target strain amplitude. This is expressed mathematically for 2 ×
0.112 + 𝜀𝜀𝑚𝑚𝑚𝑚𝑚𝑚 < 𝜀𝜀 < 𝜀𝜀𝑚𝑚𝑚𝑚𝑚𝑚 as:
(𝟎𝟎.𝟓𝟓(𝜺𝜺−𝜺𝜺𝒎𝒎𝒎𝒎𝒎𝒎 )−𝑪𝑪) 𝑩𝑩
𝒘𝒘(𝜺𝜺) = � � Eq. 5-2
𝜺𝜺𝒂𝒂 −𝑪𝑪

where the constants B and C are obtained from the ANL fatigue strain-life data which is assumed
to obey the Langer form for an S-N curve of the form given by Equation 5-2. εa is the strain
amplitude in percent (%):
𝐥𝐥𝐥𝐥(𝑵𝑵) = 𝑨𝑨 − 𝑩𝑩 ∙ 𝐥𝐥𝐥𝐥(𝜺𝜺𝒂𝒂 − 𝑪𝑪) Eq. 5-3

The value of C used by Currie et al. [66] is consistent with the ANL value of 0.112. However,
the paper proposes that a value of 0.056 may provide better agreement with the data and
recommends further testing to optimize this value.
The weighting function in Equation 5-1 ranges from 0 to 1 and is plotted in Figure 5-1 below:

Figure 5-1
SNW weighting function

The following summation then calculates the weighted Fen according to “SNW”:
𝜺𝜺
𝑭𝑭𝒆𝒆𝒆𝒆 = ∑−𝜺𝜺
𝒂𝒂
𝒂𝒂
(𝒘𝒘(𝜺𝜺𝒊𝒊+𝟏𝟏 ) − 𝒘𝒘(𝜺𝜺𝒊𝒊 ))𝑭𝑭𝒆𝒆𝒆𝒆,𝒊𝒊 Eq. 5-4

This is an extension of the MRA which does not account for the higher proportion of damage in
the top of the cycle. Thus, the Fen in MRA is calculated as follows:
𝟏𝟏 𝜺𝜺𝒂𝒂
𝑭𝑭𝒆𝒆𝒆𝒆 = ∑−𝜺𝜺 𝒂𝒂
(𝜺𝜺𝒊𝒊+𝟏𝟏 − 𝜺𝜺𝒊𝒊 )𝑭𝑭𝒆𝒆𝒆𝒆,𝒊𝒊 Eq. 5-5
𝟐𝟐𝜺𝜺𝒂𝒂

These calculations are based on best estimate approaches, and are far removed from a design
curve approach, which would have factors of 12 on life and 2 on stress. This would shift the
predictions to well below the measured lifetimes. However, since the results are obtained from a

5-20
13321451
Assessment Methods

precisely controlled experiment, and the model has been developed using data from similar tests,
it is not necessary to apply the transference factors used to generate the design curve, to account
for any uncertainty in loading conditions, surface roughness, and material properties. Thus, when
the curve is used in a plant assessment, there is confidence that it provides a conservative
lifetime.
Summary
Whilst the MRA provides an improved means of calculation of strain rate for complex
isothermal transients compared to simple methods based on average strain rate, it does not
predict the experimental observation that fast-slow transients are more damaging than slow-fast
transients. The reason for this is well understood, and can be explained by the principle of
enhanced environmental damage at the peak tensile loading of the cycle, i.e., there is a strain
amplitude dependency on Fen. The MRA also substantially under-predicts fatigue life for out-of-
phase thermal transients. Therefore, improved prediction methods which recognize that the
degree of fatigue damage accumulation varies with position in the cycle show promise in being
able to remove unnecessary conservatism in assessment of complex and thermal transients.
Further refinement of the models and experimental data to underpin them is required. It is noted
that this KG refers specifically to stainless steel in a PWR primary environment, but it would be
beneficial to investigate whether similar behavior is observed for other materials such as ferritic
steels, and to environments relevant to BWRs. KGs 39, 41, 42 and 43 should, therefore,
remain open.
Mean Stress Effects

5.3.3 Knowledge Gap 6 and 44


Knowledge Gap Definition
Knowledge Gap 6 (Previously Classified as Category A, Medium Priority)
Stainless steels exhibit significant strain hardening and cyclic hardening so that a sharply defined
yield stress does not exist. The Modified Goodman correction is used to adjust zero mean stress,
fatigue endurance data to account for mean stress. The influence of using a higher yield stress in
the Modified Goodman correction is to shift the influence of the mean stress correction toward
low cycle fatigue. The extent to which this happens depends on the magnitude of the yield stress
assumed.
Knowledge Gap 44 (Previously Classified as Category C, Medium Priority)
The extent to which mean stress influences both low cycle fatigue and high cycle fatigue of
stainless steel in air, and an appropriate means by which it should be accounted for, are
considered to be significant KGs.
Research and Development Need
Knowledge Gap 6
Further analysis is required to consider how the Modified Goodman correction should be applied
to stainless steel endurance data.

5-21
13321451
Assessment Methods

Knowledge Gap 44
Further analysis may be required to address how mean stress influences both low cycle fatigue
and high cycle fatigue, which is important to prevent undue conservatism.
Ancillary Information from 2011 KG Report
Knowledge Applicable
Chapter Section Subsection(s) Page(s)
Gap Materials
Critical Review of Design Mean Stress
6 9 9-2 All
Code Developments Correction
Cyclically Variable
44 10 Assessment Methods 10-10 All
Parameters

Background
Fatigue tests are generally conducted under constant stress or strain. However, due to cyclic
hardening and softening, it is difficult to keep the stress and strain constant during the fatigue
tests, especially for stainless steels. This phenomenon must be well understood to ensure that the
results from the laboratory can be confidently applied to plant components.
Generally, tensile mean stress has always been considered as detrimental, while compressive
stress has been considered to be beneficial. However, mean stress is more detrimental in strain
control than in stress control, since shakedown to R = -1 occurs after relatively few cycles in the
latter case.
Review of Activities Relevant to KG 6 and 44
Taheri et al. [79] proposed a model for fatigue damage accumulation under variable amplitude
loading for austenitic steels in strain control, which does not require an elastic-plastic
constitutive law. The paper considers loading sequences and highlights the inadequacy of a linear
damage summation rule when random sequencing occurs. By analyzing the order in which high
(H) and low (L) amplitude loadings are applied, several conclusions are drawn. One conclusion
is the effect of a H load followed by a L load in stress control which, due to pre-hardening,
causes a lower strain amplitude cycle and hence a longer fatigue life. The converse is true with
strain controlled tests, as the work hardening from the H cycle will increase the stress cycle and
hence increase damage, as both stress range and strain range contribute to fatigue damage.
Therefore, the model can explain the larger detrimental effect of a tensile mean stress in strain
controlled tests than in stress controlled tests. There is a recommendation to consider this
enhanced hardening due to mean stress, which could be significant enough to overcome the
detrimental effect of actual mean stress in stress controlled tests. For strain controlled tests, the
effect of enhanced hardening may increase the damage caused by mean stress. This effect is
observed only at low strain amplitudes, as reported in an earlier paper by Taheri et al. [79] using
a Smith-Watson-Topper model, which does require an elastic-plastic constitutive law. In work by
Kamaya [80], discussed in KG 11, it was shown that both pre-strained and ratcheting tests
showed similar fatigue lives for the same effective strain range and that the reduction in fatigue
life could be conservatively predicted by assuming the effective strain range to be equal to the
total strain range (by assuming the crack mouth was never closed during the fatigue tests).

5-22
13321451
Assessment Methods

Asada et al. [39] demonstrated that, for austenitic steels, the fatigue life with mean strain was
almost the same as that with no mean stress, and the difference between the Modified Goodman
and other mean stress correction approaches was not significant. Therefore, the DFC
Subcommittee in Japan did not adopt an approach for mean stress correction for austenitic steels.
In contrast, the use of the Smith-Watson Topper approach was recommended for carbon and low
alloy steels.
Summary
The influence of mean stress (or strain) on fatigue life is complex in stainless steels, due to the
sequential hardening, softening and secondary hardening behavior that occurs, dependent on
composition. Test results depend critically upon the way in which tests are conducted. All the
data discussed above are for tests in air, and there is still a need to understand the effects of mean
stress during tests in water. However, the major effects are observed at low strain amplitudes
(i.e., in the high cycle fatigue region), where environmental effects are expected to be minimal.
Work in both high-temperature air and water is underway within the INCEFA+ project (see
Section 3). Therefore, KGs 6 and 44 should remain open, pending the results of the INCEFA+
tests and other ongoing work.

5.4 Fatigue Crack Growth

5.4.1 Knowledge Gap 23 (Previously Classified as Category C, High Priority)


Knowledge Gap Definition
ASME XI does not include a fatigue crack growth law for wetted flaws [in austenitic stainless
steels]. A Code Case (CC) has been proposed based on an extensive database but it is not
currently incorporated in the Code.
Research and Development Need
There is a need to follow ASME CC developments and consider the need for further data to
support development of the draft CC.
Ancillary Information from the 2011 KG Report
Chapter Section Subsection(s) Page(s) Applicable Materials
Multiple Multiple Multiple 2-12 to 2-15, 4-2 to 4-3, 9- All
17 to 9-19, 10-4

Background
Crack growth rates for austenitic stainless steels in both PWR primary and BWR coolant
environments are known to be significantly faster than those in air based on a substantial body of
test data. KG 23, and the discussion below relates specifically to PWR environments for which a
crack growth reference curve is provided in ASME Code Case N-809. No reference curve is
currently available for BWR environments (either NWC or HWC); however, EPRI has initiated a
project in 2018, the objective of which is to revise CC N-809 to add BWR environments. This is
the subject of KG 24, which is discussed in Section 4.2.2.

5-23
13321451
Assessment Methods

Review of Activities Relevant to KG 23


ASME Code Case N-809 [5] has now been issued and provides guidance and FCG laws for
austenitic stainless steels for wetted flaws in a PWR environment; the derivation of the model is
described by Cipolla and Bamford et al. [159]. The model is based on a large data set, mainly
from the US (Mills et al.), UK (Tice et al.) and Japan (Nomura et al.), which was then analyzed
by Mills [109]. Code Case N-809 uses equations very close to those developed by Mills which
include influences of ∆K, stress ratio, loading rate and temperature. The data underlying the
Mills model and the reference curves show an Arrhenius dependence from 150 °C to about
340 °C, with crack growth rate increasing with temperature which is reflected in the model, but
the model includes an upturn between 150°C and ambient temperature to reflect a small number
of Japanese data points. This latter behavior appears inconsistent with more recent data which
does not show such behavior [67].
Example calculations using the model are also reported in a second paper [159], in which its
application to a reactor coolant pipe-to-nozzle weld was considered. The calculations indicate
substantially greater crack extension over a given number of cycles than the previous model in
ASME Section XI Appendix C (where a factor of 2 was applied to the ASME air curve for PWR
conditions). This is not unexpected because the model recognizes that substantial levels of
enhancement can occur in PWR water, especially at longer rise times. Comparisons with an
earlier model in the JSME Code are also provided. The paper acknowledges that there are several
refinements that could be made to the model to make it more realistic (and less conservative). In
particular, a sensitivity study on the effect of ∆K threshold on crack extension indicated that
increasing its value from the current 1.1 MPa√m to 1.5 or 2 MPa√m significantly reduced total
crack growth, and it was suggested that higher thresholds may be applicable at lower R ratios.
Recent data do appear to indicate that higher threshold values may be appropriate [108], as
discussed in the context of KG 22 (Section 4.2.1). There is currently on-going activity within
ASME Section XI committees to revise the K threshold value in CC N-809.
Negative R-ratio effects are considered by Watson et al. [137], and guidance is provided for FCG
calculations on stainless steel components based on the observation that only a fraction of the
increasing portion of the compressive stress region of a transient appears to contribute to crack
growth. Environmental effects on R-ratio are discussed and oxide-induced crack closure is
proposed as a potential reason why a reduction in the effective stress intensity factor range for
cracking might be expected. This phenomenon is reported as likely to reduce the FCG rate. Some
ongoing testing is underway to underpin this judgment, with preliminary results reported in
[138]; see further discussion under KG 28. Therefore, a gap still exists to confirm to what degree
oxide on crack flanks influences the effective stress intensity factor range. There is currently on-
going activity within ASME Section XI committees directed at revising the crack growth
relationship in the compressive loading region in CC N-809.
A variation of CC N-809 is given in Reference [139], which rearranges the FCG equation to
enable the rate of change of stress intensity factor (K) with respect to time to be entered into the
crack growth law. This mathematical manipulation provides the ability to weight the portions of
the cycle according to the K rate, because it has been demonstrated experimentally using multi-
stage loading waveforms that a slow rise time in the higher K region of the cycle has a greater
effect than if it is in the lower K region [133]. This methodology has been shown to reduce

5-24
13321451
Assessment Methods

conservatism when treating long rise times in the lower portion of a load cycle. Further testing at
different temperatures and load ratios will be required to provide the necessary validation of the
model. Further details of the modeling approach and the supporting test data are provided under
KG 28.
Summary
ASME XI Code Case N-809 was published in 2015 and provides reference fatigue crack growth
curves for austenitic stainless steels in PWR environments. The crack growth relationship in CC
N-809 provides a reasonable description of published crack growth rates from laboratory tests;
additional data not considered in the database from which the curves were derived are also
identified in the basis document for future consideration. Further developments are under
consideration, including improved definitions of threshold and R-ratio effects, including
behavior under negative R loading. It is also judged that the change in temperature dependence
in the current curves warrants further consideration. This KG should remain open, but be
combined with KG 22 (∆K thresholds) and KG 24 (CG for stainless steel).

5.4.2 Knowledge Gap 28 (Previously Classified as Category A, High Priority)


Knowledge Gap Definition
Very few [crack growth] data are available under plant-representative loading conditions, and the
influence of complex loading conditions (including hold times and spectrum loading) waveforms
and combined loading are not well quantified. Crack growth data are obtained under isothermal
conditions whereas many facility plant transients involve simultaneous temperature and load
cycling (either in-phase or out-of-phase).
Research and Development Need
Further tests are required to investigate the influence of different loading waveforms on crack
growth rates and better represent the temperature and loading conditions experienced in actual
plants.
Ancillary Information from the 2011 KG Report
Chapter Section Subsection(s) Page(s) Applicable Materials
10 Laboratory Data Crack Growth Data 10-8 All

Background
KG 28 specifically regards crack growth data collected under plant-relevant loading conditions.
However, sufficient advancement has occurred that a new methodology has been developed that
requires consideration from the assessment community. Potentially a new Gap or modification of
an existing Gap is required to properly handle new methods of this type. To ensure that this
information is presented to the correct audience, KG 28 has been split so that the data is dealt
with in Section 4 and the assessment method is considered here. A more optimal place for this
Gap will be suggested in a later Section of this document.
As noted elsewhere, plant transients differ significantly from the continuous triangular or sine
wave cyclic loading waveforms, which are typically applied during laboratory testing under
isothermal loading, usually at positive stress ratio. The need was therefore identified for
additional data to clarify the impact of complex loading conditions, which include hold times,

5-25
13321451
Assessment Methods

spectrum loading, non-isothermal loading, waveform shape and negative R loading. Details of
recent crack growth studies in this area are described in Section 4.3.1. As a result of new
experimental observations in this area, it has become clear that existing methods of calculating
damage due to EAF in this area are inadequate and often excessively conservative for many plant
transients, especially those that are thermal in origin. Improved assessment methods have
recently been proposed and are discussed below.
Review of Activities Relevant to KG 28
FCG for austenitic stainless steels in deaerated PWR environments for use in ASME Section XI
flaw evaluation applications is contained within Code Case N-809 [5]. The FCG law is
dependent on temperature, R ratio and environment. Reference [104] discusses temperature
effects during environmentally-assisted fatigue crack growth. It highlights the increase in crack
growth rates as the temperature exceeds 150 °C, but also points out that there is a departure from
this trend when the temperature falls below 150 °C. Recent work [108; 110] also suggests that
threshold ∆K values also appear higher than the value of 1.1 MPa√m assumed in Code Case N-
809, with values between 2 and 5 MPa√m being proposed (see Section 5.4.1).
Emslie et al. modified the Code Case N-809 relationships to develop an approach for fatigue
crack growth under complex loading using the weighted K-rate (WKR) model [139]. The WKR
method accounts for the rise time effect by expressing the FCG law in terms of a K rate rather
than rise time and weights the K rate according to the position within the load cycle based on the
exponent in the FCG law. This approach was based on the test data investigating loading
waveforms and spectrum loading performed by Platts et al. [133; 136]. The basis of the model is
the experimental observation that the damage through a complex loading cycle is not constant
but is weighted towards the upper portion of the rising load portion of the cycle. The proposed
WKR model, therefore, expresses the fatigue crack growth law in terms of the rate of change of
Stress Intensity Factor, K, along with the K range, ∆𝐾𝐾. This is derived from the crack growth law
in ASME Code Case N-809 which expresses it in the following form:
𝒅𝒅𝒅𝒅
= 𝑪𝑪𝟏𝟏 𝑻𝑻𝟎𝟎.𝟑𝟑
𝒓𝒓𝒓𝒓𝒓𝒓𝒓𝒓 ∆𝑲𝑲
𝟐𝟐.𝟐𝟐𝟐𝟐
Eq. 5-6
𝒅𝒅𝒅𝒅

where a is the extension in crack depth, N is the number of applied load cycles, 𝑇𝑇𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟 is the load
rise time, ∆𝐾𝐾 is the stress intensity factor range, and 𝐶𝐶1 is a constant that is calculated from the
R-ratio, environment and temperature. Equation 5-7 is obtained by Equation 5-5 and Equation 5-
6 to give an expression that contains the rate of change of K during the rising load portion of the
cycle:
∆𝑲𝑲
𝑲𝑲̇𝒓𝒓𝒓𝒓𝒓𝒓𝒓𝒓 = Eq. 5-7
𝑻𝑻𝒓𝒓𝒓𝒓𝒓𝒓𝒓𝒓

𝒅𝒅𝒅𝒅
= 𝑪𝑪𝟏𝟏 𝑲𝑲̇−𝟎𝟎.𝟑𝟑
𝒓𝒓𝒓𝒓𝒓𝒓𝒓𝒓 ∆𝑲𝑲
𝟐𝟐.𝟐𝟐𝟐𝟐
Eq. 5-8
𝒅𝒅𝒅𝒅

Using the revised formulation shown in Equation 5-7 enables the degree of environmental
enhancement to be related to the rate of change of K during the rising load portion of the
waveform (K̇ rise ) or K rate. Thus, the K rate is weighted depending on its position in the
waveform as observed in the tests reported by Platts et al. [133; 136]. The analysis of these data
demonstrated that the crack growth rates were strongly dependent on the position of the K rates
within the waveform. For example, a long rise close to minimum load did not significantly
contribute to the environmental enhancement. The approach assumes that most of the crack

5-26
13321451
Assessment Methods

extension occurred in the rising portion of the waveform, an assumption confirmed by recent
testing at Amec Foster Wheeler [133], and that it was an incremental process. Using these
assumptions, Emslie et al. stated that crack extension would be consistent with the ΔK exponent
in the crack growth law which enabled a relationship between applied load and crack extension
to be developed in which the load in the early stages of a rise (bottom of the waveform) will
contribute much less to crack extension than the load being applied during the later stages of a
rise (top of the waveform). As indicated by the equations above, the WKR method gives the
same rates for sawtooth loading as the fatigue crack growth law described in CC N-809. The
WKR method was also found to give a good prediction of crack growth rates for a range of
complex waveforms, which were not well described by the use of Code Case N-809 with an
average rise time. The WKR method, therefore, allows the impact of a waveform shape to be
accounted for in crack growth rate calculations, and so enables the removal of potential
conservatism when applying Code Case N-809. Further validation of the model has been applied
to additional data including limited data obtained under non-isothermal conditions [67].
In the case of non-isothermal loading, only limited data are available for crack growth [136].
Nevertheless, the presently available data reinforces the conclusion of Green et al. [156]
recommending that - for simple assessment methods - the average metal temperature of the cycle
should be used for fatigue assessments. The data described in Reference [136] indicated that the
crack growth rates for out-of-phase loading corresponded to the rate found for isothermal tests
conducted at the lowest temperature of the transient cycle. For in-phase loading, the crack
growth rates were found to be close to those measured during isothermal tests at the highest
temperature of the transient cycle. Thermo-mechanical fatigue endurance data, generated since
2012, showed the following general correlation between the fatigue lives for in and out-of-phase
loading: Fatigue Life using Maximum Temperature < Fatigue Life for In-Phase Loading <
Fatigue Life for Out-of-Phase Loading < Fatigue Life using Minimum Temperature. This
correlation agrees with the conclusions presented for the crack growth data which suggests the
longest lives should be found for the out-of-phase tests.
Summary
Although the reference fatigue crack growth curves for austenitic stainless steels in PWR
environments provided in ASME XI Code Case N-809 provides a reasonable description of
published crack growth rates from laboratory tests, subject to further developments discussed in
Section 5.4.1, it nevertheless substantially overestimates crack growth for complex and
non-isothermal transients. Improved methodologies have recently been published which appear
to provide a much better predictive capability for plant transient loading. Further development of
these methods is warranted and it is therefore considered that the assessment aspects of KG 28 be
combined with other work relating to improved assessment methods for crack growth in stainless
steel (see KG 23, Section 5.4.1). Therefore, KG 28 should remain open as further data is
required to underpin this new method. Additionally, the analysis side of this Gap should be
considered for relocation.
It is also appropriate to evaluate the potential relevance of these improved assessment
methodologies to stainless steels in BWRs (see Section 4.2.2) and to other materials such as
carbon/low alloy steels and nickel-based alloys (Section 4.2.5). This will need to be supported by
experimental data for complex transients and non-isothermal conditions for these materials.

5-27
13321451
Assessment Methods

5.5 Welds

5.5.1 Knowledge Gap 38 (Previously Classified as Category C, Medium Priority)


Knowledge Gap Definition
While there are some data concerning the behavior of welded features, there is a lack of data to
account for aspects such as geometric stress concentration factor, weld defects, residual stress,
and multiaxiality, all of which may be influential.
Research and Development Need
There is a general need to understand whether weld related features can lead to differences in
EAF behavior compared to parent materials.
Ancillary Information from the 2011 KG Report
Chapter Section Subsection(s) Page(s) Applicable Materials
9 Other Design Code Treatment of 9-12 to 9-13 All
Aspects Weldments

Background
The analysis approach used by ASME Section III assumes that the fatigue lives of weld metal are
bounded by (equal to or longer than) those for parent material. The validity of this assumption is
the subject of KG 37 which is discussed in Section 2.2.2. Welded features also have
characteristics which can make them more susceptible to fatigue cracking. Transitions between
weld material and parent material often have a discontinuity that acts similarly to a notch. ASME
Section III also specifies the use of a stress concentration factor to account for the influence of
geometric and other features associated with a weld on fatigue life, which is greatest for an
undressed field weld.
There is also a region next to the weld which has different microstructure due to the extreme
heating that the base material experiences during the welding process. This heat affected zone
(HAZ) has different material properties from the base material remote from the weld and hence
different fatigue properties. The weld material normally has higher strength compared to the
parent material, which may suggest improved fatigue endurance. However, the lower fracture
toughness that is normally associated with welds could limit the amount of allowable crack
growth based on the reduced limiting defect size. Fatigue endurance data include initiation, as
well as short and long crack growth. Therefore, to fully understand the influence of weld
material on fatigue endurance, the detailed history of the welding process is required. If this is
not available, there are simplified methods which provide correction factors to account for welds
by modifying the S-N curve.
Residual stress is also a common feature of welds and this would be particularly relevant to high
cycle fatigue where applied stresses are below yield and mean stress correction is a factor. For
low cycle fatigue, the stresses exceed the yield strength of the material and may cause some
reduction in this residual stress. However, evaluation of weld residual stress is not a
straightforward assessment, and requires detailed finite element analysis to estimate the strain.
Appendix 4 of the R5 procedure [25] also provides specific guidance on the treatment of
weldments without the need for more complex finite element analyses. Residual stress also

5-28
13321451
Assessment Methods

influences the mean strain when considering fatigue endurance and R-ratio for crack growth
assessments. Both these aspects can be assessed with current methods for mean stress correction
and R-ratio calculation.
Review of Activities Relevant to KG 38 and Summary
Reference [160] outlines the main aspects of ongoing research in Germany on environmental
influences on the fatigue assessment of austenitic and ferritic steel components including welds.
A key focus of the work is to establish a practicable engineering assessment procedure,
maintaining design code methodologies as far as possible. This procedure will be underpinned by
an extensive experimental program that will investigate the influence of microstructure on EAF.
Preliminary investigations have identified the transition between austenitic steel and weld
material to show the lowest fatigue life in air [161].
ASME Section III specifies the use of a stress concentration factor to account for the influence of
geometric and other features associated with a weld on fatigue life, which is greatest for an
undressed field weld.
The general approach to analyze weldments is to treat them as though they are made up of parent
material, and the difference in behavior of the weldment compared to the parent material is taken
into account by using a Fatigue Strength Reduction Factor (FSRF). These FSRFs are derived
from fatigue tests on actual weldments by comparing the observed weldment endurance with the
parent material fatigue endurance curve. In R5 Volume 2/3 (2012) [162], two separate
assessment routes were provided for weldments in the undressed (as-welded) and dressed
conditions using different FSRFs to enhance the strain range to reflect the reduced endurance of
the weldment. For undressed (as-welded) weldments, where the detailed surface geometry is not
known, the FSRF was also used as a strain concentration factor in determining the stress.
However, Dean et al. [163] considered this approach to be overly pessimistic since the FSRF
includes the material fatigue endurance reduction effect, which should not affect the stress.
The new R5 procedure (2014) [25] was also simplified by adopting a single route for both
dressed and undressed (or as-welded) weldments. In Appendix A4 of the R5 Volume 2/3
procedure, FSRFs were replaced by Weld Strain Enhancement Factors (WSEFs) and a Weld
Endurance Reduction (WER). The WSEF accounts for the strain enhancement due to the
weldment geometry and the material mismatch between the weldment zones, while the WER
accounts for the fatigue endurance reduction due to the presence of small imperfections (e.g.,
inclusions, porosity etc.) in the weldment constituent materials. For calculation of Fen, the WSEF
would not be imposed in the stress calculation as this would over-estimate the strain rate.
The procedures outlined in other codes, such as KTA [17], specify different factors and values to
use to account for welds. The CEN Design standard for unfired pressure vessels EN 13445 [21]
includes different fatigue curves for parent steel and weld metal. The curves for welds differ
depending on the class of welded joint, which range from full penetration butt welds (least
reduction in life compared to non-welded materials) to fillet welds (greatest reduction in life).
Progress against this Gap appears to be limited by the lack of significant advancement in the
amount of supporting data. Therefore, this KG should remain open.

5-29
13321451
Assessment Methods

5.6 Geometric Features

5.6.1 Knowledge Gap 40 (Previously Classified as Category C, Medium Priority)


Knowledge Gap Definition
The method of stress indices commonly used for simplified piping analysis requires further
development for corrosion fatigue assessments.
Research and Development Need
Further testing and analysis work is required to develop and validate simplified assessment
methods for specific components.
Ancillary Information from the 2011 KG Report
Chapter Section Subsection(s) Page(s) Applicable Materials
9 Other Design Code Component Specific 9-11 to 9-12 All
Aspects Rules

Background
In addition to the full assessment methodology of the design codes, simplified rules (using stress
indices) can be used for certain components such as piping. Also, for all components, the vessel
rules for determining the sign of the strain rate can be used. Additionally, for some component
types, specific rules relate to the way the strain rate may be calculated and signed (positively or
negatively) using component-specific simplifications. Both the vessel and component-specific
rules represent an increase in the calculation complexity when consideration of corrosion fatigue
is required. For piping, the JSME simplified rules and ASME NB-3600 [1] use the simplified
method of stress indices, which generally does not require time histories of stress or strain. For
the EFEM, the method of stress indices is retained with the additional requirement of an analysis
of the strain rate history. The method of stress indices introduces complications as the specified
stress index is appropriate to the position of maximum stress in a pipe cross section,
corresponding to the maximum strain rate, whereas the Fen is maximized at the position of
minimum strain rate. Consequently, it is not evident in which region the maximum EAF occurs.
For example, the test work on a pipe bend of Reference [125] reported in Section 7 of Reference
[6] showed that both the location of initiation and crack orientation were different in air
compared with a PWR environment, with the failure occurring in the region of highest stress in
the former.
Review of Activities Relevant to KG 40 and Summary
Reference [164] calculates stress indices for a non-radial branch connection using non-linear
finite element analysis. A similar methodology could be employed to obtain best estimate, or
lower-bound values for the stress index to be used in the Fen calculation.
R5 considers geometric influences on strain amplitude via Fatigue Strength Reduction Factors
(FSRFs). However, these are not required if a full finite element analysis has been undertaken.
In RCC-M, the stress index table B3680 is being updated to align with ASME NB-3600 [165].
Progress against this Gap appears to be limited by the lack of significant advancement in the
amount of supporting data. Therefore, this Knowledge Gap should remain open.

5-30
13321451
Assessment Methods

5.7 Other Knowledge Gaps

5.7.1 Knowledge Gap 35 (Previously Classified as Category B, High Priority)


Knowledge Gap Definition
For many PWR and BWR plants, there is a lack of knowledge of actual plant transients, which is
important because of the sensitivity of EAF to temperature and strain rate variations.
Research and Development Need
Plant monitoring is required to obtain actual plant transient profiles, and/or detailed
thermodynamic modelling. Long-term plant monitoring has been successfully carried out in US
and German plants.
Ancillary Information from the 2011 KG Report
Chapter Section Subsection(s) Page(s) Applicable Materials
8 International Status N/A 8-2 to 8-3 All

Background
Plant transient data can be obtained from calculated temperatures and pressures, or directly from
instrumentation within the plant. Analysis techniques range from one-dimensional thermal
hydraulic calculations to very detailed three-dimensional Computational Fluid Dynamics (CFD).
The one-dimensional codes provide data on the average temperature across the section of the
component, and do not account for any local changes in fluid properties due to complex
geometry or mixing. This leads to uncertainty in the actual metal temperatures at certain regions,
and in some cases more complex analysis using CFD is justified. Where detailed plant transient
data is available, significant improvements in strain amplitude calculation and transient cycle
counts can be obtained.
Review of Activities Relevant to KG 35 and Summary
Wermelinger et al. presents a new fatigue monitoring system implemented by KernKraftwerk
Gosgen-Daniken (KKR) which utilizes thermocouples surrounding primary components [166].
The quality of the data which is obtained from this new system is reported to provide more
accurate fatigue assessments and result in highly stressed components being kept in operation
longer than originally planned. Thermocouples were also installed by E.ON at fatigue relevant
locations within the primary circuit [167]. The more detailed transient data obtained from these
measurements has the potential to improve operation with particular regard to component
fatigue. Westinghouse has implemented a method to automate transient cycle counting, to more
easily evaluate the impact on fatigue life from testing of components [168]. This information has
provided the plant engineer with the justification necessary for continued operations without
significant expense. EPRI and Structural Integrity Associates have performed fatigue monitoring
in operating plants since 1986 [169; 170; 171; 172].
The methods presented above show benefit can be gained in terms of plant life when more
accurate transient monitoring systems are installed. This was also concluded by the NRC in
NUREG/CR-6260 during their initial investigations of the impact of EAF on plant components
[173] This is particularly relevant to aging plants where more refined assessments are often

5-31
13321451
Assessment Methods

required to justify continued operation. More benefit would also be gained by using this
information in laboratory tests on plant-realistic waveforms or on EAF testing of actual
components. This would allow further development and validation of existing and emerging
assessment methods. Therefore, KG 35 should remain open.

5.7.2 Knowledge Gap 45 (Previously Classified as Category C, Medium Priority)


Knowledge Gap Definition
There is a lack of understanding of the extent to which the inclusion of environmental effects on
fatigue will influence inspection programs with regard to how to inspect, when to inspect, and
where to inspect.
Research and Development Need
A review of the implications of corrosion fatigue on inspection requirements is required.
Ancillary Information from the 2011 KG Report
Chapter Section Subsection(s) Page(s) Applicable Materials
10 Discussion of Inspection Requirements 10-11 to All
Knowledge Gaps and Mechanistic 10-12
Understanding

Background
Plant inspection requirements use fatigue assessments to guide the frequency and extent of in-
service component inspections (and maintenance). EAF has a marked impact on fatigue initiation
and crack growth and, for that reason, may influence the inspection requirements. The rankings
used to prioritize the components in an inspection program are typically based on air fatigue
assessments, but it is unclear if these rankings and, in fact, locations will still be valid when EAF
is active. To determine what effect EAF may have on inspection programs a deeper
understanding of EAF mechanisms will be needed.
Review of Activities Relevant to KG 45
Inspection requirements for nuclear components are outlined in the ASME Code Section XI
[174]. Nonmandatory Appendix L provides flaw tolerance evaluation procedures which inform
the inspection intervals based on allowable crack growth. Environmental enhancement of these
fatigue crack growth rates in a PWR or BWR environment puts more onerous requirements on
these inspection intervals and hence increases the effort in obtaining license renewal.
Nonmandatory Appendix L provides methods for performing fatigue assessments to determine
acceptability for continued service of reactor coolant system and primary pressure boundary
components and piping subjected to cyclic loadings. This nonmandatory appendix is applicable
only in the absence of any flaw at the location of concern which is larger than allowed by the
applicable acceptance standard referenced in Table IWB-3410-1.
Appendix L refers to Nonmandatory Appendix C, which gives detailed guidance on determining
the end of life flaw size as well as the tolerable flaw size. The tolerable flaw size is based on
linear elastic fracture mechanics (LEFM) and proximity to plastic collapse concepts. An end-of-
life flaw size is calculated according to the geometry, transient specification, and relevant crack

5-32
13321451
Assessment Methods

growth laws. Two recent examples where Nonmandatory Appendix L was successfully applied
in the license renewal process was Turkey Point Units 3 and 4 in 2013 and St. Lucie Units 1 and
2 in 2016. Both cases addressed pressurizer surge line inspections due to high calculated values
of environmentally assisted fatigue usage. Four approaches were considered:
1. Further refinement of the fatigue analysis to lower the CUF below 1.0, or
2. Repair the affected locations, or
3. Replacement of the affected locations, or
4. Manage the effects of fatigue by an NRC approved inspection program.
Option 3 could also involve a design modification of the replaced component to reduce the level
of fatigue damage. However, it is recognized that this is likely only to be practicable for new
plant designs.
The extent of and basis for exceeding the fatigue usage for the susceptible regions of Turkey
Point Units 3 and 4 and St. Lucie Units 1 and 2 are not detailed in the NRC submittals [175] and
[176], i.e., it is not clear at what point in plant life the usage will be exceeded or whether the
primary cause of the fatigue initiation relates to environmental effects. While, in this case,
Option 1 may not be viable for these particular plants, it may be applicable in other cases
dominated by environmental effects. This would be the preferred option noting the advantages in
reduced dose burden in limiting through-life inspection and maintenance activities as well as
reduced plant shut down.
Of the remaining options, Option 4 is desirable as it reduces the potential to cause damage to the
plant, dose burden and plant downtime, and hence has been adopted in the plant life extension
justifications of Turkey Point Units 3 and 4 and St. Lucie Units 1 and 2, which are detailed
below. The Code Case N-809 crack growth laws for austenitic steels in a PWR water
environment were under development at the time of the Turkey Point submittal in 2013, so a best
estimate approach was used based on evaluation of crack growth data by Mills [109]. This data
analysis formed the basis for ASME Section XI Code Case N-809 which was published in 2015
[5] and used for the St. Lucie submittal in 2016 [176]. Both the flaw evaluations and crack
growth analyses utilized stresses obtained from elastic FEA.
Saint Lucie Aging Management Plan
A flaw tolerance evaluation was performed specifically for St. Lucie Nuclear Plant Units 1 and
2, in order to assess the operability of the pressurizer surge line by using ASME Section XI
Appendix L methodology and to determine the successive inspection schedule for the surge line
welds with a postulated surface flaw [175]. Two bounding locations were evaluated in detail.
The two bounding locations of concern were the hot leg surge nozzle-to-pipe weld and the
adjacent elbow base material, which is a Cast Austenitic Stainless Steel (CASS) material. The
bounding argument was based upon the fact that these locations saw the most limiting transients
in terms of EAF. The calculated crack growth rates and the allowable flaw size indicated an
allowable operation period of greater than ten years.
Following this ASME XI Appendix L assessment, the welds are to be examined in accordance
with the requirements of ASME Code, Section XI, Subsection IWB, “Requirements for Class 1
Components of Light-Water Cooled Plants,” every ten years.

5-33
13321451
Assessment Methods

The proposed surge line inspection program will rely on volumetric examinations to detect
cracks that may be formed because of EAF. It was determined by NRC staff that the use of
volumetric examinations, such as ultrasonic and radiographic tests, is acceptable because, when
performed in accordance with ASME Code Section XI, these examinations are capable of
detecting discontinuities, such as cracks, that initiate from the inside diameter of welds in piping
[176].
Turkey Point Aging Management Plan
A similar process was followed for Turkey Point Units 1 and 2 [177], where the bounding
locations of concern were the pressurizer surge nozzle-to-safe-end weld and the hot leg surge
nozzle-to-pipe weld. Again, in accordance with Appendix L, Table L-3420-1, for the allowable
operating periods obtained, the successive inspection schedule for pressurizer surge line welds
was determined to be ten years for either an axial or circumferential postulated flaw.
In the NRC’s review of the licensee’s submittal, additional information regarding the data
underlying the referenced fatigue crack growth curves of W. J. Mills was requested [178]. These
data, which now form the basis of ASME Section XI Code Case N-809, show the rise time
dependency of crack growth rates for different stress intensity factor ranges (ΔK) and R-ratios.
The crack growth rates obtained from the W.J Mills formulation are more conservative than
assuming twice the crack growth rate in air. Given that twice the air rate has been previously
accepted in industry for PWR conditions, and the inspection intervals are shorter than the time to
reach a limiting flaw size, the NRC determined that the licensee’s approach was conservative.
Summary
For component sites where CUF values are greater than 1.0, the application of Nonmandatory
Appendix L from ASME XI to the St Lucie Units 1 and 2 and the Turkey Point Units 3 and 4,
using the modified environmentally enhanced fatigue crack growth laws to support the in-service
inspection program, has been accepted by the NRC. However, the preferred option would be to
demonstrate that the CUF remains less than 1.0 or, where this is not possible, to develop a more
accurate understanding of the level of crack growth so that in-service activities, which have a
number of disadvantages, can be minimized. This is the focus of the other activities discussed in
this report. It is acknowledged that other international codes and regulators may have different
requirements in terms of the acceptability of the in-service inspection and maintenance
requirements. Overall, recognizing that the in-service inspection and maintenance activities are
dependent on the progression of the other activities concerning environmental fatigue initiation
and crack growth, as well as the different requirements for these activities across the
international community, it is considered that KG 45 covers too broad a subject area to be
included in a review of EAF knowledge gaps and hence can be considered closed.

5.8 Summary
This subsection summarizes the key observations made in this section of this report.
There has been significant progress made in the last 6 years in terms of generating new data for
design fatigue curves, crack growth rates and an improved mechanistic understanding of EAF. It
will be the combination of these three developments that eventually leads to new semi-empirical
Fen factors and crack growth laws that align estimates of fatigue lives with actual field
observations. As environmental fatigue is very much dependent on the rate of increase of strain,

5-34
13321451
Assessment Methods

and plant transients have variable strain rates, modern calculation methods have been designed to
account for such variations.
With respect to calculations of fatigue crack initiation, methods that integrate across the relevant
portions of a loading cycle, accounting for changes in temperature and strain rate, as well as
weighting the Fen more in the most damaging, i.e., maximum tensile portion of the cycle, are
showing the most promise for more accurately predicting the fatigue lives of actual components.
More data is required to fully underpin these models and improve their predictive capability. A
focused international effort to generate the necessary test data would be helpful in achieving this
desired outcome.
For the calculation of fatigue crack growth, a method that accounts for the rate of change of the
crack driving force, as well as the varying temperature throughout the loading cycle, provides the
most benefit in terms of reducing unnecessary conservatism in plant assessments. An updated
ASME crack growth law is now available for austenitic stainless steels in PWR environments,
and application of the WKR method which can account for varying increases in stress intensity
factor through the cycle shows promising trends. Furthermore, the method can account for
varying temperature throughout the loading cycle, which has significant benefits for cases when
the temperature is out-of-phase with the applied strain.
One key revision to the RCC-M code document was proposed by Métais et al. [48] and suggests
there should be a reduction in the reserve factor on stress to generate the stainless steel design
curve. Factors of 10 on life and 1.4 on strain were proposed [44] based on statistical analysis of
French data, as well as a method to account for the environmental effects already covered by the
fatigue design curves. The rules allow previous calculations of CUFs using air design curves to
be valid as long as the Fen is below a certain value for a certain transient type. For example,
according to these rules, a thermal shock transient can have a Fen of up to five before a more
detailed analysis needs to be performed and additional Fen values applied. Work is ongoing to
support these new rules and both methods are integrated as Règle en Phase Probatoire (RPP) to
the RCC-M code [37].
Further work is also proposed for assessing environmental effects in welds [160; 161] and will
provide further data to improve current assessment methodologies. The KGs associated with
welds (37, 38 and 40) are still open due to the ongoing work in this area.
Transient definitions using plant monitoring and analysis has been discussed under KG 35. The
precise definition of transients through improved instrumentation and detailed thermal hydraulics
calculations could significantly reduce the uncertainty in plant assessments. Given that Fen is
strongly dependent on the temperature and strain rate, any improvements in this area could
provide considerable benefit.
Additionally, concern has been raised over the new limits applicable to the Fen expressions in
NUREG/CR-6909 Rev 1 [3], in particular, the maximum threshold relating to the strain rate (2.2
% s-1, 7.0 % s-1, and 5.0 % s-1 for CS/LAS, SS, and Ni-alloys, respectively). Having this limit
set so high means that the vast majority of plant-relevant transients and laboratory testing will
still be subject to a Fen when the other parameters are within the limits for an environmental
effect. This could lead to excessive conservatism being applied to transients that have rates
greater than 0.4 % s-1 where there are no data to justify an environmental effect.

5-35
13321451
Assessment Methods

Notable progress has been made against KG 45 during the license renewal of Turkey Point
(Units 3 and 4) and St. Lucie (Units 1 and 2). In these two cases the NRC accepted in-service
inspection programs for component sites where the CUF was greater than 1.0. By using a
combination of crack growth assessments and non-destructive crack detection methods, the
operators were able to demonstrate adequate conservatism. Although progress has been made
against understanding the impact of EAF on in-service inspection and maintenance programs,
there are still uncertainties to be considered. Alternative codes and standards may have different
requirements that could mean that this route to determining maintenance and inspection periods
may not be generally accepted. Furthermore, there is significant benefit, in terms of safety and
dose, in improving current Fen approaches to maintain CUF values to less than 1.0.
Improvements in these approaches could reduce the number or extent of in-service inspections.
Given that the resolution of KG 45 will depend on the consensus of the international community
on appropriate assessment methodologies as well as being dependent on the progression of other
EAF activities, KG 45 may be too ambiguous to address within this document and so has been
deleted.
Most of the KGs discussed in this section of the report relate to Hypotheses 1 to 6. KG 2 applies
to all hypotheses as defined in the Roadmap document [7]. The KGs in this assessment section
were all given either high or medium priority by the EPRI Expert Panel. However, a significant
number of these Gaps were later reprioritized by the EPRI focus group, in some cases from high
to low priority. In general, the gaps that were downgraded to low priority by the EPRI focus
group related to Hypotheses 5 and 6. However, the effort put into addressing the KGs discussed
in this section by the nuclear industry more closely reflects the prioritization performed by the
expert panel for the original 2011 Gap. This indicates that, since the EPRI focus group assessed
the priority of the Gaps, a shift in the needs of the industry has occurred. This section
summarizes the advancements made towards the high-priority gaps as defined in the Roadmap
report.
Of all the Gaps considered within this section, KGs 2, 28 and 43 have matured the most since the
last update in 2012. KG 2 is a high-priority Gap, and is directly related to EdF’s Fen -integrated
approach that has been accepted as an RPP in the RCC-M code and, as Fen-threshold, is currently
being considered for acceptance as an ASME Code Case. This attempts to account for some
aspects of EAF that are already accounted for in the current fatigue design curves. Other
alternatives to NUREG/CR 6909 are currently under development. These developments include
the total life approach, which characterizes a component’s life from an understanding of fatigue
initiation followed by crack growth, which includes compensation for decaying strain fields
lowering the crack driving force as a crack propagates further into a component.
KG 28 is a high-priority gap discussed in this section to highlight the accomplishment of the
WKR method for predicting the crack growth rates that rise from complex waveforms. This
methodology is a significant improvement over existing methods for handling the impact of
complex waveforms on crack growth rates. The WKR method is still under development, with
the goal of extending its applicability to non-isothermal waveforms. This method is the direct
result of a sufficiently mature testing program that specifically targeted the resolution of this
matter and highlights the strength of collaboration between the testing and assessment
communities.

5-36
13321451
Assessment Methods

KG 43 is a medium priority Gap in Hypothesis 1 that was downgraded by the EPRI focus group
to low priority. In this section, several related Gaps (39, 41 and 42) were combined with this Gap
and addressed together as they all relate to how variable parameters are handled in assessments.
The major advance in this area is the publication of the SNW method that can adequately predict
the fatigue lives of specimens tested with complex and non-isothermal waveforms when used in
place of the MRA for determining the strain rate and temperature parameters for use in the Fen
equation in NUREG/CR-6909. This is a new method that was published in 2017 that requires
further validation against a larger range of waveforms.
Overall, large amounts of progress have been achieved against some of the Gaps summarized in
this section. However, in some cases, the advancement has been in an area that was previously
assigned a low priority. These lower priority Gaps have been advanced as a result of the industry
identifying opportunities to make significant reductions in over-conservatisms by tackling these
Gaps now.

5-37
13321451
13321451
6
SUMMARY OF OBSERVATIONS AND PROPOSED
UPDATE TO KNOWLEDGE GAPS

6.1 Introduction
This overall aim of this report is to provide an update to the KG analyses on EAF in LWR
environments which were published by EPRI in 2011-2012 [6; 7]. The original 2011 KG report
(1023012) identified areas where significant uncertainties regarding the influence of LWR
environments on both fatigue life and fatigue crack growth existed. The issues identified have
potentially significant consequences for the assessment methodologies proposed to address such
effects, such as those in NUREG/CR-6909, ASME Section III Code Cases for incorporating
LWR environment effects on fatigue life, and ASME Section XI Code Cases for fatigue crack
growth. The 2012 Roadmap Report (1026724) provided a number of hypotheses which, together
or in part, might provide an explanation for the observed discrepancies between laboratory
fatigue data in LWR environments and the exceptional service experience regarding LCF failures
in operating plants, together with Roadmaps which were aimed at guiding future work in this
area.
A thorough review of available published work, over the 6-year period since the completion of
the above reports, has identified that significant advances in the understanding of EAF in LWR
environments have been made, in response to major international research efforts on this topic.
This research has been carried out in response to perceived concerns about the relevance and
applicability of extant assessment procedures for fatigue in LWR environments described in
NUREG/CR-6909 and related NRC and other international regulatory guidance. As a result of
significant new experimental data and improved understanding of the factors influencing EAF, a
number of modified or alternative assessment procedures have been proposed or are in course of
development.
The KGs identified in the earlier EPRI reports provided valuable guidance to researchers with
respect to the scope of required EAF studies. Evaluation of more recent data and developments
in assessment methods and further research over the last six years has significantly improved the
understanding of many of the issues identified. As a result, some of the Gaps can be considered
resolved. Nevertheless, some issues are still outstanding, so a need has been identified to develop
a revised list of KGs to provide some guidance and focus on these remaining issues. It is also
apparent that it would be beneficial to merge some of the original gaps, and some clarifications
and additions are required. Based on the review of recent EAF studies summarized in Sections 2
to 5 of this report, this section provides an overview of the status of the existing KGs. A
simplified and updated list of KGs that reflects current understanding, together with new issues
and assessment approaches that have been identified since the original review, is provided in
Section 7.

6-1
13321451
Summary of Observations and Proposed Update to Knowledge Gaps

6.2 Status of Original Knowledge Gaps

6.2.1 General Issues


Discrepancies between predictions from EAF models and plant procedures
KGs 2 and 4 specifically identify that there is inadequate understanding of the reasons for the
differences between relatively good plant experience regarding fatigue failures and predictions
from the Fen-based procedures developed from S-N test data in NUREG/CR-6909. KG 2 states
that a wide-ranging investigation is required into the basis of the Fen (or similar) approach and its
application to plant transient analysis, whereas KG 4 suggests that excessive conservatism in
fatigue/EAF design rules and/or the influence of complex loading may provide at least a partial
explanation.
Recent work reviewed in Section 3 of this report and summarized in the remainder of this section
has identified a number of factors which contribute to the observed discrepancies between the Fen
model predictions and plant experience. Related studies concerning crack growth are described
in Section 4. Improved assessment methodologies have been proposed which provide significant
improvement in predictive capability for austenitic stainless steels in PWR environments; these
are discussed in Section 5. Further development and supporting data are likely to be required and
the applicability to BWR environments and other alloys need to be evaluated with appropriate
laboratory data.
Stakeholders’ testing database
KG 46, and some of the other KGs identified in the original gap analysis report, expressed the
need for further analyses of existing data, whilst other gaps highlighted the need for additional
data in a number of areas. Following a review of the KGs report by an expert panel of key
stakeholders, it was recommended that a comprehensive, and regularly updated, database of EAF
test data would be set up so as to enable the identification of specific data gaps and provide a
sound basis for international collaboration on EAF. No such database is yet available, although
international collaboration on EAF testing and assessment methodologies over the last few years
has resulted in some limited initiatives directed at data collation. However, there has been little
progress in establishing a database which is readily accessible.

6.2.2 Fatigue Endurance Data in Air and Fatigue Design Curves


Material Variability and Material-Specific Design Curves
KG 9 relates specifically to austenitic stainless steel and suggests that design curves based on
data for specific grades of stainless steel may be more appropriate in some cases. Several
national programs have examined such an approach, such those in Germany [14], Finland [179],
France [13] and Japan [11], and the NRC has recommended a similar approach [147]. The
French, German and Finnish studies for stainless steel have suggested that the mean air data line
underpinning the curve revised in ASME Section III in 2009 may be excessively conservative,
especially in the high cycle regime. However, the German work also showed that the fatigue
limit decreased with increasing temperature, so that there was little benefit above about 240 °C.
The Japanese work indicated that the fatigue limit for both austenitic and ferritic steels increases
with increasing tensile strength. This may provide a partial explanation for the heat-to-heat
variability in the fatigue limits of Types 304 and 316 grades of stainless steels reported by some

6-2
13321451
Summary of Observations and Proposed Update to Knowledge Gaps

researchers [15; 16]. Overall, the published air fatigue data on stainless steel do appear to show
that significant variability between the fatigue data for different material grades exists, but the
reasons are not adequately understood. Most of the recent data suggest that the effect is greatest
in the high cycle region, due to variability in the fatigue limit and there is some evidence that this
is related to tensile strength. This observation appears to extend to carbon and low alloy steels
and is recognized in both the Japanese DFC proposed design curves and in the CEN design
standard for unfired pressure vessels, EN-13445 [21]. The recent data on stainless steels, which
indicate a detrimental effect of elevated temperature on fatigue in air, are inconsistent with
NUREG/CR-6909 and the ASME Section III air design curve. Further work is therefore needed
to improve the understanding of the reasons for the influence of material variability on fatigue
life as well as the effects of temperature and material strength. There may also be a requirement
for studies on alloys other than stainless steel. The availability of all existing data, in the form of
a shared EAF database (refer to the discussion under KGs and 4 above), would be highly
beneficial for such an investigation.
Transference Factors for Design Fatigue Curves
The other main topic of importance concerning air fatigue design analysis identified in the KGs
relates to the transference (or adjustment) factors used to derive the design fatigue curve from the
mean data curve for each class of alloys. KG 12 addresses the factor on life, which applies at
medium to high strain amplitudes, whereas KG 3 relates to the factor of 2 on stress. For the
former factor, there is a discrepancy between NUREG/CR-6909, which uses a factor of 12 for all
alloys, and ASME Section III (2009 onwards) which uses 12 for austenitic stainless steel and
nickel-based alloys but retains the factor of 20 for carbon and low alloy steels. The reasons for
this difference are unclear. In addition, NUREG/CR-6909 defines separate curves for carbon and
low alloy steels, whereas the ASME Code retains one curve for both material groups. As
described in NUREG/CR-6909, the factor of 12 on life is made up of several contributing
factors, identified as material variability and data scatter, size and geometry, surface finish and
loading sequence. The main reason for the decrease in the factor from 20 to 12 is attributed to the
size factor. In addition, Revision 1 of NUREG/CR-6909 reviewed the contributions of each of
these factors again and concluded that an overall adjustment factor of 10 may be more
appropriate, but retained the factor of 12 in the recommended design curves to retain consistency
with the current ASME Code. It was also recommended\ that both the 2 and 12 factors be
revisited together before adopting a change in the factor of 12. Several more recent studies have
reconsidered these recommendations, with some variability in the conclusions drawn.
Significant differences in the effects of surface finish are observed between different studies and
between ferritic and austenitic materials, and the issue is further complicated by inconsistencies
between different parameters for describing surface roughness. More importantly, several studies
suggest that surface finish effects are lower in high-temperature water environments than in
ambient air which has implications for the use of an environmental factor, Fen, applied to the air
design curve; this aspect is discussed in Section 6.2.3. The Japanese DFC Subcommittee has
proposed an alternative approach to developing fatigue design curves to explicitly treat each of
the sub-factors, rather than combining them as in the ASME Section III and the
NUREG/CR-6909 approaches. The same philosophy is adopted in the CEN EN-13445 Code,
although the sub-factors are treated differently. There appears to be a reasonable consensus
regarding the value of the material scatter (2.1-2.8) and size effect (1-1.4) sub-factors, although

6-3
13321451
Summary of Observations and Proposed Update to Knowledge Gaps

there is some evidence that the size effect sub-factor may be negligible. Although a factor of
1.0-2.0 for loading sequence effects is included in the NUREG/CR-6909 adjustment factor, it is
possible that they may be better addressed by alternative approaches, as discussed for EAF in
Section 6.2.3.
In contrast to the transference factor on life, very little discussion of how the factor of 2 on stress
was derived is included in NUREG/CR-6909 because the authors did not address that factor;
rather, they retained its use to be consistent with the original ASME Code approach, but
recommended that it be revisited since a valid technical basis for the factor is absent.
NUREG/CR-6909 Rev 1 [3] included the basis for the factor of 2 on stress/strain and described
its scope in more detail, but reinforced the statement in the Rev 0 document that more data would
be required to substantiate the value currently used and to confirm its scope. Several proposals to
reduce this factor have been made. A French proposal to reduce the factor on stress for austenitic
stainless steel to 1.4 based on statistical analysis of datasets on a limited number of 304/316 (and
L) grades of stainless steel [31] appears to be supported by Finnish/German data on stabilized
321 and 347 grades [40]. The Japanese DFC have proposed a slightly lower factor of 1.34 for
stainless steel and a factor of 1.22 for ferritic steel. It is considered that the basis of these
proposed reductions require further validation. Once again, the availability of all existing data, in
the form of a shared EAF database (refer to the discussion under KGs and 4 above), would be
highly beneficial for such an investigation.
Weld Metal Data and Assessment
KG 37 relates to the relative paucity of fatigue data on weld metal but notes that existing data
appear to be bound by base metal data, as assumed in ASME Section III and NUREG/CR-6909.
Fatigue data on weld metal in LWR environments is especially limited. Very little additional
data have been identified to address this issue, but it is noted that EN-13445 [21] provides
separate design curves for base and weld metals, although the justification underlying the curves
is not provided. A more significant concern for weldments, however, is likely to be the influence
of geometrical features, weld defects and residual stress, which are covered by KG 38. Several
different approaches have been developed to address these issues, including stress concentration
factors (ASME Section III) which are defined for specific weld types (e.g., dressed and
undressed welds). The EN-13445 curves provide factors which allow for the type of weld (e.g.,
full penetration versus fillet weld). The 2014 version of the EdF Energy R5 Code uses weld
strain enhancement factors to address geometrical and mismatch effects, as well as a weld
endurance reduction term which accounts for an effect due to the presence of inclusions or pores.
There is therefore a need for further data, improved understanding and standardization of the
methods for dealing with welds in fatigue design. In weldments, especially those made using
dissimilar materials, the mismatch in strength can cause a local strain concentration which
reduces fatigue life; these features have been described as metallurgical notches [23] and can be
taken into account in a similar way to geometrical features. The effect is most severe in
conventional membrane-loaded fatigue tests on dissimilar weldments which can produce
premature failures which may not reflect behavior of many plant components.

6-4
13321451
Summary of Observations and Proposed Update to Knowledge Gaps

6.2.3 LWR Environment Effects on Fatigue Design


Data Deficiencies Defining Fen equations, Parameters and Thresholds
Several KGs identify data deficiencies which potentially impact assessment approaches for
specific classes of alloys, whilst others relate to specific areas of uncertainty which might vary in
significance for different alloy classes. KG 13 relates to the lack of sufficient data to demonstrate
that the Fen approach is applicable to the full range of relevant controlling parameters as
independent variables, and KG 10 identifies that the database underpinning thresholds for EAF
is incomplete. KG 46 identifies a need for a comprehensive, and regularly updated, database
which would enable the identification of specific data gaps and provide a sound basis for
international collaboration on EAF. The concept of an EAF database is a recurring theme that
manifests itself in many of the KGs, and its potential benefits to resolving KGs in the future
cannot be overemphasized. NUREG/CR-6909 Rev. 1 used a larger database to redefine the Fen
equations and redefined some thresholds, primarily so that the equations reduced to 1.0 for
situations when environmental effects are not present; details are provided in Section 3.2.2.
However, at the present time, much of the data underlying the analyses are not publically-
available. Nevertheless, international collaboration on EAF assessment methodologies has
improved significantly over the last few years. Further collaboration and data sharing is highly
encouraged.
For austenitic stainless steels, changes are detailed in NUREG/CR-6909 Rev. 1 [3] to the
temperature range over which EAF needs to be considered. The temperature threshold was
reduced from 150 °C to 100 °C, and the strain rate threshold was significantly increased from
0.4 % s-1 to 7 % s-1. Although the detailed technical basis for the latter change is not presented,
for most plant-relevant transient conditions, it effectively negates the benefit of a correction to
the Rev. 0 equations in Rev. 1. The revisions to the Fen equations makes it possible for a Fen
value of 1 to be obtained when EAF thresholds are not exceeded, rather than the minimum of
approximately 2 based on the earlier equations (see KG 1).
Some further uncertainties relating to stainless steel are identified in KG 20. These include the
known dependence of Fen on dissolved oxygen (DO) which is beneficial for stainless steel but
was excluded from the NUREG/CR-6909 Rev. 0 expressions. This was specifically identified as
a concern in KG 47. Although a DO dependence is now included in the Rev. 1 Fen equation for
stainless steel (except for sensitized material), it results in a step transition in Fen at 0.1 ppm DO,
which does not resolve the declared intent of KG 47 which was to develop a Fen expression
which is a continuous function of DO. However, the Japanese EFEM found that for stainless
steel there was insufficient evidence to support an expression for DO dependence for Fen. This
could be due to the lack of sufficient data available to develop such a relationship and/or
differences between the analytical methods applied to the datasets in NUREG/CR-6909 Rev 1
and the EFEM. The discrepancy between these methods highlights the inconsistency between
conclusions developed by different international groups when analyzing similar datasets. This
emphasizes the need for a single database from which the international community could work
from to reach a consensus, as original identified in KG46. Related topics under KG 20 include
oxygen effects in PWR environments and the relevance of PWR data to BWR NWC conditions,
although it is judged that these effects are unlikely to be of major significance since oxygen
appears to be beneficial for EAF in stainless steels and there seems little difference between data
in deoxygenated pure water and PWR coolant. Further efforts are now underway by EPRI to

6-5
13321451
Summary of Observations and Proposed Update to Knowledge Gaps

revise CC N-809 for FCG rates of stainless steels in PWR environments to BWR conditions,
which may shed some light on this topic. Another issue noted under KG 20 was a possible effect
of steel sulfur content, which is known to retard corrosion fatigue crack growth under some
loading conditions. Although relevant fatigue endurance data are very limited, it is judged that
sulfur is less likely to influence fatigue endurance because dissolved sulfur species may not
accumulate in short cracks, and environmental effects will be limited for longer cracks at the
higher ∆K levels in membrane-loaded fatigue specimens. Finally, it is noted that the value of the
strain amplitude threshold for EAF is not well defined because of the lengthy test times required
to obtain data at low strain amplitudes.
In the case of carbon and low alloy (ferritic) steels, NUREG/CR-6909 Rev. 1 also includes
several changes to the Fen equations and threshold values which are claimed to result from fitting
to a significantly larger database, much of which is not in the public domain. This includes
changes to the sulfur, dissolved oxygen and temperature parameters, an increase in the upper
strain rate threshold from 1 to 2.2 % s-1, and a decrease in the lower threshold from 0.001 to
0.0004 % s-1. However, the new DO dependencies do not resolve the issues raised in KG 47. An
additional issue, identified in KG 34, is the lack of data in PWR secondary environments,
although NUREG/CR-6909 Rev. 1 states that the Fen equations are intended to apply to this
environment. This is potentially an issue for steam generator shells and nozzles, for example,
where there are CUF calculations of record that must be addressed for EAF. KG 5 identifies a
deficiency in ASME Code Case N-761 which relates only to high strength steels; this is
considered a minor issue since the Fen approach is now used in preference to S-N curves specific
to water environments.
NUREG/CR-6909 Rev. 1 also develops modified Fen expressions for nickel-based alloys, and
recommends its use with the austenitic stainless steel fatigue design curve because of inadequate
data. KG 21 also identifies a need for further data to support improved relationships for Alloy
600 and its weld metals. More recent data on high chromium nickel alloys and weld metals
(Alloys 690 and 52M) [60] show similar behavior to earlier data which underpin the earlier
Japanese data on Alloy 600 and related weld metals [61]. Both sets of data suggest that the use of
stainless steel Fen relationships for nickel-based alloys is overly conservative.
Although a number of changes have been made to the Fen equations and threshold values in
NUREG/CR-6909 Rev. 1 [3] compared to the earlier report, there are still significant data gaps
which adversely impact the revised expressions, especially the definition of threshold conditions
below which environmental effects are not significant. The interactions between some
influencing parameters are not well defined. EAF data for nickel-based alloys are still limited.
Although SCC is likely to be dominant at low strain rates for Alloy 600 and related weld metals
(182, 132 and 82), this is not the case for Alloy 690 and its weld metals, although environmental
effects on fatigue appear to be significantly smaller than for austenitic stainless steel. A
comprehensive stakeholders’ database on EAF is still not available, although international
collaboration on this topic has developed significantly since the issue of the EPRI KG report in
2011.
Applicability of Air Fatigue Design Transference Factors to LWR Environments
Since the Fen assessment methodology applies the environmental factor to the fatigue design
curve, there is an implicit assumption that the design transference factors are the same in room
temperature air and high-temperature water. However, at the time the KG report was published

6-6
13321451
Summary of Observations and Proposed Update to Knowledge Gaps

in 2011, two separate studies showed that surface finish effects appeared to be lower in high-
temperature water environments than in ambient air. Those studies included one on carbon steel
and the other on austenitic stainless steel [8; 16]. Since the observations from these two studies
are inconsistent with the Fen approach of applying the Fen factor to the design curve that
incorporates the air transference factor for surface finish, KG 19 suggested that the design
margin might be reduced in LWR environments. More recent studies in both France and the UK
[15; 16; 43] have confirmed this observation for austenitic stainless steels in PWR environments,
although there is some variation between the different studies in the value of the factor by which
the air transference factor should be reduced under PWR conditions. Further studies are
underway to improve the understanding of the effects of surface finish in LWR environments,
including the INCEFA+ European collaborative project as well as work in Japan. It has also been
suggested that, since the size effect factor in NUREG/CR-6909 is attributed to an increased
probability of initiating cracks on a large surface in plant components rather than on small
fatigue specimens that such a factor should be less applicable to rough surfaces where cracks are
more likely to initiate. A recent French proposal suggests removing the size factor for rough
surfaces [44]. Other studies, however, suggest that size effects are negligible for tensile loading,
even in air [36; 180].
For austenitic stainless steels in PWR environments, there is now a significant body of evidence
that the Fen approach overestimates environmental effects when applied to the ASME Section III
fatigue design curve, mainly because surface roughness effects are lower in water compared to
air. This results in an under-prediction of fatigue life (or an over-prediction in usage factor) by a
factor of about three. Several assessment approaches to address this discrepancy have been
proposed, as discussed in Section 6.2.4, although more validation data are likely to be required.
Relevant data for other materials and for all materials under BWR conditions are very limited, so
this issue therefore warrants further study to assess its applicability for these materials and
environments.
Strain Amplitude Effects on Fen
Since the NUREG/CR-6909 approach applies a fixed value of Fen to the fatigue life calculated
using the ASME Section III design curve, it follows that the approach assumes that the
environmental effect is independent of strain amplitude, i.e., it is constant for low and high cycle
fatigue above the fatigue endurance limit and the strain threshold for EAF. KG 7 identifies that
fatigue life is made up of several different processes (crack nucleation, plus microstructurally
short, mechanically short and long crack growth stages), and the contributions of each of these
stages to overall fatigue life differ for low and high strain amplitudes. Therefore, it would be
expected that Fen should vary with strain amplitude. Recent studies [27; 54; 55; 56] appear to
support this observation and, in particular, appear to indicate a significant effect on crack
nucleation, but there are insufficient data to support the use of a strain amplitude-dependent Fen.
However, recent work in support of the development of a total life assessment methodology [57],
which is similar to the EdF Energy R5 approach [25] in treating the individual nucleation and
short/long crack growth phases separately (see Section 5.1.1), essentially resolve this concern
since the Fen methodology would only be used for a much smaller crack size (e.g., 250 µm), with
crack growth calculations covering the remainder of the fatigue life. This methodology, which is
in a preliminary phase of development, provides a potentially better means of addressing this
issue, and has the additional benefits of dealing with the through-wall strain gradients caused by
thermal transients in thick-walled components; see Section 6.2.4.

6-7
13321451
Summary of Observations and Proposed Update to Knowledge Gaps

Effects of Mean Stress on Fatigue Life and EAF


Several of the KGs relate to mean stress (or strain) effects on fatigue life. The need for additional
data to determine the extent to which both positive and negative mean stress could influence
fatigue life is identified in KG 11, as well as the requirement for further analysis in KG 44. KG
14 focuses specifically on the lack of data relating to possible mean stress effects on the fatigue
threshold, whilst KG 6 notes that mean stress could be especially important for austenitic
stainless steels due to sequential cyclic strain hardening, softening and secondary hardening.
Positive mean stress has generally been regarded as detrimental, with negative mean stress being
beneficial. However, since most fatigue tests are performed under strain control, conduct of
testing is complicated; testing under strain control with mean strain shows there is no effect of
the mean strain on fatigue life because the loading shakes down to R = -1. Testing of austenitic
stainless steels in air under strain control, in which a positive mean stress is actively maintained,
showed a significant (30 % for 50 MPa mean stress) reduction in life above 105 cycles [79],
whereas a much smaller effect of mean stress has been observed under load control. A simple
explanation for the effect is that hardening of the material due to the applied mean stress reduces
the strain amplitude for a given stress amplitude. In stress-controlled tests, although the
maximum stress is increased which would be expected to reduce fatigue life, this is offset by the
reduction in plastic strain due to hardening. In fact, some studies on stainless steel under stress
control have shown an increased fatigue life with a non-zero (tensile or compressive) mean stress
[54] which was attributed to a reduction in strain amplitude due to hardening. At higher tensile
mean stresses, a reduction of fatigue life has been observed in stress control [80], which was
attributed to an increase in strain amplitude due to ratcheting.
Only very limited data on mean stress effects in LWR environments are available, so this topic
remains a KG. However, as the largest effects of mean stress (in air) are observed at low strain
amplitude where environmental effects are expected to be lower, this may not be a major
concern. This is consistent with Japanese data on stainless steels in PWR environments which
show no effect at a strain amplitude of 0.6 % [39]. Data on 316L stainless steel in a BWR HWC
environment using hollow specimens under stress control showed more complex behavior than
in air, with a reduction in fatigue life for small tensile mean stresses, but an increase for
compressive and higher mean stress [54]. The difference between air and water data was
tentatively explained as being due to an additional positive mean stress in the hollow specimens
used for the water tests due to internal pressurization.
In summary, data on the influence of mean stress (or strain) on fatigue life are limited and are
often difficult to interpret, especially for stainless steels due to cyclic hardening and softening.
Results depend on the method of testing, i.e., strain versus stress controlled. There are very few
data in LWR environments, although testing under these conditions is planned as part of the
INCEFA+ collaborative project [12].

6.2.4 Influence of Complex and Non-Isothermal Loading Transients on Fatigue


Life in LWR Environments
As described above, the value of Fen is dependent on several factors, including strain rate and
temperature. The Fen equations in NUREG/CR-6909 [3; 8] were developed from test data
predominantly conducted at constant temperatures and strain rates (fully-reversed, triangular or
sawtooth loading). Actual component loading histories encountered during plant service are far

6-8
13321451
Summary of Observations and Proposed Update to Knowledge Gaps

more complex, with both strain rates and temperatures varying throughout each operational
transient, so that the application of a constant value of Fen is unclear and may be inappropriate
for such transients. Most operating nuclear plants also experience long periods of operation
under relatively constant loading, interrupted by relatively infrequent large transients in load and
temperature. For example, it is not uncommon for nuclear plants to operate for 18- or 24-month
cycles with only a heat-up event in the beginning and a cool-down event at the end.
Several KGs relate more specifically to the lack of understanding of the implications of such
loading conditions for EAF relative to standard cyclic membrane-loaded fatigue test specimens,
and identify the need for more data to develop improved understanding and assessment
approaches that are appropriate for application of actual components and transients. KG 18
focusses on the need for data under complex, but isothermal, loading conditions and highlights
the need for improved strain rate calculation methods for such conditions and the proportioning
of damage to different portions of the loading cycle; the need for data to understand the influence
of hold times is also identified under this KG. KG 15 states that more data with variable
temperatures and strain rates within cycles are required, including out-of-phase variations of
temperature and strain rate which are more representative of actual component thermal cycles.
KG 16 relates mainly to the requirement for data under mixed thermal/mechanical loading,
especially for thick section components for which through-wall stress gradients are present.
Several other KGs (8, 39, 40, 41, 42 and 43) relate to the requirement for the development of
improved assessment procedures to account for EAF under these more complex loading
conditions relevant to plant components; improved assessment procedures are discussed later in
Section 6.2.7.
A method for accounting for complex loading, called the modified rate approach (MRA) or the
integrated Fen approach 3, is proposed in NUREG/CR-6909, based on an earlier Japanese
proposal, whereby the load cycle is broken up into a number of strain increments and then
integrated to give the Fen for the cycle. KG 17 observed that only limited data were available to
support the use of the MRA approach. More recent data have shown that, for non-triangular,
isothermal testing, this method provides an improved predictive capability compared to use of an
average strain rate over the rising load portions of the cycle [66]. The method has also been
successfully applied to environmental fatigue calculations for plant applications [181]. However,
recent testing using multi-stage loading cycles has demonstrated that the environmental
enhancement is significantly greater when the slow portion of the strain rate is closer to the top
rather than the bottom of the waveform [66]. The MRA method does not take this behavior into
account and so under-predicts fatigue life under some loading conditions. Moreover, recent
thermo-mechanical test data indicate clearly that fatigue lives are greater for loading in which the
temperature and load are out-of-phase than when they are in-phase, which again is not accounted
for by the MRA method. The former condition is much closer to that produced by thermal
transients in LWR plants, for which MRA is likely to over-predict fatigue damage. Several new
methods for calculation of Fen under complex loading, when both strain rate and temperature
vary during the load cycle, have recently been proposed to account for the above effects and

3
The latter term should not be confused with Fen-integrated which has been used by EDF to describe the proportion of
the Fen value in the NUREG/CR-6909 equations (for stainless steel in a PWR environment) which is believed to be
already incorporated in the ASME Section III fatigue design curve; see Section 5.2.1.

6-9
13321451
Summary of Observations and Proposed Update to Knowledge Gaps

produce significant improvement in predictive capability compared with previous methodologies


[66]. These are discussed further in Section 5.3. The models have so far been applied to stainless
steels in PWR environments and still require additional validation for application to plant-
relevant transients. No relevant test data on ferritic steels or nickel-based alloys are available
against which to assess the models for these materials, nor for any materials in BWR
environments.
In order to evaluate possible effects of hold periods, testing has been carried out in ambient
temperature air, with cyclic loading interrupted by periods of holding at elevated temperatures,
either at or above normal LWR operating temperatures [70]. A beneficial effect of hold periods
was observed at low strain amplitudes (<0.4 %) and was tentatively attributed to an influence on
cyclic hardening. Only limited data are available on hold time effects in PWR environments; no
effect was observed but this may have been because the tests were conducted at a high strain
amplitude of 0.6 %. In the absence of relevant LWR data at low strain amplitudes, a tentative
conclusion from these studies is that, whilst there is a possibility of a benefit of hold times under
LWR conditions, any effect is likely only at low strain amplitudes, may be relatively small and is
likely to saturate for long hold times.

6.2.5 Prototypical and Component Testing


The database of environmental fatigue data is based on small scale round bar and tubular
specimens. The applicability of these small scale specimens to plant assessments from a
geometry, surface finish, and loading point of view is not fully understood. As discussed above,
recent testing has improved understanding by considering more plant-realistic temperature
cycling and variable strain rate tests, as well as more representative surface finishes. These have
been used to validate the latest fatigue endurance models and good agreement between
prediction and experiments have been obtained. However, there is still a lack of test data to
support the assumption that these models are applicable to a full scale component. KG 33
identified the need for more data using component-like features with plant-representative
loading, including the effects of more complex geometrical features that are present in actual
plant components, such as elbows, bends and changes in bore. Differences in constraint between
plant components and small scale tests are likely to be important, so KG 36 observed that there
was a need for more testing that addresses the influence of multiaxial loading. The proposed
EPRI prototype component testing that will begin in 2018 will help provide data to support this
need. It was also recognized in KG 35 that improved understanding of actual plant transients is
required. As there is now more than 30 years of plant monitoring experience available, there is a
wealth of plant operational transient data that could be investigated for this KG.

6.2.6 Fatigue Crack Growth


Conventional Cyclic Loading
The main focus of recent work on crack growth has been on austenitic stainless steels. KG 23
noted that ASME Section XI did not include a crack growth reference curve for wetted flaws in
stainless steel materials, but recognized that an ASME Code Case was under development for
PWR environments. Code Case N-809 has now been published to address this deficiency, but it
only addresses 316, 304, and the corresponding L grade stainless steels in PWR environments.
The equations are also stated to be applicable to equivalent cast stainless steel grades, although

6-10
13321451
Summary of Observations and Proposed Update to Knowledge Gaps

supporting data appear to be quite limited. KG 24 identified a need to develop a crack growth
reference curve for BWR NWC and HWC environments. An extension of the ASME CC N-809
work has been initiated in EPRI in 2018 to determine similar relationships for BWR
environments which should address this gap; it will be necessary to determine if available data
are sufficient to fulfill this requirement.
KG 22 identified a lack of data on stabilized grades of stainless steels. More recent data [75] has
suggested that Types 347 and 321 grades behave similarly to 304, and there also appears to be
little difference between low (L) and high carbon grades [53; 76]. The latter observation conflicts
with the Code Case N-809 expressions which give equations showing faster rates for Types 304L
and 316L compared to Types 304 and 316. The Code Case was based on an evaluation of data on
several steel heats [109] in which other factors other than carbon content might have contributed
to variability. The developed curves show an Arrhenius temperature dependence above 150°C,
but below this temperature, there is an upturn in growth rates which seems inconsistent with
more recent data.
In contrast to PWR environments, crack growth data are much more limited in BWR HWC
environments. Although the available data appear to indicate similar behavior to PWR
conditions, CC N-809 was not confirmed to be applicable in for BWR environments. Faster rates
in BWR NWC conditions indicate that an alternative approach is required for oxidizing
environments, as identified in KG24. EPRI’s new work initiated in 2018 should shed further
light on this issue.
A second issue, identified by KG 22, is the lack of threshold data in PWR environments. Code
Case N-809 uses a ∆K threshold value of 1.1 MPa√m, based on high R data used in the model
developed by Mills [109]. More recent data [108; 110] suggest threshold values between 2 and 5
MPa√m may be more appropriate. It is noted that the ASME Section XI Task Group on Flaw
Evaluation Reference Curves is currently investigating the inter-relationship between the crack
growth threshold and R-ratio, with a view to updating CC N-809, if appropriate.
KG 25 observes that crack growth rates significantly below the CC N-809 curve are observed for
stainless steels in PWR environments under some conditions, but the reasons for retarded growth
are not well understood. Substantial further work to address this issue of retardation has recently
been undertaken by several laboratories and it is now confirmed that a major factor influencing
retardation is the sulfur content of the steel, with higher sulfur contents enhancing the
retardation, and retardation occurring for sulfur contents as low as 0.006 % at low loading
frequencies. Evidence has been obtained that enhanced corrosion caused by MnS dissolution
from the steel produces a sulfur-rich environment in the crack enclave which promotes
retardation [118], although there is still uncertainty about the precise mechanism (see KG 31).
However, it has also been discovered that the sulfur may get depleted under certain
circumstances, e.g., at high water flow rates, very long cycle rise times or during hold periods, so
that it may not be possible to realize a benefit from retardation under plant operating conditions
[53]. Work on this topic is, therefore, now judged to be of lower priority than previously
envisaged.
In contrast to stainless steels, sulfur is detrimental to ferritic steels in that it enhances crack
growth in LWR environments. Enhanced crack growth rates for ferritic steels in both PWR and
BWR environments are the result of maintaining a critical high sulfur anion concentration in the
crack tip environment combined with a critical range of strain rate (see KG 30). The current

6-11
13321451
Summary of Observations and Proposed Update to Knowledge Gaps

ASME Section XI ferritic crack growth reference curves (which are based on limited data from
the 1980s) are not necessarily conservative in BWR normal water chemistry under some
conditions, especially at low loading frequencies. As identified in KG 29, ASME Code Case N-
643-2 does not apply to BWR environments, although data from Seifert [107] in BWR HWC
conditions appear to show similar behavior to PWR conditions, leading Seifert to suggest that
CC N-643-2 could be extended to cover this environment. However, the significantly higher
growth rates that can occur for oxidizing NWC conditions indicate that further work is required
to develop a crack growth reference curve for this case. EPRI’s new 2018 effort will investigate
this.
Apart from KG 28, which identifies the important need for crack growth data under complex
loading and non-isothermal conditions to be more relevant to operating plant components (see
Section 4.3.1), the remaining crack growth KGs concern nickel-based alloys. KG 26 observes
that there are significant differences between reference curves for nickel-based alloys in ASME
Section XI and curves proposed by the Japanese. It is also observed that, for Alloy 600 and its
weld metals, corrosion fatigue crack growth rates are substantially lower than for austenitic
stainless steels, although the database is significantly more limited. KG 27 identifies a lack of
relevant crack growth data for Alloy 690 and its weld metals. Only limited new work on these
topics has been identified, with the consensus that these materials experience lower corrosion
fatigue crack growth rates compared to stainless steels, and the dominance of SCC as a failure
mechanism for Alloy 600 and the associated weld metals (Alloys 182, 132 and 62) means that
corrosion fatigue remains a lower priority issue for these alloys. KG 32 queries whether
interactions between SCC and corrosion fatigue may be significant for Alloy 690, which is
normally highly resistant to SCC, unless in a heavily cold-worked condition. This issue may be
more relevant to lower chromium alloys (Alloy 600 and related weld metals) for which SCC is a
more significant issue. Some laboratory data do suggest that SCC crack growth in heats of Alloy
600 of medium SCC susceptibility can be enhanced using periodic unloading despite the fact that
the expected fatigue contribution to propagation is small. This observation raises the possibility
that enhancement of PWSCC due to dynamic straining during cyclic loading may result in
greater crack growth than predicted by a corrosion fatigue plus SCC superposition model.
Crack Growth Under Plant-Relevant Loading
As already stated, the simultaneously occurring thermal and mechanical load cycling
experienced by operating LWR plants produces complex waveform shapes, often containing
both tensile and compressive portions. These waveforms differ substantially from the simple
triangular or sine wave cyclic loading waveforms which are typically applied during laboratory
crack growth testing and are almost always applied under positive stress ratios. Many nuclear
plant components also experience long periods at relatively constant load, interspersed by
relatively infrequent transients. KG 28 therefore identified a need for additional crack growth
data to investigate the influence of different loading waveforms that better represent complex
waveform shapes, thermal transients, hold time effects and negative R loading. These studies are
complementary to the complex loading and non-isothermal fatigue life testing discussed earlier.
A number of recent studies have aimed to address the above issues, focusing predominantly on
austenitic stainless steels in PWR environments. Isothermal testing using four-part linear
waveforms has shown that the environmental contribution to fatigue crack growth is greater
when the slow rising load ramp is in the upper (higher ∆K) part of the cycle. Use of Code Case

6-12
13321451
Summary of Observations and Proposed Update to Knowledge Gaps

N-809 with an average rise time for the waveform provided an excessive over-prediction of
crack extension under these conditions because it did not take account of the variation in damage
produced in different parts of the loading cycle. A modified assessment methodology, the WKR
approach, produces a more realistic prediction under these conditions, and is discussed further in
Section 5.4.2. Testing was also performed using mechanical loading cycles more representative
of the loads produced by typical plant transients, with a similar conclusion being drawn
regarding weighting of fatigue damage. Testing has also been extended to thermo-mechanical
loading, with temperature being varied either in-phase or out-of-phase with mechanical loading.
The former condition produced similar crack growth to isothermal testing at the maximum test
temperature, whereas out-of-phase cycling produced lower crack growth, and is more relevant to
conditions produced by thermal transients. The WKR method showed a better predictive
capability than the MRA method, although some further development of the model may be
required to improve its accuracy for thermo-mechanical cycles.
Another area studied, relating to plant-relevant loading includes effects on subsequent crack
growth of overloads (where the preceding transient had a stress/strain range higher than the
following transient) and underloads (where the preceding transient had a stress/strain range lower
than the following transient). In general, overloads were found to produce some degree of
retardation of crack growth during subsequent cycling, whilst underloads at low ΔK could
produce transient high growth rates below those determined in decreasing ΔK threshold testing
conditions. Testing has also been carried out using hold times, either within individual cycles
(trapezoidal loading), or with longer hold times between series of consecutive cycles, the latter
being more relevant to plant operation. Hold time effects were complex, with retarded rates
resulting from the use of trapezoidal loading waveforms occurring in medium and high sulfur
steels similar to triangular waveforms with the same total period, but not in very low sulfur
steels. In contrast, long hold times interspersed between periods of triangular loading had a
smaller effect than a series of repeated trapezoidal cycles with the same overall hold period.
Recent testing has also studied crack growth under negative R ratio conditions, which appears to
indicate that only a portion (≤1/3) of the compressive rising load portion of the waveform
contributes to crack growth.
It is noted that all the above studies focused on austenitic stainless steels in PWR environments,
so relevance to BWR environments and to other materials such as carbon/low alloy steels and
nickel-based alloys have not been evaluated. This would require relevant experimental data for
complex transients and non-isothermal conditions similar to the recent PWR studies on stainless
steels.

6.2.7 EAF Assessment Methods


The main underlying driver for preparation of the original EPRI EAF gap analysis and research
Roadmap reports was the fact that the available assessment methods to account for
environmental effects on both fatigue endurance life and fatigue crack growth, for example those
in NUREG/CR-6909 and ASME Code Cases, can result in high fatigue usage factors and
substantially higher wetted crack growth rates for PWR and BWR components. As identified in
KGs 2 and 4, these predictions are not consistent with field experience where there have been
no reported component failures attributed to LCF, and they continue to challenge justifications
for safe long-term nuclear plant operation. There is also a potential impact on plant management
activities, such as in-service inspection intervals or component replacement policies (see

6-13
13321451
Summary of Observations and Proposed Update to Knowledge Gaps

KG 45). Consequently, a need was identified to resolve a number of uncertainties concerning the
understanding and treatment of EAF for LWR components with the aim of developing more
realistic, but adequately conservative, assessment methodologies. In support of this goal, this
report provides a review of recent developments which have led to proposals for modified EAF
assessment guidance that are less prohibitive in the demonstration of safe long-term plant
operation.
In addition to the conservatisms inherent to ASME Code requirements, designers have also often
included additional conservatisms in their analyses. One example is the way in which design
transients are defined, such as the assumption of step changes in temperature to maximize
stresses. Combination of such an assumption with a slow rise time averaged over a real or
assumed transient (which will maximize the environmental factor, Fen) adds excessive
conservatism. In addition, transients are combined in the most severe ordering possible per
ASME Code procedures, which ignores the order of actual plant history, and is especially
unrealistic when time-dependent processes due to environmental effects are present. Another
significant conservatism in the ASME Code procedure is the use of the simplified elastic-plastic
strain correction factor, Ke. This methodology was intended to simplify analyses to avoid the
need to perform more complicated and costly elastic-plastic analyses. Some researchers and
plant owners have argued that these conservatisms compensate for the maximum expected
environmental effect, which also provides a potential explanation for the apparent discrepancies
between fatigue life and plant operating experience, as identified in KG 2. When submitting
license extension applications, many plant operators have resorted to the use of more realistic
transient profiles based on improved knowledge of actual plant transients (see KG 35), and a few
instances of elastic-plastic analyses for critical component locations in order to offset EAF
effects.
The basis of most fatigue design assessments for nuclear plants, such as those in ASME Section
III and other international codes and standards, is a fatigue design curve which is derived from
tests on small-scale, membrane-loaded, polished specimens carried out in air, most commonly at
ambient temperature. Although often referred to as “fatigue initiation” tests, several stages of
cracking occur up to the defined failure criterion (often identified as 25% load drop when tests
are performed under strain control), usually identified as nucleation, microstructurally short
crack coalescence, and mechanically short crack growth (see Section 5.1.1). For typical
specimen sizes used to develop the database, the failure criterion corresponds to a final crack of
about 3 mm. A mean strain amplitude (εa) versus life (N) curve is derived by best fits of the
experimental data from a number of tests. To obtain design curves, the mean air curves in
NUREG/CR-6909 were adjusted by a factor of 12 on life and 2 on strain, whichever resulted in a
greater influence on fatigue life. The transference factor of 12 in NUREG/CR-6909 was divided
into subfactors that cover material variability and data scatter, surface finish, size and loading
history, as shown in Table 6-1. NUREG/CR-6909 Rev. 1 supported the retention of a factor of
12 on life as the adjustment factor even though further analysis suggested a factor of about 10
was justified.

6-14
13321451
Summary of Observations and Proposed Update to Knowledge Gaps

Table 6-1
Factors on Life Applied to the Mean Fatigue ε–N Air Curve to Account for the Effects of
Various Material, Loading, and Environmental Parameters [3].
5th - 95th Percentiles
Mean Life Factors 95th Percentile Life
Parameter for Lognormal
for CS, LAS, SS Factors for CS, LAS, SS
Distribution
Material variability and
2.1-2.8
data scatter
Size effect 1.0-1.4
Surface finish 1.5-3.5
Loading history 1.0-2.0
TOTAL ADJUSTMENT 3.15-27.4 5.0, 4,8, 4.9 10.2, 9.0, 9.6

It should be noted that a transference factor of 10 on life for austenitic stainless steels has
recently been proposed for incorporation into the French RCC-M Code in 2016, which is
consistent with the updated NUREG/CR-6909 analysis [3]. This relates to KG 12 which suggests
there may be conservatism in the value of factor of 12 (for all alloys) in air. The factor of 20 used
to derive the ASME Section III design curves for carbon and low alloy steels has even more
conservatism than the NUREG/CR-6909 analysis shown in the above table, so it warrants
revision. In addition, the ASME Code joint design fatigue curve for carbon and low alloy steel
could be improved by separating it into the two curves defined for these materials in
NUREG/CR-6909.
Since the NUREG/CR-6909 EAF assessment method applies the Fen factor to the design curve, it
inherently assumes that the transference factors between the mean data and the design curve
remain the same in high-temperature water compared to room temperature air. In fact, as
discussed in Sections 2.2.1 and 3.2.1, there is now significant evidence that this is not the case,
predominantly due to a smaller effect of surface finish in high-temperature water compared to
ambient air. As a consequence, a portion of the environmental factor, Fen, calculated using the
NUREG/CR-6909 equations is already accounted for in the ASME air design curve. A number
of approaches to modifying the NUREG/6909 methodology to account for this conservatism
have been proposed. A correction factor, Fen-integrated, was proposed by Métais et al. [44] to reduce
the applicable Fen value by a factor of 3 relative to that in NUREG/CR-6909. A very similar
approach, but using the term Fen-threshold, is now the subject of an ASME Code Case submission
[182], and a further modification, Fen-incorporated, is under development in the UK. A 2018 paper
reviews these approaches in detail [49]. A further proposed modification to the RCC-M Code
permits the use of a larger Fen-integrated value of 5 if the component region under investigation is
exposed to primarily hot and cold shock transients [37]. This is in recognition of the fact that the
out-of-phase cycling of stress and temperature further reduces the overall environmental effect
on fatigue, as reported by Le Duff et al. [16]. However, alternative approaches to address thermal
shock transients have also been developed and are discussed below.
These approaches, currently only validated for stainless steels in PWR environments, partially
address KG 19 which identifies that differences in surface finish effects in air and water may
make it possible to reduce the fatigue design margin in high-temperature water environments. It
should be noted that even for austenitic stainless steels in PWR environments, there is some

6-15
13321451
Summary of Observations and Proposed Update to Knowledge Gaps

variability in surface finish observations between different laboratories and further validation
data would be beneficial. As noted above, there may be additional conservatisms in some of the
other air transference factors when applied to a water environment.
In contrast to the factor on life, less attention has been given to justification of the factor of 2 on
stress/strain, which is the subject of KG 3. However, a recent statistical analysis of French high
cycle fatigue data [13] has proposed that a factor of 1.4 on stress is sufficient for austenitic
stainless steels relevant to French PWRs and this has been proposed for incorporation in the
French RCC-M Code [44]. The reduced factor arises from decreased data scatter due to
evaluation of data specific to a smaller number of steel grades (304L and 316L). Furthermore, a
later paper concludes that inclusion of Finnish/German data on stabilized 321 and 347 grades
into the analysis still supports a factor of 1.4 [40]. The effect of the change is to bring the RCC-
M fatigue design curve much closer to the 2017 ASME Section III stainless steel design curve.
Several KGs relating to assessment methodology concern the potential differences in the
observed behavior between actual plant loading conditions and cyclically-loaded isothermal
laboratory tests. KG 18 relates to the method for combining non-contiguous cycle pairs which
has not been investigated for EAF. KG 39 addresses methods for determining an appropriate
strain rate for complex loading cycles, and KG 43 concerns the assessment of EAF when both
strain rate and temperature vary during a cycle. KG 41 is related to KG 39 but is concerned
specifically with the use of the strain rate specific EAF S-N curves in ASME Code Case N-761
rather than the Fen approach in NUREG/CR-6909 and Code Case N-792-1. As discussed in
Section 5.3.2, significant new experimental data have recently been generated which provide a
basis for developing new assessment methodologies. Thus, S-N testing of stainless steels in
PWR environments has indicated that, for fatigue cycling with varying strain amplitude, fatigue
lives vary depending on the whether the slower portions of the waveform (which produce a
higher degree of enhanced fatigue damage) were closer to the top or bottom of the loading cycle.
Thus, slow ramps near the bottom of an increasing strain cycle had a smaller effect than when
they occurred near the upper portion of the loading cycle. Essentially similar effects are seen in
crack growth tests (Section 4.3.1), with higher crack growth rates when the slow rise portion of a
waveform is near the top of a cycle than vice versa.
Several new methods for calculation of Fen under complex loading have recently been proposed
to account for the above effects. These methods apply a weighting of Fen increments depending
on the position within the strain cycle, such that slow strain rates in the upper portion of the cycle
contribute more to damage than those in the lower part of the cycle. This was justified based on
observations where fatigue life was shown to have a linear dependence on plastic strain rather
than total strain amplitude. In practice, the magnitudes of plastic strains during the rising portion
of a cycle can only be obtained from load-displacement hysteresis loops which are not normally
available from elastic plant component analyses. The “strain-life weighted” (SNW) model
provides an approach which avoids this need by deriving the required weighting curve directly
from an appropriate S-N curve, and has been shown to provide a much better prediction of EAF
for complex cycling waveforms than the use of the MRA. The MRA does not allow for
weighting of fatigue damage according to the position within the strain cycle [66].

6-16
13321451
Summary of Observations and Proposed Update to Knowledge Gaps

S-N tests using thermo-mechanical loading also show a difference in fatigue life depending on
whether the thermal and mechanical cycles are in-phase or out-of-phase. This difference is not
accounted for by the MRA method which significantly under-predicts life for out-of-phase
loading, which is more relevant to thermal transients in actual plant components. The SNW
approach has also been applied to these test data and showed improved predictive capability with
a small degree of over-prediction for the out-of-phase condition.
Only very limited data using complex and non-isothermal loading cycles are available for ferritic
steel, and none for nickel-based alloys. This, together with similar stainless steel data in BWR
environments, is a significant data omission and prevents evaluation of the suitability of the new
assessment approaches to other alloys and environments.
KG 28 relates to crack growth behavior under plant-representative loading. Relevant data are
quite limited, although crack growth testing of stainless steels in PWR environments has been
performed using complex load cycles, either bilinear up and down ramps or cycles more closely
representing the stress (and hence K) variation during actual component thermal transients.
Observations are analogous to S-N test results where the observed crack growth rates for
complex cycles are greater when the longer rise time portions of the waveform are in the lower
part of the cycle. [133; 136] A modelling approach for fatigue crack growth under complex
loading, the weighted K-rate (WKR) model, has been developed based on these data [139],
which weights the fatigue crack growth through a complex loading cycle towards the upper
portion of the rising load portion of the cycle in accordance with experimental observations
(Section 5.4.2). The model was found to give good predictions of crack growth rates for a range
of complex waveforms, which were not well described by the use of Code Case N-809 with an
average rise time.
Additional thermo-mechanical crack growth tests were also carried out in which both
temperature and load were varied throughout the loading cycle, both in- and out-of-phase.
Similar to the lower observed life times from S-N testing, higher growth rates were observed for
the in-phase than for the out-of-phase TMF cycling. The WKR method was also applied to these
tests and found to give reasonable agreement; it is possible that this might be improved if a better
representation of temperature dependence was included than that given in the CC N-809
equations [67]. As for fatigue life, data for other alloys and BWR environments are needed to
ascertain the applicability of the new assessment approach for these conditions.
An additional issue for thermal transient loading applies to thick-walled components, where a
significant stress or strain gradient exists through the wall thickness. Since the ASME Section III
design approach is based on data obtained from small membrane-loaded specimens, the through-
wall gradient impact on fatigue life is not accounted for. As a result, it is not clear how 25 % load
drop in a membrane loaded laboratory specimen relates to crack initiation and growth in a thick-
walled component where the strain decreases significantly through the wall thickness. This is
further complicated by the fact that thick-walled components typically have substantial residual
stress present from welding. Only a limited number of studies relating to thermal fatigue of
thick-walled components in LWR environments have been reported. These effects are considered
to be very important as usage factors are substantially over-predicted using the current
NUREG/CR-6909 Fen methodology.

6-17
13321451
Summary of Observations and Proposed Update to Knowledge Gaps

The issue of crack growth is usually addressed as a part of in-service inspection. If engineering
cracks are detected during inspections, flaw evaluation may be performed that includes crack
growth assessment using the appropriate reference crack growth curves, such as Code Case N-
809 for austenitic stainless steels or N-643-2 for ferritic steels in PWR environments. Subsequent
augmented inspection intervals are influenced by the predicted crack growth, so future
inspections are impacted by excessive conservatisms in prediction methodologies, as recognized
in KG 45. Similarly, a postulated defect may be assumed in the case where CUF is calculated to
be greater than unity and a similar crack growth analysis performed (e.g., a flaw tolerance
assessment). The total life assessment methodology approach aims to treat the individual
nucleation and short/long crack growth phases separately so as to enable strain gradient effects
for thermal shock transients to be incorporated (see KG 16 and Sections 5.1.1 and 6.2.6). The
basis of the approach is that the Fen methodology would only be used for estimating initiation of
smaller crack sizes (e.g., 250 µm). Fracture mechanics crack growth calculations would then be
performed to cover the remainder of the component fatigue life, taking account of the calculated
through-wall strain gradient. The total life approach is similar in concept to the EdF Energy R5
approach [25], although the latter does not incorporate LWR environmental effects. To enable its
use, it would be necessary to develop crack growth relationships for short crack growth, say from
250 µm up to 3 mm. A first attempt to do this based on striation counting on tested S-N
specimens has been reported (Sections 3.2.4, 3.5.1 and 5.2.1) and an EPRI-funded experimental
study on short crack propagation is now underway.
An additional identified conservatism arises from assuming that a cumulative usage factor of
unity for thick-walled components subjected to thermal transients corresponds to an engineering
size crack on the order of 3 mm. Whereas a 3-mm crack represents 25 % load drop in small-scale
laboratory specimens, the equivalent state in actual components is not defined. An even greater
level of conservatism arises under the assumption that a calculated CUF equal to unity represents
through-wall leakage. The total life approach provides an alternative means of addressing this
issue by calculating a usage factor based on growth of mechanically short cracks.

6-18
13321451
7
RATIONALIZED AND CONSOLIDATED KNOWLEDGE
GAPS

7.1 Introduction
The aim of this final section of the report is to develop a revised and simplified list of KGs on
EAF that recognize substantial developments in data generation, knowledge, assessment
methodologies and mechanistic understanding since the original KG review in 2011/12. In the
previous section, the major developments and outstanding issues were summarized and reviewed
with reference to the previously identified KGs. In that summary, as in the original report, data
generation activities have been considered separately from assessment methods. This approach
creates closely related KGs addressing data and assessment issues. Since the purpose of the data
generation is to support assessment method development, it is now considered more appropriate
to combine these two aspects, so this approach was adopted in developing the new listing of KGs
which is presented below. The gaps are grouped under the headings A to F below. Relevant KGs
from the previous gap analysis and Roadmap reports are given in parentheses. Some of the
original KGs relate to more than one of the new KGs, so they are listed multiple times in the list
below.
A. General issues
1. Further studies on both fatigue endurance and fatigue crack growth to support improved EAF
assessment models and address discrepancies between predictions from EAF models and
plant procedures (KGs 2 & 4)
2. Universal stakeholders’ EAF database with appropriate screening of data (KG46)
B. Fatigue endurance data in air and fatigue design curves
1. Further work to address material variability effects and, where appropriate, develop material-
specific fatigue design curves (KG9)
2. Revised transference factors on life and stress for design fatigue curves (KGs 3 & 12)
3. Data, improved understanding and standardization of fatigue assessment methods for welds
(KGs 37 & 38)
C. LWR Environment effects on fatigue design
1. Data deficiencies relating to defining Fen equations, especially temperature and strain rate
thresholds (KGs 1, 5, 10, 13, 20, 21, 34 & 47)
2. Improved EAF design methods that incorporate differences between transference factors
between LWR environments and air (KG19)
3. Improved assessment methods which recognize strain amplitude effects on Fen (KG7)
4. Effects of mean stress on fatigue life and EAF (KGs 6, 11, 14 & 44)

7-1
13321451
Rationalized and Consolidated Knowledge Gaps

D. Influence of complex and non-isothermal loading transients on fatigue life in LWR


environments
1. Improved assessment methodologies under complex and non-isothermal loading more
relevant to plant (KGs 8, 18, 39, 40, 41, 42 & 43)
2. Validation data under plant-relevant loading to support new assessment methods, including
applicability to alloys other than stainless steel (KGs 15, 16 &17)
3. Improved total life methodologies to incorporate influence of strain gradients for thermal
shock transients (KG16
E. Fatigue crack growth in LWR environments
1. Improvements to reference crack growth curves for austenitic stainless steels, including
threshold and R ratio effects (including negative R), and extension to BWR (NWC & HWC)
conditions (KGs 22, 23, 24, 25 & 31)
2. Reference crack growth curves for carbon and low alloy steels in BWR environments (KGs
29 & 30)
3. Improved reference crack growth curves for nickel-based alloys (KGs 26, 27 & 32)
4. Assessment methods for crack growth under plant-relevant loading which recognize
weighting of damage through the loading cycle (KGs 28 & 45)
F. Prototypical and component testing (KGs 33, 35 & 36)

7.2 General Issues

7.2.1 Discrepancies Between Predictions from EAF Models and Plant Procedures
(Previous KGs 2 & 4)
Two KGs in the previous report relate to the overarching concern that there is inadequate
understanding of the reasons for the differences between relatively good plant experience
regarding fatigue failures and predictions from the Fen-based procedures developed from S-N test
data in NUREG/CR-6909.
It was recommended that a bottoms-up investigation into the basis of the proposed Fen factor
approach and fatigue design rules be carried out. Several conservatisms in the ASME fatigue
design procedures and transient specifications used for the initial fatigue design of many existing
PWR and BWR plants provide a partial explanation for the observed discrepancies between the
generally good fatigue performance of these reactors and predicted high usage factors for some
locations obtained using the NUREG and ASME Fen-based assessment procedures. More modern
finite element assessment methods and the use of actual transient histories have enabled life
extensions for many base-load plants to 60 years even when using the current Fen relationships
and fatigue life methodology. However, additional conservatisms in that approach need to be
addressed for longer lifetimes or load-following situations. There is still no specific definition of
what a CUF of unity in an actual component means, nor is there a satisfactory benchmark for
calculated CUF that aligns with anything meaningful in actual components.

7-2
13321451
Rationalized and Consolidated Knowledge Gaps

Although recent work reviewed in this report has confirmed that these issues are indeed
significant, a number of other significant factors have also been identified which may contribute
to the observed discrepancies between predictions and plant experience. Improved assessment
methodologies have been proposed which provide significantly better predictive capability for
austenitic stainless steels in PWR environments. Related approaches have also been developed
for crack growth and total life assessment. Further development and supporting data are likely to
be required and the possible applicability of these approaches to BWR environments and to other
alloys evaluated with appropriate laboratory data. EAF testing of actual components will provide
promising results for these issues.
NEW GAP A1: Further studies on both fatigue endurance and fatigue crack growth are
required to develop EAF life and crack growth models which take account of improved
understanding which is not reflected in the current Fen-based procedures.

7.2.2 Universal Stakeholders’ Database (Previous KG 46)


The need for a comprehensive, and regularly updated, database of EAF test data which would
enable the identification of specific data gaps and provide a sound basis for international
collaboration on EAF was identified following a peer review of the previous gap analysis report.
The revision 1 of NUREG/CR-6909 used a larger database to redefine the Fen equations and
redefined some thresholds. Despite this, much of the data underlying the analysis are not
publically-available. Although international collaboration on EAF testing and assessment
methodologies has improved significantly over the last few years, there has been only limited
progress in establishing a database which is readily accessible.
NEW GAP A2: There is a need for development of a universal stakeholders’ database
for EAF test data which would provide a consistent basis for development of improved
EAF assessment methodologies worldwide, and expert screening of all data should be
considered.

7.3 Fatigue Endurance Data in Air and Fatigue Design Curves

7.3.1 Material Variability and Material-Specific Fatigue Design Curves (Previous


KG 9)
Recent studies indicate significant heat-to-heat variability in fatigue behavior for austenitic
stainless steels. The reasons for such variability are not fully understood, although the effect is
more pronounced in the vicinity of the fatigue limit which appears to vary with material strength.
The latter effect is recognized in the CEN design standard, EN-13445. The use of
material-specific S-N curves has been proposed by some plant operators as a means of reducing
data scatter which introduces excessive conservatism.
NEW GAP B1: Further work is required to improve understanding of the reasons for
material variability effects, including temperature and strength, on fatigue life, and
especially on the fatigue limit for austenitic stainless steels, as well as for other
materials.

7-3
13321451
Rationalized and Consolidated Knowledge Gaps

7.3.2 Transference Factors for Design Fatigue Curves (Previous KGs 3 & 12)
Several different proposals concerning the values of the transference factor on life, which differ
between NUREG/CR-6909 and ASME Section III, have been made in recent years.
Nevertheless, there is no consensus on the data supporting some of the sub-factors or how they
should be combined. NUREG/CR-6909 provides evidence that a factor of 10 is appropriate for
all alloys, but recommends the retention of a factor of 12 for consistency with ASME Section III
and that the factor of 2 on stress should also be jointly considered. ASME Section III has adopted
the value of 12 for stainless steel and nickel-based alloys, but retains an overly conservative
factor of 20 for carbon and low alloy steels.
In contrast to the transference factor on life, NUREG/CR-6909 did not consider revision of the
factor of 2 on stress. A basis for this original factor does not exist in ASME Section III. Several
proposals to revise this factor have been made. A French proposal to reduce the factor on stress
for austenitic stainless steels to 1.4 based on statistical analysis of datasets on a limited number
of 304/316 (and L) grades of stainless steels appears to be supported by Finnish/German data on
stabilized 321 and 347 grades. The Japanese DFC has proposed a slightly lower factor of 1.34 for
stainless steels and a factor of 1.22 for ferritic steels.
NEW GAP B2: Further analysis is required to develop revised transference factors on
life and stress that are inter-related, along with a supporting technical basis. Generation
of additional test data may be necessary, and availability of existing data (Gap A2) is
needed to support this gap.

7.3.3 Weld Metal Data and Assessment (Previous KGs 37 & 38)
Relatively limited fatigue endurance data are available on weld metals, especially in LWR
environments, although existing data appear to be bound by base metal data, as assumed in
ASME Section III and NUREG/CR-6909. However, the EN-13445 standard provides separate
design curves for base and weld metals, although the justification underlying the curves is not
provided. A further issue for weldments is the influence of geometrical features, weld defects
and residual stress, which are dealt with by a range of different approaches including stress
concentration or fatigue strength reduction factors for different classes of welds, but it is unclear
which approaches are most appropriate.
NEW GAP B3: There is a need for further data, improved understanding and
standardization of the methods for dealing with welds in fatigue design.

7.4 LWR Environment Effects on Fatigue Design

7.4.1 Data Deficiencies Defining Fen Equations, Parameters and Thresholds


(Previous KGs 1, 5, 10, 13, 20, 21, 34 & 47)
The original KG analysis and Roadmap documents identified several Gaps relating to Fen based
methodologies, specifically those documented in NUREG/CR-6909. These Gaps mainly cover a
general lack of data for all materials that form the basis of the transformed parameters, their
thresholds and the way in which the parameters are combined to produce the environmental
factor. Furthermore, the update provided in Sections 3 to 5 indicates that the currently adopted

7-4
13321451
Rationalized and Consolidated Knowledge Gaps

method for applying Fen is overly conservative for application to plant components in the nuclear
industry. A list of some of the main Gaps associated with the Fen methodology in
NUREG/CR-6909 include:
• There is lack of clarity in the means by which the transformed parameters for strain rate,
temperature and dissolved oxygen were obtained and their thresholds defined
• The validity of using a simple multiplication of parameters to calculate Fen may be overly
conservative and “double-count” for some aspects of fatigue life
• Fatigue thresholds may be influenced by environment, and they may also vary between
material heats
• The lack of a strain amplitude dependence is not consistent with currently available data or
mechanistic understanding
• EAF test data do not cover the full necessary range of relevant parameters (e.g., temperature
and threshold value of 7% s-1 for strain rate parameter of stainless steel)
• There are only limited EAF data for some materials, especially nickel-based alloys
• The calculation of Fen values for some complex, plant-realistic loading conditions may be
unduly conservative due to the way cyclically variable parameters are handled within
NUREG/CR-6909 (see Section 7.5).
To fully address all the above gaps in EAF data would require substantial additional investment
in data generation and is likely to be a prohibitively expensive and time consuming undertaking
(as can be seen from the 2012 Roadmap report). Therefore, it is considered prudent and
pragmatic to focus future data generation towards areas of most benefit whilst retaining sight of
the total scope of the gaps in fatigue data. Noteworthy progress was recently made in generating
data to support advanced methodologies for use in combination with the NUREG/CR-6909 Fen
approach (see, for example, Sections 7.4.2 and 7.5). These data have so far only been developed
for stainless steels in PWR environments, and they are not fully validated or verified. In support
of those methodologies, there would be a significant benefit from generating sufficient data to
properly define the temperature and strain rate thresholds for stainless steels. This is due to the
methods utilizing weighted averages and applying them to out-of-phase thermo-mechanical
fatigue waveforms and plant-realistic testing being more reliant on data collected at the extreme
ranges of strain rates and temperatures. Availability of a database containing all relevant existing
data (Gap A2) would enable expert review and analysis of all that data, as well as helping to
identify relevant data gaps so as to focus future research activities.
NEW GAP C1: Data need to be generated that better define the temperature and strain
rate thresholds for stainless steel, which have a significant impact on predictions of
recently derived weighted average models for austenitic stainless steel.

7.4.2 Applicability of Air Fatigue Design Transference Factors to LWR


Environments (Previous KG 19)
Since the Fen assessment methodology applies the environmental factor to the fatigue design
curve, there is an implicit assumption that the design transference factors are the same in room
temperature air and high-temperature water. However, a number of studies have now established
that surface finish effects appear to be lower in a PWR environment than in air. As a
consequence, a portion of the environmental factor, Fen, calculated using the NUREG/CR-6909

7-5
13321451
Rationalized and Consolidated Knowledge Gaps

equations may already be accounted for in the ASME Section III air design curve and so should
not be included in the value of Fen used to modify fatigue usage factor or life. In a similar
fashion, it has also been suggested that the size effect factor in NUREG/CR-6909 may be
unnecessarily conservative for rough surfaces.
A number of approaches to modifying the NUREG/6909 methodology to account for these
observations have been proposed and involve the use of a correction factor which reduces the
value of Fen calculated using NUREG/CR-6909 by about a factor of 3. The factor is referred to as
Fen-integrated in a draft amendment to the French RCC-M Code (RPP, effectively a code case) and
as Fen-threshold in a draft ASME Section III Code Case. These Code Cases currently only relate to
stainless steels in PWR environments. Limited data suggest that similar effects of surface finish
in LWR conditions may be applicable to ferritic steels.
In contrast to the transference factor on life, no technical basis exists for the factor of 2 on stress
used by ASME Section III and NUREG/CR-6909. Although the NRC justifies lower factors on
life of 10 for all materials in NUREG/CR-6909, Rev. 1, they also recommend that the factor of 2
on stress be revisited in conjunction with the factor on life before the factor on life is further
adjusted. Several proposals to reduce the factor on stress have been made, including a French
proposal to reduce the factor on stress for austenitic stainless steels to 1.4 based on statistical
analysis of limited datasets, and a Japanese DFC proposal for factors of 1.34 for stainless steels
and 1.22 for ferritic steels.
NEW GAP C2: Revised factors on stress and life need to be developed, applicable
specifically to LWR environments, which can be used to develop improved EAF
assessment methods to be incorporated in ASME Section III.

7.4.3 Strain Amplitude Effects on Fen (Previous KG 7)


Since the NUREG/CR-6909 approach applies a fixed value of Fen to the fatigue life calculated
using the ASME Section III design curve, it follows that the Fen value is independent of strain
amplitude, i.e., it is constant for low and high cycle fatigue up to the threshold strain amplitude.
Since fatigue life is made up of several different processes, and the contributions of each of these
processes to fatigue life differ for low and high strain amplitudes, it is expected that Fen should
vary with strain amplitude. Although recent studies appear to support this observation, there are
insufficient data to support the use of a strain amplitude-dependent Fen. However, recent work in
support of the development of a total life assessment methodology, in which the individual
nucleation and short/long crack growth phases are assessed separately, overcome this concern
since the Fen methodology would only be used for much smaller crack sizes (e.g., 250 µm), with
crack growth calculations covering the remainder of the fatigue life.
NEW GAP C3: The total life approach, which provides a better means of addressing
strain amplitude effects on Fen, should be further developed to realize the additional
benefits for thermal transients which impose a strain gradient in thick-walled
components.

7-6
13321451
Rationalized and Consolidated Knowledge Gaps

7.4.4 Effects of Mean Stress on Fatigue Life and EAF (Previous KGs 6, 11, 14 &
44)
The data underlying NUREG/CR-6909 are based on fully reversed (R = -1) data which
corresponds to zero mean stress. S-N data with either positive or negative applied mean stress are
limited and, for austenitic stainless steels, the effects are difficult to predict due to sequential
cyclic strain hardening, softening and secondary hardening. Since most fatigue tests are
performed under strain control, the conduct of testing including mean stress is complicated.
Testing under strain control with mean strain shows no effect on fatigue life because the loading
shakes down to R = -1, whereas testing under strain control in which a positive mean stress is
actively maintained showed a significant reduction in life above 105 cycles. This was attributed
to hardening of the material due to the applied mean stress, thereby reducing the effective strain
amplitude for the applied stress amplitude. Since the largest effects of mean stress are observed
at low strain amplitudes where environmental effects are expected to be lower, this may not be a
major concern for EAF, although supporting data are very limited.
NEW GAP C4: Further testing for mean stress effects on stainless steels and other alloys
is required.

7.5 Influence of Complex and Non-Isothermal Loading Transients on


Fatigue Life in LWR Environments

7.5.1 EAF Endurance Data Under Complex and Non-Isothermal Loading


(Previous KGs 8, 18, 39, 40, 41, 42 & 43)
The Fen equations were developed from test data predominantly conducted at constant
temperatures and strain rates (e.g., triangular or sawtooth loading). Actual loading histories
encountered during service are far more complex, with both strain rates and temperatures varying
throughout the cycle. Therefore, use of a constant value of Fen throughout the loading cycle is
over-simplified and conservative. Most operating nuclear plant also experience long periods of
operation under relatively constant loading, interrupted by relatively infrequent large transients
in load and temperature. As a result, a need was identified for testing under conditions more
relevant to plant components in which strain rates and/or temperatures are varied throughout the
applied loading cycle.
Significant progress has been achieved over the last few years with performing tests to address
these issues. The majority of the testing to date has been performed on austenitic stainless steels
in PWR environments. Three main types of tests have been performed to date. The first two
involved isothermal tests using two-part linear strain up-ramps (and corresponding strain down-
ramps) with different rise times, and similar tests but with gradually changing strain rise times
that more closely simulate actual plant transients. In either case, the lower and higher strain rise
time portions could be positioned in different parts of the cycle. The third test type was thermo-
mechanical loading in which the thermal and mechanical components could be varied either in or
out-of-phase with each other. These tests led to two main conclusions. First, a slow strain ramp
near the top of the strain cycle produced a greater environmental effect on fatigue life than if it
was positioned close to the bottom of a strain cycle. Second, in-phase thermo-mechanical
loading produced more fatigue damage than out-of-phase loading. The use of the standard Fen
methodology, even when combined with the modified rate approach (MRA), significantly under-

7-7
13321451
Rationalized and Consolidated Knowledge Gaps

predicted fatigue life for some of the loading conditions which most closely represented common
thermal transients in plant components. New methods for calculation of Fen under complex
loading, where both strain rates and temperatures vary during the load cycle, have recently been
proposed to account for the above effects. This is because that the MRA does not account for the
fact that more environmental enhancement occurs during the higher strain portion of the cycle.
The latest methods account for this with weighting functions that scale the environmental
damage throughout the cycle. Plant representative temperature cycling tests on austenitic
stainless steel in PWR environments were used to validate the models, and good agreement was
obtained between predicted and experimental test lifetimes.
Plant transients that occur due to start-up, shut-down or power adjustments may be separated by
long periods of time where the plant is operating within its on-load fuel cycle (stable operation).
To evaluate possible effects of hold periods (as a result of stable operation) on the fatigue life of
stainless steel specimens, testing has been carried out in ambient temperature air with the cyclic
loading interrupted by periods of holding at elevated temperatures. These hold periods were
conducted at elevated temperatures relevant to or above normal LWR operating temperature.
Following the hold period, the test temperature was reduced to ambient prior to the restart of
cyclical loading. A beneficial effect of hold periods was observed for stainless steels at low strain
amplitudes (<0.4 %) and was tentatively attributed to an influence on cyclic hardening. Only
limited data are available on hold time effects in PWR environments; no effect was observed but
this may have been because the tests were performed at a high strain amplitude of 0.6 % that
induced significant cyclic hardening.
NEW GAP D1: Recently developed assessment methods for austenitic stainless steel
which weight the damage through a complex loading cycle require further supporting
data from true plant-realistic transients, including thermal shock tests, to provide the
necessary evidence that these models are fit for purpose, and to provide the mechanistic
understanding of the actual damage process through the cycle to support the weighting
used in the models.

7.5.2 Treatment of Cyclically Variable Parameters in Fen Equation (Previous KGs


15, 16 & 17)
For a cycle which has varying strain rate and temperature, the use of averaged Fen parameters is
not appropriate, as demonstrated by the results of isothermal variable strain rate tests and in- and
out-of-phase thermo-mechanical tests on austenitic stainless steel. The modified rate approach
integrates the Fen throughout the cycle to account for changes in these parameters and provides
improved predictions over the standard Fen equation. The limitation of MRA is in its lack of
weighting throughout the cycle, as it is now known that more environmental enhancement occurs
during the higher strain portion of the cycle. The latest methods account for this with weighting
functions that scale the environmental damage throughout the cycle. The approach has only so
far been validated for austenitic stainless steel in PWR environments.
NEW GAP D2: Appropriate validation data are required to determine the applicability
of improved EAF assessment methods, which enable weighting of damage according to
the position in the loading cycle, to austenitic stainless steels in BWR environments, as
well as to other alloys.

7-8
13321451
Rationalized and Consolidated Knowledge Gaps

7.5.3 Influence of Strain Gradients for Thermal Shock Transients (Previous KG


16)
An additional issue for thermal transient loading applies to thick-walled components, where a
significant stress or strain gradient exists through-wall. Since the ASME Section III design
approach is based on data obtained from small membrane-loaded specimens, the through-wall
gradient is not accounted for in fatigue design curves, nor does the laboratory specimen failure
criterion of a 3 mm deep crack (or 25 % load drop) relate to anything specific in an actual plant
component. Only a limited number of studies relating to thermal fatigue of thick-walled
components in LWR environments have been reported but, in general, cracking appears to be
substantially over-predicted using the current ASME Section III CUF and NUREG/CR-6909 Fen
methodologies.
The total life assessment methodology approach treats the individual crack initiation and
short/long crack growth phases separately so as to enable strain gradient effects to be
incorporated into fatigue life assessments. The basis of the approach is that the Fen methodology
is only applied to the initial, smaller crack size (e.g., 250 µm), followed by crack growth
calculations to cover the remainder of the fatigue life, accounting for the variation in the through-
wall strain gradient.
NEW GAP D3: The total life approach should be further developed and benchmarked
against component testing to provide a more representative assessment methodology for
fatigue life assessment, especially for thermal transients which produce through-wall
strain gradients (see Gap C3).

7.6 Fatigue Crack Growth in LWR Environments

7.6.1 Reference Crack Growth Curves for Austenitic Stainless Steels (Previous
KGs 22, 23, 24, 25 & 31)
Although ASME Section XI Code Case N-809 provides relationships for corrosion fatigue crack
growth of austenitic stainless steels in PWR environments, it is applicable to 304, 304L, 316 and
316L and corresponding cast grades only. Recent data on stabilized grades 321 and 347 appear to
show similar crack growth behavior. Some other limitations have also been identified regarding
threshold and low temperature behavior, and the relationships do not take account of retarded
growth rates which occur for higher sulfur steels at long rise times. No equivalent relationship is
available for BWR HWC or NWC conditions.
NEW GAP E1: Further work to investigate R ratio, sulfur content and ∆K threshold
effects should be pursued and curves for BWR NWC and HWC environments should be
developed, for implementation into a revision to Code Case N-809.

7.6.2 Reference Crack Growth Curves for Carbon and Low Alloy Steels (Previous
KGs 29 & 30)
ASME Code Case N-643-2 provides reference crack growth curves for ferritic steels PWR
environments which recognize the detrimental effect of steel sulfur content. This Code Case is
not applicable to BWR environments. Limited data suggest that it may be possible to
demonstrate the relevance of the PWR Code Case to BWR HWC conditions, although this may

7-9
13321451
Rationalized and Consolidated Knowledge Gaps

require additional supporting data. However, higher growth rates can occur for oxidizing NWC
conditions, so further work is required to develop a crack growth reference curve in these
environments.
NEW GAP E2: ASME Code Case N-643-2 for carbon and low alloy steels should be
expanded to cover BWR environments.

7.6.3 Reference Crack Growth Curves for Nickel-Based Alloys (Previous KGs 26,
27 and 32)
The earlier gap reports identified significant differences between reference crack growth curves
for nickel-based alloys in ASME Section XI and curves proposed by the Japanese. It is also
observed that, for Alloy 600 and its weld metals, corrosion fatigue crack growth rates in PWR
environments are substantially lower than for austenitic stainless steels, although the database is
significantly more limited. There are also only limited crack growth data for Alloy 690 and its
weld metals. Only limited new work on these topics has been identified, with the consensus
appearing to be that the lower corrosion fatigue crack growth rates compared with stainless
steels, and the dominance of SCC as a failure mechanism for Alloy 600 and the associated weld
metals (Alloys 182, 132 and 82), mean that corrosion fatigue remains a lower priority issue for
these alloys. However, some laboratory data for Alloy 600 suggest that SCC crack growth rates
in heats of Alloy 600 of medium SCC susceptibility can be enhanced using periodic unloading.
The possibility that enhancement of PWSCC due to dynamic straining during cyclic loading may
result in greater crack growth than predicted by a corrosion fatigue plus SCC superposition
model may warrant further study.
NEW GAP E3: The differences between the ASME Section XI fatigue crack
growth curves for nickel-based alloys and the Japanese curves for these
materials should be investigated.

7.6.4 Crack Growth Under Plant-Relevant Loading (Previous KGs 28 and 45)
The significant differences between operating LWRs and conventional laboratory crack growth
testing, which usually use continuous triangular (or occasionally sine wave) loading on small
compact tension specimens at positive stress ratio in a constant temperature autoclave, have
already been noted. Plant transients often involve simultaneously occurring thermal and
mechanical load cycling which produces complex waveform shapes, often containing both
tensile and compressive portions. Most nuclear plant components also experience long periods at
relatively constant load interspersed by relatively infrequent transients. Only relatively limited
data are available using more plant-relevant loading, predominantly on austenitic stainless steels
in PWR environments. Testing has included bilinear up- and down-ramp temperature, and strain
loading waveforms. It has been observed that the environmental contribution to fatigue crack
growth is greater when the slow rising strain load ramp is in the upper (higher ∆K) part of the
cycle; this is analogous to similar observations for fatigue endurance. Testing has also been
extended to thermo-mechanical loading, with temperature being varied either in-phase or out-of-
phase with mechanical loading. The former condition produced similar crack growth to
isothermal testing at the maximum test temperature; whereas, out-of-phase cycling produced
lower crack growth and is more relevant to conditions produced by thermal transients. In
response to the need for a modelling approach for fatigue crack growth under complex loading,
the weighted K-rate (WKR) model was developed based on these data [139]. The model weights

7-10
13321451
Rationalized and Consolidated Knowledge Gaps

the fatigue crack growth through a complex loading cycle towards the upper portion of the rising
load portion of the cycle in accordance with experimental observations. This method was found
to give a good prediction of crack growth rates for a range of complex waveforms, which were
not well described using Code Case N-809 with an average rise time (Section 6.2.7). Other areas
studied related to plant-relevant loading include effects of underloads (where the preceding cycle
has a stress/strain range lower than the next cycle) and overloads, the influence of hold times,
and crack growth under negative R ratio conditions.
NEW GAP E4: ASME Section XI Code Case N-809 overestimates crack growth for
complex and non-isothermal transients due to enhancement of crack growth being
weighted towards the upper (higher ∆K) portion of the transient, so further development
and validation of improved methods is needed. Further crack growth data under negative
R conditions is required since these conditions are relevant to many thermal transients.

7.7 Prototypical and Component Testing (Previous KGs 33, 35 & 36)
The discrepancies between plant experience and laboratory data is a key driver for a considerable
number of KGs and EAF research, but very few tests have been performed directly investigating
plant-relevant features and geometries. The applicability of tests on small scale, membrane-
loaded round bar or tubular specimens, which were used to develop the fatigue design curves, to
actual plant components and related assessments is not fully understood. Nevertheless, recent
laboratory specimen testing has improved understanding by considering more plant-realistic
temperature cycling and variable strain rate tests, as well as more representative surface finishes.
These test results have been used to validate the latest fatigue endurance models, and good
agreement between prediction and experiments has been obtained. However, there is still a lack
of test data to support the assumption that these models are applicable to full-scale components,
including the effects of geometrical features such as: elbows, bends and changes in bore,
differences in constraint between plant components and small scale tests, scale effects, and
multiaxial loading. It has also been recognized that improved understanding of actual plant
transients is required, perhaps supported by increased use of in-situ monitoring. EPRI’s proposed
prototype component testing would help provide data to support this requirement and would
provide a means to evaluate improved assessment methodologies under conditions more relevant
to plant components. A flexible test facility will be required in order to incorporate the effects of
different component geometries, constraint and multiaxial loading. In addition, an improved
understanding of actual plant transients is required to support the application of the new models
to plant assessments.
NEW GAP F1: Testing involving the combination of actual plant components with
typical plant thermal loading, as well as realistic component surface finishes, is required
to more fully understand the application of the fatigue endurance models to actual plant
components.

7-11
13321451
13321451
8
CONCLUSIONS

Established fatigue initiation life curves, such as those given in Section III of the ASME Boiler
and Pressure Vessel Code, provide the design basis for NPP components. These curves do not
explicitly account for the deleterious effects for exposure to high-temperature light water reactor
(LWR) water environments on component fatigue initiation life.
To address the effects of the high temperature water environments, the NRC, in Regulatory
Guide 1.207 [2], NUREG-1801, Rev. 2, and NUREG-2191, prescribed environmentally assisted
fatigue (EAF) assessment guidance for fatigue initiation life in new and operating NPPs in the
U.S. The technical bases for this guidance is documented in NUREG/CR-6909 Rev 1 [3],
developed by Argonne National Laboratory based on their review of small specimen fatigue
initiation life test data. The NRC’s guidance uses an environmental factor, Fen, by which the
fatigue design lifetime is reduced in LWR environments compared to air. ASME has also
proposed Code Cases for assessing fatigue life in LWR environments either using an approach
similar to that prescribed by the NRC, or using fatigue curves fitted to the experimental data in
water environments.
In addition to the environmental effects on fatigue life, crack growth test data also indicate that
significant environmental enhancement of fatigue crack growth can occur under LWR
environmental conditions. Although reference crack growth curves or Code Cases for defect
assessment of flaws exposed to LWR environments are available in the ASME Section XI Code
for some materials and reactor environments, not all relevant structural materials and
environments are addressed. For example, ASME Code Case N-809, which was published in
2015, only addresses austenitic stainless steels in PWR environments.
Designers and plant owners pursuing nuclear plant life extension and new build licenses use
NUREG/CR-6909 (or similar) guidance to demonstrate safe operation of plant components with
respect to the fatigue aging mechanism over their intended lifetimes. However, the guidance can
often result in very high cumulative usage factors for some components, which presents
significant challenges when validating safe, long-term operation of nuclear plant components.
The high usage factors are not consistent with operational experience where there have been no
low cycle fatigue failures identified to-date.
To resolve uncertainties in the currently available guidance, EPRI solicited expert help to
identify gaps in the industry’s understanding of EAF. The results of that effort was published in
2011 in a KG Analysis report. Subsequent review and prioritization led to a second Roadmap
report that was published in 2012. These documents highlighted the technical areas where the
uncertainties existed with respect to EAF, and identified the research and development needed to
address them. This resulted in 47 KGs which were reviewed by nuclear industry experts who
allocated priorities to the various gaps. This prioritization was used by the Roadmap report to
plan out activities to guide research towards resolving the KGs that have the best cost-to-benefit
ratio.

8-1
13321451
Conclusions

Since the KG reports were published, there have been a number of significant relevant research
activities completed worldwide. These activities were primarily aimed at improving the
understanding of the differences between predictions of environmental fatigue behavior based on
laboratory data and plant component experience. As a result of these activities, a number of
alternative methodologies for EAF life assessment have been proposed which are aimed at
refining the procedures in NUREG/CR-6909 and the related ASME Code Cases. NUREG/CR-
6909 Rev 1 has also been published to address some of the identified areas of concern in the
original model equations.
Following a detailed review of research since the original KGs were developed in 2011,
significant additional laboratory data has been developed in LWR environments, with the major
focus being on austenitic stainless steels in PWR primary coolant environments, which was the
greatest need identified by plant operators worldwide. Much of this testing was performed under
conditions intended to simulate plant operating conditions, such as complex loading transients,
thermo-mechanical loading and surface finish. However, test data on plant-representative
geometrical features and representative component sizes are still very limited. The observations
from some of these more recent, advanced experimental studies provide a partial explanation for
the observed discrepancies between the predictions from the simple Fen-based models based on
small specimen fatigue endurance tests under continuous cyclic loading and plant field
experience. Improved predictive models have been developed which provide a better description
of observed behavior in these recent laboratory tests; however, full validation for plant-
representative component geometries and loading histories is not available.
Based on a review of these data, the existing 47 KGs have been reviewed and a revised, shorter
list of outstanding issues which warrant further study has been developed. The rationale for the
new KG listing is described in Sections 6 and 7 of this report. A listing of the new KGs is
provided in Table 8-1.

8-2
13321451
Table 8-1
Listing of New Knowledge Gaps

Gap Relevant
Knowledge Gap Description Research Needs
Number Previous Gaps
General Issues
A1 Further studies on both fatigue endurance and Most of the KGs identified in this report are
fatigue crack growth are required to develop EAF life ultimately aimed at addressing
and crack growth models which take account of discrepancies between plant experience and
2&4
improved understanding which is not reflected in the EAF assessment model predictions.
current Fen-based procedures. Necessary research is therefore identified
against the remaining gaps.
A2 There is a need for development of a universal The availability of a complete, consistent
stakeholders’ database for EAF test data which and verified EAF database would be a major
would provide a consistent basis for development of advance and could enable expert review
46
improved EAF assessment methodologies and analysis of all the data, which would
worldwide, and expert screening of all data should support new method development.
be considered.
Fatigue Endurance Data in Air and Fatigue Design Curves
B1 Further work is required to improve understanding of This may require material specific testing
the reasons for material variability effects, including and assessment method development.
temperature and strength, on fatigue life, and 9
especially on the fatigue limit for austenitic stainless
steels, as well as for other materials.
B2 Further analysis is required to develop revised Generation of additional test data may be
transference factors on life and stress that are inter- necessary, and availability of existing data 3 & 12
related, along with a supporting technical basis. (Gap A2) is needed to support this gap.
B3 There is a need for further data, improved Work may include testing of weld metals and
understanding and standardization of the methods development of revised assessment 37 & 38
for dealing with welds in fatigue design. approaches for weld features.

8-3
13321451
Conclusions

Table 8-1 (continued)


Listing of New Knowledge Gaps

Gap Relevant
Knowledge Gap Description Research Needs
Number Previous Gaps
LWR Environment Effects on Fatigue Design
C1 Data supporting new assessment
Data need to be generated that better define the
approaches are so far confined to
temperature and strain rate thresholds, which have a
austenitic stainless steels in PWR 1,5 10, 13, 20,
significant impact on predictions of recently-derived
environments and applicability to BWR 21, 34 & 47
weighted average models for austenitic stainless
conditions and to other alloys needs to be
steel.
evaluated.
C2 Several outstanding issues have been
identified including variability in surface
Revised factors on stress and life need to be finish effects between different laboratories,
developed, applicable specifically to LWR differences between ASME Section III and
environments, which can be used to develop NUREG/CR-6909 life factors for ferritic 19
improved EAF assessment methods to be steels, possible differences between PWR
incorporated in ASME Section III. and BWR environments for stainless steel,
and different proposals for the factor on
stress/strain.
C3 The total life approach, which provides a better
Further development work and associated
means of addressing strain amplitude effects on Fen,
testing to support the total life approach is
should be further developed to realize the additional 7
recommended. The testing identified
benefits for thermal transients which impose a strain
against KG D3 is also relevant.
gradient in thick-walled components.
C4 Further testing for mean stress effects on stainless Data in LWR environments is very limited
6, 11, 14 & 44
steels and other alloys is required. and may be required.

8-4
13321451
Conclusions

Table 8-1 (continued)


Listing of New Knowledge Gaps

Gap Relevant
Knowledge Gap Description Research Needs
Number Previous Gaps
Influence of Complex and Non-Isothermal Transients on Fatigue Life
D1 Recently developed assessment methods for Further development and validation testing
austenitic stainless steel which weight the damage to support the use of new proposed
through a complex loading cycle require further assessment methods for austenitic
supporting data from true plant-realistic transients, stainless steels in PWR environments
including thermal shock tests, to provide the under plant loading conditions, including 8, 18, 39, 40,
necessary evidence that these models are fit for thermal shock transients. 41, 42 & 43
purpose, and to provide the mechanistic
understanding of the actual damage process through
the cycle to underpin the weighting used in the
models.
D2 Appropriate validation data are required to determine Validation testing under plant relevant
the applicability of improved EAF assessment loading conditions is required on ferritic
methods, which enable weighting of damage steels and nickel-based alloys to determine
according to the position in the loading cycle, to relevance of new assessment methods. 15, 16, 17
austenitic stainless steels in BWR environments, as Testing is also needed to confirm
well as to other alloys. applicability for stainless steels in BWR
environments.
D3 The total life approach should be further developed Testing in an LWR environment under
and benchmarked against component testing to thermal shock loading of components with
provide a more representative assessment plant-relevant wall thickness is required to 16
methodology for fatigue life assessment (see Gap enable benefit to be claimed for a through-
C3). wall stress gradient.

8-5
13321451
Conclusions

Table 8-1 (continued)


Listing of New Knowledge Gaps

Gap Relevant
Knowledge Gap Description Research Needs
Number Previous Gaps
Fatigue Crack Growth in LWR Environments
E1 Further work to investigate R ratio, sulfur content and Further work is planned to revise R ratio
∆K threshold effects should be pursued and curves and threshold effects in ASME Code Case
for BWR NWC and HWC environments should be N-809 but may require additional
developed, for implementation into a revision to Code supporting data, including data at negative
Case N-809. R. Evidence for the non-linear temperature 22, 23, 24, 25
dependence for temperatures below 150°C & 31
is also limited and further testing is needed.
Applicability to stabilized grades of
stainless steel and to BWRs should be
considered.
E2 ASME Code Case N-643-2 for carbon and low alloy Further testing to provide an adequate
steels should be expanded to cover BWR database for BWR (HWC and NWC) 29 & 30
environments. environments may be required.
E3 The differences between the ASME Section XI fatigue Further fatigue crack growth data may be
crack growth curves for nickel-based alloys and the required to resolve the differences. The
Japanese curves for these materials should be possibility that enhancement of PWSCC 26, 27 & 32
investigated. due to dynamic straining during cyclic
loading may warrant study.
E4 ASME XI Code Case N-809 overestimates crack Further development of the new
growth for complex and non-isothermal transients due assessment methods aimed at assessing
to enhancement of crack growth being weighted crack growth under complex, and
towards the upper (higher ∆K) portion of the transient, especially non-isothermal transient loading 28 & 45
so further development and validation of improved is required, together with appropriate
methods is needed. supporting data.
Applicability to other alloys and to BWR
environments also needs to be evaluated.

8-6
13321451
Conclusions

Table 8-1 (continued)


Listing of New Knowledge Gaps

Gap Relevant
Knowledge Gap Description Research Needs
Number Previous Gaps
Prototypical and Component Testing
F1 Testing involving the combination of actual plant Such testing is essential to provide
components with typical plant thermal loading, as well validation for existing or modified EAF life
as realistic component surface finishes, is required to assessment procedures and crack growth 33, 35, & 36
more fully understand the application of the fatigue curves.
endurance models to actual plant components.

8-7
13321451
13321451
9
REFERENCES

1. ASME. Boiler and pressure vessel code, Section III Div-1, Subsections NB, NC and ND.
s.l. : American Society of Mechanical Engineers.
2. R Tregoning. Guidelines for Evaluating The Effects of Light-Water Reactor Water
Environments in Fatigue Analyses of Metal Components. s.l. : U.S. Nuclear Regulatory
Commission, June, 2018. Regulatory Guide 1.207, Revision 1.
3. O K Chopra and G L Stevens. Effect of LWR Water Environments on the Fatigue Life of
Reactor Materials Final. s.l. : Office of Nuclear Regulatory Research, May, 2018.
NUREG/CR-6909, Rev 1.
4. H S Mehta. An Environmental Factor Approach to Account for Reactor Water Effects in
Light Water Reactor Pressure Vessel and Piping Fatigue Evaluations. Palo Alto, USA :
EPRI, December, 1995. TR-105759.
5. ASME. N-809 Cases of ASME B&PV Code Section XI. Code Case N-809 Reference
Fatigue Crack Growth Curves for Austenitic Stainless Steels in Pressurized Water
Reactor Environments. s.l. : ASME, 2015.
6. D R Tice, D Green and A Toft. Environmentally Assisted Fatigue Gap Analysis and
Roadmap for Future Research: Gap Analysis Report. Palo Alto, USA : EPRI, December,
2011. CA.2011. 1023012.
7. D Tice, D Green and A Toft. Environmentally Assisted Fatigue Gap Analysis and
Roadmap for Future Research: Roadmap. Palo Alto, USA : EPRI, 2012. 1026724.
8. O K Chopra and W J Shack. Effect of LWR Coolant Environments on the Fatigue Life of
Reactor Materials. Argonne, Illinois : U.S. Nuclear Regulatory Commission, February,
2007. NUREG/CR-6909 Rev. 0 and ANL-06/08.
9. W E Cooper. The Initial Scope and Intent of the Section III Fatigue Design Procedure.
Clearwater, Florida : Welding Research Council, Inc., Technical Information from
Workshop on Cyclic Life and Environmental Effects in Nuclear Applications, January,
1992.
10. G L Stevens, O K Chopra and R L Tregoning. Observations and Recommendations for
Further Research Regarding Environmentally Assisted Fatigue Evaluation Methods.
Stuttgart, Germany : 40th MPA-Seminar, Materials Testing Institute, October, 2014.
ADAMS Accession No: ML14258A039.

9-1
13321451
References

11. S Asada, T Hirano and T Sera. Study on a New Design Fatigue Evaluation Method.
Boston, Massachusetts : ASME 2015 Pressure Vessels and Piping Conference PVP2015,
July, 2015. PVP2015-45089.
12. K Mottershead, M Bruchhausen, T Métais, S Cicero, D Tice and N Platts. INCEFA-Plus
(Increasing Safety in Nuclear Power Plants by Covering Gaps in Environmental Fatigue
Assessments). Vancouver, Canada : ASME 2016 Pressure Vessels and Piping Conference
PVP2016, July, 2016. PVP2016-63149.
13. G Blatman, T Métais, J Le Roux and S Cambier. Statistical Analyses of High Cycle
Fatigue French Data for Austenitic SS. ANaheim, USA : ASME 2014 Pressure Vessels
and Piping Conference PVP2014, July, 2014. PVP2014-28409.
14. X Schuler, K Herter and J Rudolph. Derivation of Design Fatigue Curves for Austenitic
Stainless Steel Grades 1.4541 and 1.4550 Within the German Nuclear Safety Standard
KTA 3201.2. Paris, France : ASME 2013 Pressure Vessels and Piping Conference
PVP2013, July, 2013. PVP2013-97138.
15. J W Stairmand, N Platts, D R Tice, K Mottershead, W Zhang, J Meldrum and A G
McLennan. Effect of Surface Condition on the Fatigue Life of Austenitic Stainless Steels
in High Temperature Water Environments. Anaheim, USA : ASME 2015 Pressure
Vessels and Piping Conference PVP2015, July, 2015. PVP2015-45029.
16. A J Le Duff, A Le Francois, J P Vernot and D Bossu. Effect of Loading Signal Shape and
of Surface Finish on the Low Cycle Fatigue Behavior of 304L Stainless Steel in PWR
Environment. Bellevue, USA : ASME 2010 Pressure Vessels and Piping Conference
PVP2010, July, 2010. PVP2010-26027.
17. KTA. Components of the Reactor Coolant Pressure Boundary of Light Water Reactors,
Part 2: Design and Analysis. s.l. : KTA. Standard No. 3201.2, Issue 6/96.
18. T Métais, S Courtin, P Genette, L De Baglion, C Gourdin and J Le Roux. Overview of
French Proposal of Updated Austenitic SS Fatigue Curve and of Method to account for
EAF. Seville, Spain : Fatigue of Nuclear Reactor Components 2015, September, 2015.
33.
19. J Solin, S Reese, H E Karabaki and W Mayinger. Fatigue performance of stabilized
austenitic stainless steels – experimental investigations respecting operational relevant
conditions like temperature and hold time effects. Paris, France : ASME 2013 Pressure
Vessels and Piping Conference PVP2013, July, 2013. PVP2013-97502.
20. P Trampus and B Fekete. Isothermal and Thermal-mechanical Fatigue of VVER-440
Reactor Pressure Vessel Steels. s.l. : Journal of Nuclear Materials, 2015. 464,
pp 394-404.
21. Technical Committee CEN/TC 54. Unfired Pressure Vessels Part 3: Design. : British
Standards Institution, 2014. BS EN 13445-3:2014+A2:2016.

9-2
13321451
References

22. J Rudolph, K Langschwager, A Bosch, A Scholz, M Oechsner, M Vormwald and E Lang.


Fatigue Behavior of Butt Weld Seams: Experimental Investigation and Numerical
Simulation. Anaheim, USA : ASME 2014 Pressure Vessels and Piping Conference
PVP2014, July, 2014. PVP2014-28787.
23. J Rudolph, K Langschwager, A Bosch, A Scholz, M Oechsner and M Vormwald. High
Temperature Fatigue of Welded Joints: Experimental Investigation and Local Analysis of
Butt Welded Flat and Cruciform Specimens. Boston, USA : ASME 2015 Pressure
Vessels and Piping Conference PVP2015, July, 2015. PVP2015-45572.
24. A Bosch, J Rudolph and M Vormwald. Numerical Investigations of Seam Welds Under
Low Cycle Fatigue: Proposal for Lifetime Estimation and Recommendations for Design
With Commonly Used Guidelines. Boston, USA : ASME 2015 Pressure Vessels and
Piping Conference PVP2015, July, 2015. PVP2015-45576.
25. EDF Energy. Assessment Procedure for the High Temperature Response of Structures,
R5. 2014. Issue 3.
26. X Schuler, K Herter and M C Kammerer. Fatigue Behavior of Dissimilar Weld and
Cladding Material of Nuclear Components. Seville, Spain : Fourth International
Conference on Fatigue of Nuclear Reactor Components, September, 2015. 13.
27. K Herter, X Schuler, M Hoffmann, J Mahlke and P Kopp. Fatigue Behavior of Cladding
Material for Nuclear Components. Paris, France : ASME 2013 Pressure Vessels and
Piping Conference PVP2013, July, 2013. PVP2013-97399.
28. K Herter, X Schuler, M Hoffmann and P Kopp. Fatigue Behavior of Dissimilar Welds
Used for Nuclear Piping. Paris, France : ASME 2013 Pressure Vessels and Piping
Conference PVP2013, July, 2013. PVP2013-97400.
29. J Rudolph, P Wilhelm, A Roth, M Herbst, M Kammerer, X Schuler and K-H Herter.
Proposal for an improved engineering fatigue lifetime assessment concept. Vancouver,
Canada : ASME 2016 Pressure Vessels and Piping Conference PVP2016, July, 2016.
PVP2016-63576.
30. The Americal Society of Mechanical Engineers. Criteria of the ASME Boiler and
Pressure Vessel Code for Design by Analysis in Sections III and VIII, Division 2. New
York : The American Society of Mechanical Engineers, 1969.
31. T Métais, P Genette, J Le Roux, L De Baglion, S Courtin and C Gourdin. Status of the
French Methodology Proposal for Environmentally Assisted Fatigue Assessment.
Anaheim, USA : ASME 2014 Pressure Vessels and Piping Conference PVP2014, July,
2014. PVP2014-28408.

9-3
13321451
References

32. H Kanasaki, M Higuchi, S Asada, M Yasuda and T Sera. Proposal of Fatigue Life
Equations for Carbon & Low-Alloy Steels and Austenitic Stainless Steels as a Function of
Tensile Strength. Paris, France : ASME 2013 Pressure Vessels and Piping Conference
PVP2013, July, 2013. PVP2013-97770.
33. P Wilhelm, P Stienmann and J Rudolph. Discussion of Fatigue Data for Austenitic
Stainless Steels. Anahiem, USA : ASME 2014 Pressure Vessels and Piping Conference
PVP2014, July, 2014. PVP2014-28066.
34. S Asada, Y Fukuta, H Kanasaki and T Sera. Proposal of Surface Finish Factor on
Fatigue Strength in Design Fatigue Curve. Anahiem USA : ASME 2014 Pressure
Vessels and Piping Conference PVP2014, July, 2014. PVP2014-28601.
35. O K Chopra and W J Shack. Review of the Margins for ASME Code Fatigeu Design
Curve - Effects of Surface Roughness and Material Variability. Washington, USA : U.S.
Nuclear Regulatory Commission, September, 2003. NUREG/CR-6815, ANL-02/39.
36. A Hirano, M Nakane, S Asada and T Sera. Study on Consideration of Size Effects on
Design Fatigue Curve. Anaheim, USA : ASME 2014 Pressure Vessels and Piping
Conference PVP2014, July, 2014. PVP2014-28573.
37. S Courtin, T Métais, M Triay, E Meister and S Marie. Modifications of the 2016 Edition
of the RCC-M Code to Account for Environmentally Assisted Fatigue. Vancouver,
Canada : ASME 2016 Pressure Vessels and Piping Conference PVP2016, July, 2016.
PVP2016-63127.
38. The High Pressure Gas Safety Institute of Japan. Standard for Facilities of Ultra High
Pressure Gas. 2010. KHKS-0220.
39. S Asada, T Ogawa, M Higuchi, H Kanasaki and Y Takada. Study on Mean Stress Effects
for Design Fatigue Curves. Vancouver, Canada : ASME 2016 Pressure Vessels and
Piping Conference PVP2016, July, 2016. PVP2016-63796.
40. G Blatman, J Le Roux, T Métais, K Wallin, J Solin, E Karabaki and W Mayinger.
Statistical Methods and Database Splitting for HCF Data Analysis. Vancouver, Canada :
ASME 2016 Pressure Vessels and Piping Conference PVP 2016, July, 2016. PVP2016-
63141.
41. S Asada, Y Fukuta, K Tsutsumi and H Kanasaki. Applicability of Hollow Cyclindrical
Specimens to Environmental Assisted Fatigue Tests. Waikoloa, USA : ASME 2017
Pressure Vessels and Piping Conference PVP2017, July, 2017. PVP2017-65514.
42. M Twite, N Platts, A G McLennan, J Meldrum and A McMinn. Variations in Measured
Fatigue Life in LWR Coolant Environments Due To Different Small Specimen
Geometries. Vancouver, Canada : ASME 2016 Pressure Vessels and Piping Conference
PVP2016, July, 2016. PVP2016-63584.

9-4
13321451
References

43. L De Baglion, J Mendez, J Le Duff and A Lefrancois. Influence of PWR Primary Water
on LVF Behaviour of Type 304L Austenitic Stainless Steel at 300 °C - Comparison with
Results Obtained in Vacuum or in Air. Toronto, Canada : ASME 2013 Pressure Vessels
and Piping Conference PVP2012, July, 2012. PVP2012-78767.
44. T Métais, S Courtin, P Genette, L De Baglion, C Gourdin and J Le Roux. Overview of
French Proposal of Updated Austenitic SS Fatigue Curves and of a Methodology to
Account for EAF. Boston, USA : ASME 2015 Pressure Vessels and Piping Conference
PVP 2015, July, 2015. PVP2015-45158.
45. T Poulain, J Mendez, G Hénaff and L de Baglion. Analysis of the Ground Surface Finish
Effect on the LCF life of a 304L austenitic stainless steel in air and in PWR environment.
s.l. : Engineering Fracture Mechanics, May, 2017.
https://doi.org/10.1016/j.engfracmech.2017.05.043.
46. O K Chopra. Mechanisms and Estimation of Fatigue Crack Initiation in Austenitic
Stainless Steels in LWR Environments. Argonne, USA : U.S. Nuclear Regulatory
Commission, July, 2002. NUREG/CR-6787 and ANL-01/25.
47. J Le Duff, S Courtin, A Lefrancois and A Le Pecheur. Environmentally Assisted Fatigue
Assessment Considering an Alternative Method to the ASME Code Case N-792. Toronto,
Canada : ASME 2012 Pressure Vessels and Piping Conference PVP2012, July, 2012.
PVP2012-78767.
48. T Métais, C Stephan, P Genette, A Lefrancois, J P Massoud and L de Baglion. French
Methodology Proposal For Environmentally Assissted Fatigue Assessment. Paris,
France : ASME 2013 Pressure Vessels and Piping Conference PVP2013, July, 2013.
PVP2013-97203.
49. T Métais, A Morley, L de Baglion, D R Tice, G Stevens and S Cuvilliez. Explicit
Quantification of the Interaction Between the PWR Environment and Component Surface
Finish in Environmental Fatigue Evaluation Methods for Austentic Stainless Steels.
Prague, Czech Republic : ASME 2018 Pressure Vessels and Piping Conference
PVP2018, July, 2018. PVP2018-84240.
50. Japanese Nuclear Energy Safety Organization. Nuclear Power Generation Facilities:
Environmental Fatigue Evaluation Method for Nuclear Power Plants. s.l. : U.S. Nuclear
Regulatory Commission, March, 2011. Accession Number: ML113010189.
51. K Herter, X Schuler and T Weissenberg. Fatigue Behaviour of Nuclear Materials Under
Air and Environmental Conditions. Paris, France : ASME 2013 Pressure Vessels and
Piping Conference PVP2013, July, 2013. PVP2013-97394.
52. S Asada, K Ahluwalia, Y Fukuta and D Steininger. Study on Effects of Non-Isothermal
Condition and Strain Holding on Environmentally Assisted Fatigue in PWR Primary
Water Environment. Vancouver, Canada : ASME Pressure Vessels and Piping
Conference 2016, July 2016. PVP2016-63798.

9-5
13321451
References

53. D R Tice, N Platts, A Panteli, J W Stairmand and D I Swan. Influence of Steel Sulfur
Content on Corrosion Fatigue Crack Growth of Types 304 and 316 Stainless Steels in
High Temperature Water. Ottawa, Canada : 17th International Conference on
Environmental Degradation of Materials in Nuclear Power Systems – Water Reactors,
August, 2015. ENVDEG160.
54. P Spätig, H P Seifert, M Heczko and T Kruml. Mean Stress Effect on Fatigue Life and
Dislocation Microstructures of 316L Austenitic Steel at High Temperature in Air and
Water Environment. Seville, Spain : 4th International Conference on Fatigue of Nuclear
Reactor Components, September, 2015. 43.
55. N Platts, D R Tice and J Nichols. Study of Fatigue Initiation of Austenitic Stainless Steels
in a High Temperature Water Environment and in Air using Blunt Notch Compat Tension
Specimens. Boston, USA : ASME 2015 Pressure Vessels and Piping Conference
PVP2015, July, 2015. PVP2015-45844.
56. M C Kammerer, X Schuler, K-H Herter, T Weissenberg and S Weihe. Fatigue Behaviour
of Low-Alloy Ferritic Steel and Corresponding Dissimilar Metal Weld Subjected to
Oxygenated High Temperature Water. Ottawa, Canada : NACE 17th International
Conference on Environmental Degradation of Materials in Nuclear Power Systems –
Water Reactors, August, 2015. ENVDEG035.
57. J Mann, M Twite and M G Burke. Analysis of Fatigue Crack Growth in Standard
Endurance Test Specimens in Support of Total Life Approaches to Fatigue Assessment.
Vancouver, Canada : ASME 2016 Pressure Vessels and Piping Conference PVP2016,
July, 2016. PVP2016-63238.
58. M Kamaya. Environmental Effect on Fatigue Crack Initiation and Growth of Stainless
Steel for Flaw Tolerance Assessment. Vancouver, Canada : AMSE 2016 Pressure Vessels
and Piping Conference PVP2016, July, 2016. PVP2016-63434.
59. W K O'Donnell and T P O'Donnell. Proposed New Fatigue Design Curves for Carbon
and Low Alloy Steel in High Temperature Water. s.l. : Journal of Pressure Vessel
Technology, 2009. Vol. 131, Art No: 024003-1.
60. C Jang, J Hong, T S Kim and Y S Lee. Low Cycle Fatigue Behaviours of Alloy 690 and
52M in PWR Primary Water Condition. Asheville, USA : NACE 16th International
Conference on Environmental Degradation of Materials in Nuclear Power Systems -
Water Reactors, August, 2013. ED2013-3339.
61. M Higuchi, A Hirano, K Sakaguchi and Y Nomura. Revised and New Proposal of
Environmental Fatigue Life Correction Factor (Fen) for Carbon and Low-Alloy Steels
and Nickel Base Alloys in LWR Water Environments. Vancouver, Canada : 2006 ASME
Pressure Vessels and Piping Division Conference PVP2006, July, 2006. PVP2006-
ICPVT-11-93194.

9-6
13321451
References

62. Y Nomura, T Nakamura, S Asada and M Tanaka. Effects of Continuous Strain Rate
Changing on Environmental Fatigue for Stainless Steels in PWR Environment. Bellevue,
USA : ASME 2010 Pressure Vessels and Piping Conference PVP2010, July, 2010.
PVP2010-25194.
63. A Hirano and S Mizuta. Fatigue Evaluation Under Continuous Strain Rate Changing
Condition of Carbon Steel in BWR Environment. Anaheim, USA : ASME 2014 Pressure
Vessels and Piping Conference PVP2014, July, 2014. PVP2014-28576.
64. J D Hong, C Jang and T S Kim. Effects of mixed strain rates on low cycle fatigue
behaviors of austenitic stainless steels in a simulated PWR environment. s.l. :
International Journal Fatigue, June, 2015. 82, pp, 292-299.
65. T Seppanen, J Alhainen, E Arilahti and J Solin. Direct Strain-Controlled Variable Strain
Rate Low Cycle Fatigue Testing in Simulated PWR Water. Vancouver, Canada : ASME
2016 Pressure Vessels and Piping Conference PVP2016, July, 2016. PVP2016-63294.
66. C Currie, A Morley, N Platts, M Twite and K Wright. Models for Calculating the Effect
of Environment on Fatigue Life (FEN) for Complex Waveforms and / or Non-isothermal
Conditions. Waikoloa, USA : ASME 2017 Pressure Vessels and Piping Conference
PVP2017, July, 2017. PVP2017-66030.
67. N Platts, P Gill, S Cruchley, E Grieveson and M Twite. Effect of Variable Temperature
on the Fatigue Life and Crack Growth Rates of Austenitic Stainless Steels in PWR
Coolant Environments. Hawaii, USA : ASME 2017 Pressure Vessels and Piping
Conferences PVP2017, July, 2017. PVP2017-66029.
68. J Solin, S Reese, H E Karabaki and W Mayinger. Research on Hold Time Effects in
Fatigue of Stainless Steel - Simulation of Normal Operation Between Fatigue Transients.
Boston, USA : ASME 2015 Pressure Vessels and Piping Conference PVP2015, July,
2015. PVP2015-45098.
69. J Solin, S Reese and W Mayinger. Long Life Fatigue Performance of Stainless Steel
Discussion on Fatigue Design Curves for Stainless Steels. Baltimore, USA : ASME 2011
Pressure Vessels and Piping Conference PVP2011, July, 2011. PVP2011-57942.
70. E Karabaki, M Twite, J Solin, M Herbst, J Mann and G Burke. Fatigue Performance of
Stainless Steels (304L, 347) In Experiments Simulating NPP Operation Hold Times.
Vancouver, USA : ASME 2016 Pressure Vessels and Piping Conference PVP2016, July,
2016. PVP2016-63115.
71. J Solin, K Ertugrul, J Alhainen and W Mayinger. Effects of Hot Water and Holds on
Fatigue of Stainless Steel. Vancouver, Canada : ASME 2016 Pressure Vessels and Piping
Conference PVP2016, July, 2016. PVP2016-63291.

9-7
13321451
References

72. J Solin, J Alhainen, T Seppänen, E H Karabaki and W Mayinger. Experimental research


on cyclic response, hold effects and fatigue on stainless steels. Waikoloa, USA : ASME
2017 Pressure Vessels and Piping Conference PVP2017, July, 2017. PVP2017-66103.
73. S H Reese, J Seichter, D Klucke, E Karabaki and W Mayinger. Numerical Approach for
The Consideration of Hold Time Effects. Anaheim, USA : ASME 2014 Pressure Vessels
& Piping Conference PVP2014, July, 2014. PVP2014-28191.
74. S H Reese, J Seichter, D Klucke and H. E Karabaki. Numerical Evaluation of
Environmentally Assisted Fatigue (EAF) In Consideration of Recent Updates of the
Formulas and Hold Time Effects. Boston, USA : ASME 2015 Pressure Vessels and
Piping Conference PVP2015, July, 2015. PVP2015-45020.
75. H P Seifert and H J Leber. Corrosion fatigue initiation and short crack growth behaviour
of austenitic stainless steels under light water reactor conditions. s.l. : Corrosion Science,
Volume 59, Pages 20-34, February, 2012. ISSN: 0010938X.
76. H P Seifert, S Ritter and H Leber. Effect of Static Load Hold Periods on the Corrosion
Fatigue Behaviour of Austenitic Stainless Steels in Simulated BWR Environments.
Colorado Springs, USA : 15th International Conference on Environmental Degradation
of Materials in Nuclear Systems - Water Reactors, August, 2011. No: 0547.
77. P Brown, A G McLennan, T Hill, S Medway, J Stairmand, S Jaffer and M Wright.
Measurements of Fatigue Initiation of Carbon Steel in High Temperature Water Using
Blunt Notch Compact Tension Specimens. Ottawa, Canada : 17th International
Conference on Environmental Degradation of Materials in Nuclear Power Systems –
Water Reactors, August, 2015. ENVDEG145.
78. L Vincent, J Le Roux and S Taheri. On the High Cycle Fatigue Behaviour of a Type
304L Stainless Steel at Room Temperature. s.l. : International Journal of Fatigue, May,
2012. ISSN 0142-1123.
79. S Taheri, L Vincent and J C Le Roux. A conservative Damage Accumulation Method for
the Prediction of Crack Nucleation Under Variable Amplitude Loading For Austenitic
Stainless Steels. Paris, France : ASME 2013 Pressure Vessels and Piping Conference
PVP2013, July, 2013. PVP2013-97284.
80. M Kamaya. Influence of Mean Strain on Fatigue Life of Stainless Steel (Effect of
Constant and Ratcheting Mean Strain). Anaheim, USA : ASME 2014 Pressure Vessels
and Piping Conference PVP2014, July, 2014. PVP2014-28279.
81. P Spätig and H P Seifert. Mean Stress Effect on Fatigue Life of 316L Austenitic Steel in
Air and Simulated Boiling Water Reactor Hydrogen Water Chemistry Environment.
Ottawa, Canada : NACE 17th International Conference on Environmental Degradation of
Materials in Nuclear Power Systems – Water Reactors, August, 2015. ENVDEG083.

9-8
13321451
References

82. S Lee. The Effect of Specimen Geometry on the Low Cycle Fatigue Life of Metallic
Materials. s.l. : Materials at high temperatures, 2011. pp. 33-39.
83. P Gill, P James, C Madew, A Morley and C Currie. An Investigation into the Lifetimes of
Solid and Hollow Fatigue Endurance Specimens Using Cyclic Hardening Material
Models in Finite Element Analysis. Waikoloa, USA : ASME 2017 Pressure Vessels and
Piping Conference PVP2017, July, 2017. PVP2017-65975.
84. S Utz, E Soppa, C Kohler, X Schuler and H Silcher. Thermal and Mechanical Fatigue
Loading - Mechanisms of Crack Initiation and Crack Growth. Anaheim, USA : ASME
2014 Pressure Vessels & Piping Conference PVP2014, July, 2014. PVP2014-28411.
85. GOM. ARAMIS, Optical 3D Deformation Analysis. Braunschweig, Germany : GOM-
Gesellschaft für Optische Messtechnik mbH (Optical Measureing Techniques).
86. J L Chaboche. Constitutive Equations for Cyclic Plasticity and Cyclic Viscoplasticity.
s.l. : Int J Plast, 1989. Vol. 5, pp. 247‐302.
87. H J Leber, S Ritter and H P Seifert. Thermo-mechanical and Isothermal Low-cycle
Fatigue Behavior of Type 316L Stainless Steel in High-temperature Water and Air. s.l. :
Corrosion, Volume 69, Issue 10, Pages 1012-1023, October, 2013. ISSN: 00109312.
88. N Platts, P Brown, P J Gill, R D Smith and J W Stairmand. Development of a New
Thermo-mechanical Environmental Fatigue Testing Facility to Investigate the Impact of
Strain Gradients on Fatigue Initiation. Vancouver, Canada : ASME Pressure Vessels and
Piping Conference 2016, July 2016. PVP2016-63161.
89. C Gourdin, F Rossillon, P Le Delliou, G Perez and A Fissolo. Investigations on Crack
Growth Propagation Under Cyclical Isothermal and Thermo-mechanical Loadings for a
Type 304-L Stainless Steel Used for Pressurized Water Reactor. Paris, France : ASME
2013 Pressure Vessels and Piping Conference PVP2013, July, 2013. PVP2013-97510.
90. M Niffenegger, B Niceno, M Sharabi, K G Janssens and K Reichlin. Crack Initiation in
Austenitic Stainless Steel Owing to Cyclic Thermal Shocks and Biaxial Preload. San
Francisco, USA : SMiRT22: 22nd Conference on Structural Mechanics in Reactor
Technology, August, 2013. 731.
91. D P Jones, J E Holliday, T R Leax and J L Gordon. Analysis of a Thermal Fatigue Tests
of a Stepped Pipe. San Diego, USA : ASME 2004 Pressure Vessels and Piping
Conference PVP2004, July, 2004. PVP2004-2748.
92. J Hickling, R Kilian, L Spain and J Carey. Environmental Fatigue Testing of Type 304L
Stainless Steel U-Bends in Simulated PWR Primary Water. Vancouver, Canada : ASME
2006 Pressure Vessels and Piping Conference PVP2006, July, 2006. PVP2006-93318.

9-9
13321451
References

93. P Arora, P K Singh, V Bhasin, K K Vaze, D M Pukazhendhi, P Gandhi and G Raghava.


Fatigue Crack Growth Investigations on SS Pipes Subjected to Block Loading Overloads
and Underloads. San Francisco, USA : IASMiRT: 22nd Conference on Structural
Mechanics in Reactor Technology, August, 2013. 595.
94. K Selvam. Experimental and Numerical Analysis of Flow in a Mixing Tee. Seville,
Spain : Fourth International Conference on Fatigue of Nuclear Reactor Components,
October, 2015. 40.
95. X Schuler, M C Kammerer, S Weihe, M Seidenfuß, M Zhou, E Laurien and R Kulenovic.
Thermo-mechanical Loading of Full-Scale Welded Piping Components in Hight
Temperature Water Environment. Waikoloa, USA : ASME 2017 Pressure Vessels and
Piping Conference PVP2017, July, 2017. PVP2017-65606.
96. R Gurdal and S X Xu. Comparison of Strain Range Measures and Environmental
Fatigue Calculation Methodologies for the Stepped Pipe Tests. Chicago, USA : ASME
2008 Pressure Vessels and Piping Conference PVP2008, July, 2008. PVP2008-61915.
97. D A Steininger, K Wright, M Twite, A Morley, T Métais, G Léopold and J C Le Roux.
Component Testing Proposal to Quantify Margins in Existing Environmentally Assisted
Fatigue (EAF) Requirements. Waikoloa, USA : ASME 2017 Pressure Vessels and Piping
Conference PVP2017, July, 2017. PVP2017-65995.
98. S Bradai and J C Le Roux. Equi-Biaxial Loading Effect on Austenitic Stainless Steel
Fatigue Life. Anaheim, USA : ASME 2014 Pressure Vessels and Piping Conference
PVP2014, July, 2014. PCP2014-28421.
99. G Perez, S Gourdin, S Courtin and J C Le Roux. PWR Effect on Crack Initiation under
Equi-biaxial Loading Development of the Experiment. Vancouver, Canada : ASME 2016
Pressure Vessels and Piping Conference PVP2016, July, 2016. PVP2016-63561.
100. S Lida and M Kamaya. Fatigue Crack Initiation and Growth Observation for 316
Stainless Steel Subjected to Equi-Biaxial Cyclic Loading. Boston, USA : ASME 2015
Pressure Vessels and Piping Conference PVP2015, July, 2015. PVP2015-45814.
101. P Arora and S Tarafdar. Fatigue Studies on Carbon Steel Tubes Under Multiaxial
Loading. Manchester, UK : SMiRT 23, August, 2015. 334.
102. S K Gupta, T M Fesich, X Schuler, V Bhasin, K K Vaze and E Roos. A Critical Plane
Based Model for Fatigue Assessment Under Fixed and Rotating Principle Direction
Loading. New Delhi, India : 21st International Conference on Structural Mechanics in
Reactor Technology (SMiRT-21), October, 2011. 624.
103. X Liu. Fatigue Behavior and Dislocation Substructures for 6063 Aluminum Alloy Under
Nonproportional Loadings. s.l. : International Journal of Fatigue, 2009. 31, 1190-1195.

9-10
13321451
References

104. R C Cipolla and W H Bamford. Technical basis for code case N-809 on reference fatigue
crack growth curves for austenitic stainless steels in pressurized water reactor
environments. Boston, USA : 2015 ASME Pressure Vessels and Piping Conference, July,
2015. PVP2015-45884.
105. D R Tice, N Platts and A Panteli. Mechanistic studies on Environmentally Assisted
Fatigue Crack Growth in Light Water Reactor Environments. Seville, Spain : Fourth
International Conference on Fatigue of Nuclear Reactor Components, September, 2015.
106. H P Seifert, S Ritter and H J Leber. Corrosion Fatigue Crack Growth Behaviour of
Austenitic Stainless Steels Under Light Water Reactor Conditions. s.l. : Corrosion
Science, Vol: 55, 61-75, February, 2012. ISSN: 0010938X.
107. H P Seifert and S Ritter. The influence of ppb levels of chloride impurities on the strain-
induced corrosion cracking and corrosion fatigue crack growth behavior of low-alloy
steels under simulated boiling water reactor conditions. s.l. : Corrosion Science, Volume
108, July 2016, Pages 148-159, 2016. ISSN 0010-938X.
108. E West, H Mohr and E Lord. Fatigue Threshold Behaviour of Stainless Steel in High
Temperature Air and Water. Vancouver, Canada : ASME 2016 Pressure Vessels and
Piping Conference PVP2016, July, 2016. PVP2016-63051.
109. W.J. Mills. Critical Review of Fatigue Crack Growth Rates for Stainless Steel in
Dearated Water: Parts 1 and 2. Colorado Springs : EPRI PWR Materials Reliability
Program Conference, 2010.
110. W J Mills. Accelerated and Retarded Corrosion Fatigue Crack Growth Rates for 304
Stainless Steel in an Elevated Temperature Aqueous Environment. Ashville, USA : 16th
International Conference on Environmental Degradation of Materials in Nuclear Power
Systems - Water Reactors, August, 2013. ED2013-3343.
111. D. R. Tice, N Platts, V Allen, S Medway, A Griffiths and G Ilevbare. Influence of Hold
Times on Fatigue Crack Growth of Austenitic Stainless Steel in PWR Environments and
Implications for Mechanistic Understanding. Ottawa, Canada : 17th International
Conference on Environmental Degradation of Materials in Nuclear Power Systems –
Water Reactors, August, 2015. ENVDEG159.
112. L B O'Brien, R G Ballinger, D J Paraventi, L Yo, Y Maruno and P W Stahle. The Effect
of Environment, Chemistry and Microstructure on the Corrosion Fatigue Behavior of
Austenitic Stainless Steels in High Temperature Water. Ottawa, Canada : 17th
International Conference on Environmental Degradation of Materials in Nuclear Power
Systems – Water Reactors, August, 2015. ENVDEG011.

9-11
13321451
References

113. D B Miller and J D Paraventi. Evaluation of Oxide/Metal Interfaces Formed in Type


304/304L Stainless Steel During Environmentally Enhanced and Retarded Fatigue Crack
Growth in Deaerated Pressurized Water Using High Resolution Analytical Electron
Microscopy. Ottawa, Canada : 17th International Conference on Environmental
Degradation of Materials in Nuclear Power Systems – Water Reactors, August, 2015.
ENVDEG010.
114. E A West, C Tackes, G Newsome and N Lewis. Influence of Sulfur and Ferrite on SCC
and Corrosion Fatigue Behavior of Model Heats of Stainless Steel. Ottawa, Canada :
17th International Conference on Environmental Degradation of Materials in Nuclear
Power Systems – Water Reactors, August, 2015. ENVDEG039.
115. N Platts, D R Tice and J W Stairmand. Effect of Flowrate on the Environmentally
Enhanced Fatigue Crack Propagation of Austenitic Stainless Steels in a Simulated PWR
Primary Coolant Environment. Virginia Beach, USA : 14th Int. Conf. on Environmental
Degradation of Materials in Nuclear Power Systems, August, 2009. ED203300.
116. S P Hannula, H Hanninen and S Tahtinen. Influence of nitrogen alloying on hydrogen
embrittlement in AISI 304-type stainless steels. s.l. : Metallurgical Transactions A, Vol
15, Issue 12, pp 2205-211, December, 1984. DOI: 10.1007/BF02647103.
117. K Mukahiwa, F Scenini, M G Burke, N Platts, D R Tice and J W Stairmand. Corrosion
fatigue and microstructural characterisation of Type 316 austenitic stainless steels tested
in PWR primary water. s.l. : Corrosion Science, October, 2017. DOI:
10.1016/j.corsci.2017.10.022.
118. A Panteli, N Platts and D R Tice. Mechanistic Studies on Envrionmentally-Assisted
Fatigue Crack Growth in Light Water Reactor Environments. Vancouver, Canada :
ASME 2016 Pressure Vessels and Piping Conference PVP2016, July, 2016. PVP2016-
63134.
119. N Platts, D R Tice, K Mottershead, L Mcintyre and F Scenini. Effects of Material
Composition on Corrosion Fatigue Crack Growth of Austenitic Stainless Steels in High
Temperature Water. Colorado Springs, USA : 15th International Conference on
Environmental Degradation of Materials in Nuclear Power Systems-Water Reactors,
June, 2012. DOI: 10.1002/9781118456835.ch58.
120. T Ogawa, M Itatani, H Nagase, S Aoike and H Yoneda. Proposal of fatigue crack growth
rate curve in air for nickel-base alloys used in BWR. s.l. : Nihon Kikai Gakkai
Ronbunshu, A Hen/Transactions of the Japan Society of Mechanical Engineers, Part A.
Vol 79, Issue 806, Pages 1550-1554, January, 2013. ISSN: 03875008.
121. T Ogawa, M Itatani, T Saito, H Nagase, S Aoike and H Yondeda. Technical Basis of
Fatigue Crack Growth Rate Curve for Ni-Base Alloy in BWR Water Environment.
Prague, Czech Republic : 22nd International Conference on Nuclear Engineering, July,
2014. doi:10.1115/ICONE22-31143.

9-12
13321451
References

122. M Itatani and Takuya Ogawa. Evaluation of Crack Growth of Ni-Base Alloys Under Long
Term Cyclic Loading in BWR Environment. Boston, USA : ASME 2015 Pressure Vessels
and Piping Conference PVP2015, July, 2015. PVP2015-45458.
123. J Xiao, S Y Qiu, Y Chen, Z H Fu, Z X Lin and Q Xu. Effects of Crack Tip Plastic Zone
on Corrosion Fatigue Cracking of Alloy 690(TT) in Pressurized Water Reactor
Environments. s.l. : Journal of Nuclear Materials, 2015.
http://dx.doi.org/10.1016/j.ifatigue.2014.07.056.
124. J Xiao, S Y Qiu, Y Chen, Z X Lin, Q Xu and H Y Xie. Effects of Dissolved Oxygen on
Corrosion Fatigue Cracking of Alloy 690(TT) in Pressurized Water Reactor
Environments. s.l. : International Journal of Fatigue, 2015.
http://dx.doi.org/10.1016/j.ifatigue.2014.12.007.
125. EPRI. Environmental Fatigue Testing of Type 304L Stainless Steel U-Bends in Simulated
PWR Primary Water (MRP-188). Palo Alto : EPRI, 2006. 1013028.
126. F J Perosanz, J Laneña and M S García-Redondo. Corrosion-Fatigue Behaviour of A690
in Simulated PWR Primary Water. Seville, Spain : Fourth International Conference on
Fatigue of Nuclear Reactor Components, September, 2015. 66.
127. O K Chopra, W K Soppet and W J Shack. Effects of Alloy Chemistry, Cold Work, and
Water Chemistry on Corrosion Fatige and Stress Corrosion Cracking of Nickel Alloys
and Welds. s.l. : U.S. Nuclear Regulatory Commission, March, 2001. NUREG/CR-6721,
ANL-01/07.
128. B Alexandreanu, Y Chen, K Natesan and B Shack. Cyclic and SCC Behavior of Alloy
152 Weld in A PWR Environment. Baltimore, USA : ASME 2011 Pressure Vessels and
Piping Conference PVP2011, July, 2011. PVP2011-57463.
129. B Alexandreanu, Y Chen, B Shack and K Natesan. SCC Behavior of Alloy 690 HAZ in a
PWR Environment. Baltimore, USA : ASME 2011 Pressure Vessels and Piping
Conference PVP2011, July, 2011. PVP2011-57649.
130. B Alexandreanu, Y Chen, K Natesan and B Shack. Stress Corrosion Cracking of Alloy
152 Weld Butter Near the Low Alloy Steel Interface. Ottawa, Canada : 17th International
Conference on Environmental Degradation of Materials in Nuclear Power Systems -
Water Reactors, June, 2015. ENVDEG017.
131. B Alexandreanu, B Shack, Y Chen and K Natesan. SCC Behavior of Alloy 52M/182 Weld
Overlay in a PWR Environment. Baltimore, USA : ASME 2011 Pressure Vessels and
Piping Conference PVP2011, July, 2011. PVP2011-57465.
132. B Alexandreanu, K Natesan, Y Chen and B Shack. Stress Corrosion Cracking of a 52M
Weld Overlay in a PWR Environment. Ottawa, Canada : 17th International Conference on
Environmental Degradation of Materials in Nuclear Power Systems - Water Reactors,
June, 2015. ENVDEG016.

9-13
13321451
References

133. N Platts, D R Tice, K Rigby and D I Swan. Effect of Loading Waveform and Spectrum
Loading on the Fatigue Crack Growth Rate in Simulated Light Water Reactor
Environments. Vancouver, Canada : ASME 2016 Pressure Vessels and Piping
Conference PVP2016, July, 2016. PVP2016-63148.
134. N Platts, D R Tice, A Panteli and S Cruchley. Effect of Hold Periods on the Corrosion
Fatigue Crack Growth Rates of Austenitic Stainless Steels in LWR Coolant
Environments. Waikoloa Village, Hawaii : ASME 2017 Pressure Vessels and Piping
Conference PVP2017, July, 2017. PVP2017-65787.
135. A Roth and B Devrient. Environmental Effects on Fatigue - Possible Reasons for the
Apparent Mismatch Between Laboratory Test Results and Operational Experience.
Avignon, France : Fontevraud 7, September, 2010. A031-T05.
136. N Platts, D R Tice, J W Stairmand and D I Swan. The Effect of Complex Loading
Waveforms and Non-isothermal Conditions on Environmentally Enhanced Fatigue Crack
Growth of Austenitic Stainless Steel. Asheville, USA : 16th International Conference on
Environmental Degradation of Materials in Nuclear Power Systems - Water Reactors,
August, 2013. ED2013-3268.
137. C Watson, C Currie and J Emslie. Accounting for Negative R-Ratio (Crack Closure) in
Fatigue Crack Growth Calculations on Stainless Steel Components. Boston, USA :
ASME 2015 Pressure Vessels and Piping Conference PVP2015, July, 2015. PVP2015-
45622.
138. N Platts, D R Tice and W Zhang. Negative Load Ratio Fatigue Crack Growth Rate
Testing On Austenitic Stainless Steel in a Simulated Primary Water Environment. Boston,
USA : ASME 2015 Pressure Vessels and Piping Conference PVP2015, July, 2015.
PVP2015-45843.
139. J Emslie, C Gill and K Wright. Assessment Method to Account for the Rise Time of
Complex Waveforms in Stainless Steel Environmental Fatigue Crack Growth
Calculations. Vancouver, Canada : ASME 2016 Pressure Vessels and Piping Conference
PVP2016, July, 2016. PVP2016-63497.
140. S Gosselin and N Palm. Fatigue Usage Life and Gradient Factors for ASME Class 1
Piping Fatigue Analysis. Vancouver, Canada : ASME 2016 Pressure Vessels and Piping
Conference PVP2016, July, 2016. PVP2016-63027.
141. M A Miner. Cumulative Damage in Fatigue. s.l. : Jounal of Applied Mechanics, 1945.
Vol. 12, pp A159-64.
142. P C Paris, M P Gomez and W E Anderson. A Rational Analytic Theory of Fatigue. s.l. :
The Trend in Engineering, 1961. Vol. 13, p. 9-14.
143. AFCEN. RCC-M: Design and Construction rules for PWR Nuclear Island. Paris : RCC-
M, 2012.

9-14
13321451
References

144. British Standards Institution. PD5500: Specification for Unfired Fusion Welded Pressure
Vessels. s.l. : BSI, 2015.
145. JSME. S NC-1-2001, Codes for Nuclear Power Generation Facilities - Rules on Design
and Construction for Nuclear Power Plants. 2001.
146. J Emslie, C Watson and K Wright. ASME III Fatigue Assessment Plasticity Correction
Factors for Austenitic Stainless Steels. Anaheim : Proceedings of the ASME 2014
Pressure Vessels and Piping Conference, 2014. PVP20014-28633.
147. G Stevens. Materials Reliability Program: Evaluation of Controlling Transient Ramp
Times Using Piping Methodologies When Considering Environmental Fatigue (Fen)
Effects (MRP-218). Palo Alto, USA : EPRI, September, 2007. 1015014.
148. K Wright, C Watson, J Emslie and M Twite. Total Life Approach for Fatigue Design to a
Target Reliability. Seville, Spain : 4th International Conference on Fatigue of Nuclear
Reactor Components, September, 2015. 17.
149. K Wright. An Overview of the ASME Working Group (WG EFEM) Fatigue Action Plan.
Seville, Spain : 4th International Conference on Fatigue of Nuclear Reactor Components,
September, 2015. 16.
150. H T Harrison and R Gurdal. Comparison between ASME Code-Case N-761, NUREG/CR-
6909 and Stainless Steel Component Fatigue Test Results. Anaheim, California : ASME
2016 Pressure Vessels and Piping Conference PVP-2016, July, 2014. PVP2014-28883.
151. Y J Kim. Environmental Fatigue Evaluation Considering Transient Stress History for
Piping. Vancouver, Canada : ASME Pressure Vessels and Piping Conference PVP2016,
July, 2016. PVP2016-63478.
152. T Gilman, A Allenshwaram and T Satyan. Industry's First NRC Approved Appendix L
Flaw Tolerance Evaluation to Manage Fatigue in a Surge Line. Boston, USA : AMSE
2015 Pressure Vessels and Piping Conference PVP2015, July, 2015. PVP2015-45194.
153. ASME. Fatigue Design Curves for Light Water Reactor (LWR) Environments. s.l. :
American Society of Mechanical Engineers, September, 2010. Code Case N-761.
154. Cases of the ASME Boiler and Pressure Vessel Code Section III, Division 1. N-792
Fatigue Evaluations Including Environmental Effects. 2012.
155. G Facco. Regulatory Guide 1.193: ASME Code Cases Not Approved For Use. s.l. : U.S.
Nuclear Regulatory Commission, August, 2017. Accession Number: ML16321A338.
156. D Green. Fatigue Crack Initiation and Growth in Stainless Steel Pipe Welds. Prague,
Czech Republic : ASME Pressure Vessels and Piping Conference 2009, July 2009.
PVP2009 – 77986.

9-15
13321451
References

157. Gray M A and Verlinich MM. Guidelines for Assessing Environmental Effects in Fatigue
Usage Calculations. s.l. : EPRI Report, December 2012. 1025823.
158. K Tsutsumi, H Dodo, S Kanasaki, Y Nomoto and T Nakamura. Fatigue Behaviour of
Stainless Steels under Conditions of Changing Strain Rate in PWR Primary Water. New
York : Pressure Vessels and Piping Codes and Standards, 2001. PVP Vol 419.
159. W H Bamford, R Cipolla, A Udyawar and N L Glunt. Example Analysis for
Environmental Fatigue Crack Growth in Austenitic Stainless Steel Piping Using Code
Case N-809. Boston, USA : ASME 2015 Pressure Vessels and Piping Conference
PVP2015, July, 2015. PVP2015-45967.
160. A Roth, M Herbst, J Rudolph, P Wilhelm, X Schuler, K-H, Kammerer, M Herter and T
Weissenberg. Environmental Influences on the Fatigue Assessment of Austentic and
Ferritic Steel Components Including Welds. Anaheim, USA : ASME 2014 Pressure
Vessels and Piping Conference PVP2014, July, 2014. PVP2014-28728.
161. K-H Herter, M Schuler, M Hoffman and P. Kopp. Fatigue behaviour of dissimilar metal
welds used for nuclear piping. Paris, France : ASME 2013 Pressure Vessels and Piping
Conference PVP2013, July, 2013. PVP2013-9740.
162. EDF Energy. Assessment Procedure for the High Temperature Response of Structures,
R5. 2012. Issue 3.
163. D W Dean, L C Allport and M J Chevalier. The R5 procedures for assessing the high
temperature response of structures: current status and recent developments. Manchester :
SMiRT 23, 2015. Paper ID 458.
164. S D Sajiah, Chandra, S Bhuwan, S Jalaldeen, P Selvaraj and P Chellapandi. Stress
Indices for Non-Radial Branch Connections for Piping. s.l. : 6th International Conference
on Creep, Fatigue and Creep-Fatigue Interaction, 2013.
165. C Faidy. Uncertainties and Margins in Environmental Fatigue Analysis Rules. Anaheim :
Proceedings of the ASME Pressure Vessels and Piping Conference 2015, 2014.
PVP2014-28551.
166. T Wermelinger, F Bruckmuller and B Heinz. Fatigue Monitoring In the Context of Long-
Term Operation of the Goesgen Nuclear Plower Plant Using Areva's Famosi. Boston,
USA : ASME 2015 Pressure Vessels and Piping Conference PVP2015, July, 2015.
PVP2015-45272.
167. S H Reese, J Seichter, D Klucke and H E Karabaki. Operational Lessons Learned Based
on Comprehensive Temperature Measurements at Primary Circuit Components. Boston,
USA : ASME 2015 Pressure Vessels and Piping Conference PVP2015, July, 2015.
PVP2015-45021.

9-16
13321451
References

168. T L Meikle, W R Wetmore and M A Gray. Transient and Fatigue Monitoring


Operational Feedback Application and Evaluation. Anaheim, USA : ASME 2014
Pressure Vessels and Piping Conference PVP2014, July, 2014. PVP2014-29102.
169. Structural Integrity Associates. FATIGUEPRO(TM) On-line Fatigue Monitoring System:
Demonstration at the Quad Cities BWR. s.l. : EPRI, Feb, 1989. NP-6170-M.
170. Structural Integrity Associates. FatiguePro, Version 2: Fatigue Monitoring Software.
s.l. : EPRI, December, 1997. TR-107448.
171. Structural Integrity Associates. FatiguePro Code, Version 3.0 User's Manual Only. s.l. :
EPRI, December, 2001. 1002861.
172. Structural Integrity Associates. Use of FatiguePro Fatigue Monitoring System for
Evaluation of Local Plant Instrument Data. s.l. : EPRI, October, 1997. TR-107534.
173. A G Ware, D K Morton and M E Nitzel. Application of NUREG/CR-5999 Interim
Fatigue Curves to Selected Nuclear Power Plant Components. s.l. : U.S. Nuclear
Regulatory Commission, February, 1995. NUREG/CR-6260.
174. ASME. ASME Boiler and Pressure Vessel Code, Section XI - Rules for In-service
Inspection of Nuclear Power Plant Components, Non-mandatory Appendix L Operating
Plant Fatigue Assessment. s.l. : ASME International (BPVC), 2017. ASME BPVC.XI
2017.
175. FPL. Submittal of Pressurizer Surge Line Welds Inspeciton program. Washington, USA :
U.S. Nuclear Regulatory Commission, October, 2015. Accession Number:
ML15314A160.
176. R Kalikian and P Buckberg. St. Lucie Plant, Unit NOS. 1 and 2 - Review of License
Renewal Commitment for Pressurizer Surge Line Welds Inspection Program.
Washington, USA : U.S. Nuclear Regulatory Commission, October, 2016. Accession
Number: ML16235A138.
177. FPL. Submittal of Pressurizer Surge Line Wleds Inspection Program. s.l. : U.S. Nuclear
Regulatory Commission, May, 2012. Accession Number: ML12152A156.
178. M Homiack. Review by the Office of Nuclear Reactor Regulation Related to License
Renewal Commitment for Pressurizer Surge Line Welds Inspection Program Florida
Power & Light Company Turkey Point Nuclear Generating Units 3 and 4. Washington,
USA : U.S. Nuclear Regulatory Commission, May, 2013. Accession Number:
ML13141A595.
179. J Solin, S Reese and W Mayinger. Discussion on Fatigue Design Curves for Stainless
Steels. Baltimore, USA : ASME 2011 Pressure Vessels and Piping Division Conference
PVP2011, July, 2011. PVP2011-57943.

9-17
13321451
References

180. I V Papadopoulos and V P Panoskaltsis. Invariant Formulation of a Gradient Dependent


Multiaxial High-Cycle Fatigue Criterion. s.l. : Engineering Fracture Mechanics, 1996.
https://doi.org/10.1016/S0013-7944(96)00047-1.
181. M A Gray, M C Salac, D H Roarty and E L Cranford. Strain Rate Calculation Approach
in Environmental Fatigue Evaluations. s.l. : ASME Journal of Pressure Vessel
Technology, April, 2014. PVT-11-1095; doi: 10.1115/1.4026113.
182. T Métais, S Courtin, L De Baglion, C Gourdin and J C Le Roux. ASME Code-Case
Proposal to Explicitly Quantify the Interaction Between the PWR Environment and
Component Surface Finish. Waikoloa, USA : ASME 2017 Pressure Vessels and Piping
Conference PVP2017, July, 2017. PVP2017-65367.

9-18
13321451
A
TRANSLATED TABLE OF CONTENTS

DISCLAIMER OF WARRANTIES AND LIMITATION OF LIABILITIES


THIS DOCUMENT WAS PREPARED BY THE ORGANIZATION(S) NAMED BELOW AS AN ACCOUNT OF
WORK SPONSORED OR COSPONSORED BY THE ELECTRIC POWER RESEARCH INSTITUTE, INC.
(EPRI). NEITHER EPRI, ANY MEMBER OF EPRI, ANY COSPONSOR, THE ORGANIZATION(S) BELOW,
NOR ANY PERSON ACTING ON BEHALF OF ANY OF THEM:
(A) MAKES ANY WARRANTY OR REPRESENTATION WHATSOEVER, EXPRESS OR IMPLIED, (I) WITH
RESPECT TO THE USE OF ANY INFORMATION, APPARATUS, METHOD, PROCESS, OR SIMILAR ITEM
DISCLOSED IN THIS DOCUMENT, INCLUDING MERCHANTABILITY AND FITNESS FOR A PARTICULAR
PURPOSE, OR (II) THAT SUCH USE DOES NOT INFRINGE ON OR INTERFERE WITH PRIVATELY OWNED
RIGHTS, INCLUDING ANY PARTY'S INTELLECTUAL PROPERTY, OR (III) THAT THIS DOCUMENT IS
SUITABLE TO ANY PARTICULAR USER'S CIRCUMSTANCE; OR
(B) ASSUMES RESPONSIBILITY FOR ANY DAMAGES OR OTHER LIABILITY WHATSOEVER (INCLUDING
ANY CONSEQUENTIAL DAMAGES, EVEN IF EPRI OR ANY EPRI REPRESENTATIVE HAS BEEN ADVISED
OF THE POSSIBILITY OF SUCH DAMAGES) RESULTING FROM YOUR SELECTION OR USE OF THIS
DOCUMENT OR ANY INFORMATION, APPARATUS, METHOD, PROCESS, OR SIMILAR ITEM DISCLOSED
IN THIS DOCUMENT.

THE ELECTRIC POWER RESEARCH INSTITUTE (EPRI) PREPARED THIS REPORT.

13321451
A-1
13321451
Analyse des écarts de
connaissances sur la fatigue
environnementale (EAF)
Mise à jour et révision de l’analyse des écarts de
connaissances sur l’EAF
3002013214

Rapport final, septembre 2018

Chef de projet EPRI


K. Ahluwalia

Les exigences du programme d’assurance qualité


nucléaire EPRI (EPRI Nuclear Quality Assurance
Program) s’appliquent, en totalité ou en partie, à ce
produit.

ELECTRIC POWER RESEARCH INSTITUTE


3420 Hillview Avenue, Palo Alto, California 94304-1338 ▪ PO Box 10412, Palo Alto, California 94303-0813 ▪ États-Unis
(+1) 800.313.3774 ▪ (+1) 650.855.2121 ▪ askepri@epri.com ▪ www.epri.com

13321451
A-3
EXCLUSION DE GARANTIES ET LIMITATION DE RESPONSABILITÉ
LE PRÉSENT DOCUMENT A ÉTÉ PRÉPARÉ PAR LE OU LES ORGANISME(S) NOMMÉ(S) CI-
DESSOUS ET CONSTITUE UN COMPTE-RENDU DU TRAVAIL COMMANDITÉ ENTIÈREMENT OU
PARTIELLEMENT PAR L’ELECTRIC POWER RESEARCH INSTITUTE, INC. (EPRI). EPRI, TOUT
MEMBRE D’EPRI, TOUT COMMANDITAIRE, LE OU LES ORGANISME(S) CI-DESSOUS OU TOUTE
PERSONNE AGISSANT EN LEUR NOM :
(A) NE DOIVENT NI DONNER DE GARANTIE, NI FAIRE DE DÉCLARATION, EXPRESSE OU
IMPLICITE, (I) CONCERNANT L’UTILISATION D’INFORMATIONS, D’ÉQUIPEMENTS, DE
MÉTHODES, DE PROCESSUS OU TOUT ÉLÉMENT SIMILAIRE INDIQUÉ DANS LE PRÉSENT
DOCUMENT, Y COMPRIS LA QUALITÉ MARCHANDE ET L’ADÉQUATION À UN USAGE
PARTICULIER, OU (II) QU’UNE TELLE UTILISATION N’ENFREINT PAS OU N’INTERFÈRE PAS AVEC
DES DROITS DE PROPRIÉTÉ PRIVÉE, Y COMPRIS LA PROPRIÉTÉ INTELLECTUELLE D’UNE
QUELCONQUE DES PARTIES, OU (III) QUE LE PRÉSENT DOCUMENT CONVIENT TOUTE
SITUATION DANS LAQUELLE SE TROUVERAIT UN UTILISATEUR ; OU
(B) ASSUMENT LA RESPONSABILITÉ POUR TOUT DOMMAGE OU AUTRE MANQUEMENT QUEL
QU’IL SOIT (Y COMPRIS TOUT DOMMAGE CONSÉCUTIF, MÊME SI EPRI OU TOUT AUTRE
REPRÉSENTANT D’EPRI A ÉTÉ INFORMÉ DE L’ÉVENTUALITÉ DE TELS DOMMAGES)
DÉCOULANT DE VOTRE CHOIX OU UTILISATION DU PRÉSENT DOCUMENT OU DE TOUT
ÉQUIPEMENT, INFORMATION, MÉTHODE, PROCESSUS OU ÉLÉMENT SIMILAIRE INDIQUÉ DANS
LE PRÉSENT DOCUMENT.
TOUTE MENTION RELATIVE À UN PRODUIT, PROCÉDÉ OU SERVICE PAR SON APPELLATION
COMMERCIALE, SA MARQUE DE COMMERCE, SON FABRICANT OU AUTRE, NE CONSTITUE
NULLEMENT UNE APPROBATION, UNE RECOMMANDATION OU UNE PRÉFÉRENCE DE LA PART
D’EPRI.
CE RAPPORT A ÉTÉ PRÉPARÉ PAR L’ORGANISME SUIVANT, SOUS CONTRAT AVEC L’EPRI :
Amec Foster Wheeler

LE CONTENU TECHNIQUE DE CE PRODUIT N’A PAS ÉTÉ PRÉPARÉ CONFORMÉMENT AU


MANUEL DU PROGRAMME QUALITÉ D’EPRI QUI SATISFAIT AUX EXIGENCES DE LA
RÉGLEMENTATION 10 CFR 50, ANNEXE B. CE PRODUIT N’EST PAS SOUMIS AUX EXIGENCES
DE LA RÉGLEMENTATION 10 CFR PARTIE 21.

REMARQUE
Pour en savoir plus sur l’EPRI, appelez le centre d’assistance client EPRI au (+1) 800.313.3774 ou
envoyez un e-mail à askepri@epri.com.
Electric Power Research Institute, EPRI et TOGETHERSHAPING THE FUTURE OF ELECTRICITY
sont des marques de services déposées d’Electric Power Research Institute, Inc.
Copyright © 2018 Electric Power Research Institute, Inc. Tous droits réservés.

13321451
A-4
REMERCIEMENTS

Ce rapport a été préparé par l’organisme suivant, sous contrat avec l’Electric Power Research
Institute (EPRI) :
Amec Foster Wheeler
Booths Park, Chelford Road
Knutsford, Cheshire
WA16 8QZ
Angleterre
Enquêteur principal
D. Tice
Relecteurs
G. Quirk
N. Platts
A. Davison
Auteurs
A. McLennan
P. Gill

Ce rapport décrit une étude commanditée par EPRI. J. Smith et G. Stevens, membres du
personnel de l’EPRI, ont eux aussi participé à la préparation du rapport.

Cette publication est un document interne devant être cité dans toute documentation de la
manière suivante :
Analyse des écarts de connaissances sur la fatigue environnementale (EAF) : Mise à jour et
révision de l’analyse des écarts de connaissances sur l’EAF. EPRI, Palo Alto, CA : 2018.
3002013214.

13321451
A-5
RÉSUMÉ

Les ingénieurs qui souhaitent construire ou prolonger la durée de vie d’une centrale nucléaire
sont tenus de prouver la sécurité opérationnelle de la centrale sur sa durée de vie prévue. Ils
doivent généralement pour ce faire appliquer des courbes conceptuelles établies de l’amorçage
de la fatigue. Cependant, les règles existantes et les directives y afférentes peuvent prédire des
facteurs d’utilisation cumulatifs très élevés pour certains composants lorsque les effets de la
fatigue environnementale (EAF) sont pris en considération. Par ailleurs, les facteurs d’utilisation
calculés élevés et prédits ne sont pas observés dans l’expérience opérationnelle et, à ce jour,
aucune défaillance due à la fatigue n’a pu être attribuée à une fatigue thermique de cycle faible
accrue par la fatigue environnementale.
Pour comprendre l’applicabilité des règles de fatigue et des directives EAF actuellement
disponibles, l’EPRI a identifié certains écarts au niveau des connaissances sectorielles sur la
fatigue environnementale. Ces écarts ont été documentés dans le rapport d’analyse des écarts
publié en 2011 (produit 1023012 de l’EPRI) et dans le rapport de feuille de route publié en 2012
(produit 1026724 de l’EPRI). Ces documents exposaient certains domaines techniques souffrant
d’incertitudes et identifiaient les études et les activités de développement à mettre en œuvre pour
corriger la situation. Ce travail a donné lieu à une liste de 47 écarts, ou lacunes, de
connaissances, qui ont ensuite été examinés et hiérarchisés par des experts du secteur nucléaire.
Depuis la publication du rapport d’analyse des écarts, diverses études importantes ont été menées
dans le monde entier et plusieurs méthodologies alternatives d’évaluation de la fatigue
environnementale ont été proposées. Certaines de ces activités ont abordé quelques-uns des
écarts antérieurement identifiés.
Ce rapport présente en détail les résultats des études publiquement disponibles parues depuis la
préparation du rapport d’analyse des écarts initialement publié en 2011 et constitue ainsi une
mise à jour de cette analyse des écarts. Il révèle également que de nombreuses données de
laboratoire supplémentaires ont été générées pour les réacteurs à eau légère depuis la publication
du rapport initial.
Il convient de remarquer qu’il est nécessaire de réaliser un essai représentatif de centrale à l’aide
des données géométriques et des historiques de chargement propres à chaque centrale. De ce fait,
le développement et la mise en œuvre d’un essai de fatigue environnementale qui soit
représentatif des composants d’une centrale et qui passe par ses transitoires habituelles suscitent
actuellement beaucoup d’intérêt dans le monde entier, avec notamment le soutien d’un groupe
collaboratif international de parties intéressées dans ce domaine. Un tel essai permettrait de
recueillir des données précieuses et de valider toute méthode d’évaluation proposée.

13321451
A-6
La liste existante des 47 écarts a été passée en revue et révisée, donnant lieu à une liste de
17 problèmes nécessitant des études plus poussées. Cette nouvelle liste d’écarts a été dressée
dans un souci de simplification de l’ancienne liste, qui comportait des redondances et des
chevauchements, ainsi que pour orienter les écarts restants sur les principaux problèmes
identifiés par le secteur nucléaire.
Mots-clés
Croissance des fissures
EAF
Fatigue environnementale (Environmentally Assisted Fatigue)
Évaluation de la fatigue
Écarts (ou lacunes) de connaissances
Réacteur à eau légère

13321451
A-7
SYNTHÈSE

Numéro de livrable : 3002013214


Type de produit : Rapport technique
Titre de produit : Analyse des écarts de connaissances sur la fatigue environnementale
(EAF) : Mise à jour et révision de l’analyse des écarts de connaissances sur l’EAF

PUBLIC VISÉ : Concepteurs et opérateurs de centrales nucléaires


AUTRE PUBLIC VISÉ : Communauté internationale de la recherche nucléaire

OBJET ESSENTIEL DE LA RECHERCHE


L’évaluation initiale des écarts de connaissances (rapport 1023012 de l’EPRI) publiée en 2011 fournissait
une liste des écarts au niveau de la compréhension de la fatigue environnementale (EAF) au sein du secteur
nucléaire. Depuis la publication de ce document, de nombreux travaux supplémentaires ont été réalisés par
des instituts de recherche du monde entier afin de mieux comprendre ce mécanisme de dégradation, dans
l’objectif de découvrir des méthodes efficaces et moins conservatrices pour évaluer la durée de vie en fatigue
des composants des centrales nucléaires. Compte tenu des nombreuses études menées dans ce domaine,
le présent rapport a été préparé pour mettre à jour et définir de façon plus précise les écarts de connaissances
existants (comme antérieurement définis dans le rapport 1023012 de l’EPRI).

PRÉSENTATION DE LA RECHERCHE
L’effet nocif d’un réacteur à eau légère sur les matériaux testés en laboratoire est bien connu. L’applicabilité
des résultats obtenus en laboratoire, qui utilisent des échantillons à petite échelle sous chargement
pleinement inversé, sur les composants d’une centrale nucléaire réelle est toutefois loin d’être établie. La
fatigue environnementale présente actuellement un problème significatif lors de la conception d’une centrale
nucléaire, qu’il s’agisse d’une nouvelle construction ou d’une prolongation de sa durée de vie, en raison des
facteurs d’utilisation en fatigue élevés prédits par les méthodologies actuelles qui intègrent les facteurs
environnementaux (Fen). En raison des difficultés liées à la fatigue environnementale, la communauté
internationale a mené de nombreuses études dans ce domaine. Entre 2011 et 2012, des travaux ont été
réalisés pour évaluer la compréhension de l’EAF au sein du secteur et ont abouti à une liste d’écarts de
connaissances devant permettre aux études internationales d’offrir le maximum d’avantages au secteur. Ce
document a identifié 47 écarts liés à l’amorçage de la fatigue dans l’air et dans les réacteurs à eau légère, à
la croissance des fissures de fatigue dans les réacteurs à eau légère et aux méthodologies d’évaluation de
la fatigue environnementale. Compte tenu des nombreux travaux réalisés dans le domaine de l’EAF par le
secteur nucléaire depuis 2011, il s’est avéré nécessaire de mettre à jour et réviser le document d’analyse des
écarts de connaissances sur la fatigue environnementale. Ce rapport explique les travaux réalisés par rapport
aux écarts antérieurement identifiés, souligne les chevauchements et les redondances existant au niveau de
ces écarts, et révise les écarts à la lumière des informations recueillies.

PRINCIPALES CONCLUSIONS
Un examen détaillé des études menées depuis la publication initiale de l’analyse des écarts en 2011 a été
réalisé et a identifié un grand nombre de données de laboratoire supplémentaires générées pour les réacteurs

13321451
A-8
à eau légère. Les derniers travaux se sont concentrés sur les aciers inoxydables austénitiques dans les
principaux systèmes de refroidissement des centrales à eau pressurisée. La plupart des essais ont été
réalisés dans des conditions préparées pour correspondre de plus près aux conditions opérationnelles d’une
centrale, notamment en prenant en considération des facteurs tels que les transitoires complexes de
chargement, le chargement thermomécanique et les finitions de surface. Cependant, les données d’essai sur
les géométries et les dimensions des composants d’une centrale habituelle restent limitées. Les observations
tirées de ces études expérimentales avancées expliquent en partie les écarts observés entre les prédictions
provenant des modèles fondés sur les Fen et l’expérience pratique des composants d’une centrale. Des
modèles prédictifs améliorés ont été élaborés pour mieux décrire le comportement observé dans ces derniers
essais en laboratoire. Cependant, aucune validation complète des géométries et des historiques de
chargement d’une centrale n’est disponible.
À partir de ces données, la liste existante de 47 écarts de connaissances a été passée en revue et révisée,
donnant lieu à une liste plus courte des problèmes nécessitant des études plus poussées. Cette liste révisée
est présentée ci-après, les écarts de connaissances (KG, Knowledge Gaps) d’origine étant placés entre
parenthèses :
A. Problèmes généraux
1. Études supplémentaires sur l’endurance à la fatigue et la croissance des fissures de fatigue pour
améliorer les modèles d’évaluation de la fatigue environnementale et combler les écarts entre
les prédictions des modèles EAF et des procédures suivies dans les centrales (KG 2 et 4)
2. Base de données universelle à disposition des parties prenantes avec sélection des données
(KG 46)
B. Données sur l’endurance à la fatigue dans l’air et courbes conceptuelles de la fatigue
1. Travaux supplémentaires pour étudier les effets de la variabilité des matériaux et, le cas
échéant, pour élaborer des courbes conceptuelles de la fatigue par matériau (KG 9)
2. Révision des facteurs de transfert sur la durée de vie et la contrainte pour les courbes
conceptuelles de la fatigue (KG 3 et 12)
3. Données, meilleure compréhension et normalisation des méthodes d’évaluation de la fatigue
pour les soudures (KG 37 et 38)
C. Effets de l’environnement présent dans un réacteur à eau légère sur la conception fondée sur la
fatigue
1. Manque de données pour définir les équations Fen, surtout pour les seuils de température et de
sollicitation (KG 1, 5, 10, 13, 20, 21, 34 et 47)
2. Amélioration des méthodes conceptuelles EAF qui intègrent les différences entre les facteurs de
transfert entre un réacteur à eau légère et l’air (KG 19)
3. Amélioration des méthodes d’évaluation en tenant compte des effets de l’amplitude de la
sollicitation sur Fen (KG 7)
4. Effets de la contrainte moyenne sur la durée de vie en fatigue et la fatigue environnementale
(KG 6, 11, 14 et 44)
D. Influence de transitoires de chargement non isothermique et complexe sur la durée de vie en
fatigue dans les réacteurs à eau légère
1. Amélioration des méthodologies d’évaluation sous des chargements complexes et non
isothermiques correspondant mieux aux composants d’une centrale (KG 8, 18, 39, 40, 41, 42 et
43)
2. Données de validation sous des chargements correspondant à une centrale pour soutenir de
nouvelles méthodes d’évaluation, avec notamment applicabilité aux alliages autres que l’acier
inoxydable (KG 15, 16 et 17)
3. Amélioration des méthodologies de durée de vie totale pour intégrer l’influence des gradients de
contrainte dus aux transitoires de choc thermique (KG 16)
E. Croissance des fissures de fatigue dans les réacteurs à eau légère

13321451
A-9
1. Amélioration des courbes de croissance des fissures de référence pour les aciers inoxydables
austénitiques en intégrant les effets des seuils et des rapports R (y compris R négatif), et
application aux réacteurs à eau bouillante (composition chimique normale de l’eau et de l’eau
hydrogénée) (KG 22, 23, 24, 25 et 31)
2. Courbes de croissance des fissures de référence pour les aciers au carbone et faiblement alliés
dans les réacteurs à eau bouillante (KG 29 et 30)
3. Amélioration des courbes de croissance des fissures de référence pour les alliages de nickel
(KG 26, 27 et 32)
4. Méthodes d’évaluation de la croissance des fissures sous les chargements habituels dans une
centrale en reconnaissant l’effet des dommages durant le cycle de chargement (KG 28 et 45)
F. Essais de prototypes et de composants (KG 33, 35 et 36)

POURQUOI EST-CE IMPORTANT ?


Ce travail récapitule les études menées depuis la publication du document d’analyse des écarts initial. Ce
rapport est destiné à présenter de façon indépendante les études de pointe réalisées dans le domaine de la
fatigue environnementale. Il souligne également les liens entre les différentes techniques de recherche et
méthodes d’évaluation utilisées pour faciliter la collaboration et favoriser les progrès dans ce domaine.

COMMENT APPLIQUER LES RÉSULTATS


Le travail présenté dans ce produit peut servir à justifier et orienter les programmes de recherche auxquels
certains membres attachent de l’importance. Par ailleurs, ce document identifie les liens unissant les différents
travaux de recherche et développement expérimentaux sur les méthodes d’évaluation, ce qui pourrait
permettre aux membres de l’EPRI d’identifier plus facilement des possibilités de collaboration afin de
maximiser l’impact de leurs travaux.

POSSIBILITÉS D’APPRENTISSAGE ET D’ENGAGEMENT


• Des organisations de recherche internationales trouveront sans doute le rapport utile pour orienter les
études et les essais qu’elles réaliseront à l’avenir dans le domaine de la fatigue environnementale.

CONTACTS EPRI : Kawaljit (Al) Ahluwalia, responsable technique, kahluwal@epri.com


Jean Smith, responsable technique principal, jmsmith@epri.com
Gary Stevens, responsable technique, gstevens@epri.com

PROGRAMMES : Primary Systems Corrosion Research (PSCR), Program 41.01.01 ; Boiling Water Reactor
Vessel and Internals Project (BWRVIP), Program 41.01.03 ; Pressurized Water Reactor Materials Reliability
Program (MRP), Program 41.01.04

CATÉGORIE DE MISE EN ŒUVRE : Bases de référence-techniques

Together...Shaping the Future of Electricity®

Electric Power Research Institute


3420 Hillview Avenue, Palo Alto, California 94304-1338 • PO Box 10412, Palo Alto, California 94303-0813 États-Unis
(+1) 800.313.3774 • (+1) 650.855.2121 • askepri@epri.com • www.epri.com
© 2018 Electric Power Research Institute (EPRI), Inc. Tous droits réservés. Electric Power Research Institute, EPRI et
TOGETHER...SHAPING THE FUTURE OF ELECTRICITY sont des marques de services déposées d’Electric Power Research Institute, Inc.

13321451
A-10
TABLE DES MATIÈRES

RÉSUMÉ ................................................................................................................................... V

SYNTHÈSE ............................................................................................................................. VII

1 INTRODUCTION ..................................................................................................................1-1
1.1 Contexte ....................................................................................................................1-1
1.2 Hypothèses de la feuille de route ...............................................................................1-2
1.2.1 Hypothèses .......................................................................................................1-2
1.2.2 Recommandations, option 1 et 2 .......................................................................1-4
1.3 Objectifs de la mise à jour et de la révision de l’analyse des écarts de
connaissances sur l’EAF .....................................................................................................1-4
1.4 Structure du rapport ...................................................................................................1-5

2 DONNÉES SUR L’ENDURANCE À LA FATIGUE DANS L’AIR ..........................................2-1


2.1 Introduction ................................................................................................................2-1
2.2 Génération de données sur la fatigue et courbes de données moyennes ..................2-1
2.2.1 Écart de connaissances 9 (antérieurement classé comme catégorie C,
faible priorité) ..................................................................................................................2-1
2.2.2 Écart de connaissances 37 (antérieurement classé comme catégorie C,
faible priorité) ..................................................................................................................2-4
2.3 Facteurs de transfert ..................................................................................................2-6
2.3.1 Écart de connaissances 12 (antérieurement classé comme catégorie C,
priorité moyenne) ............................................................................................................2-7
2.3.2 Écart de connaissances 3 (antérieurement classé comme catégorie C,
haute priorité) ...............................................................................................................2-11
2.4 Bases de données sur la fatigue et la fatigue environnementale ..............................2-13
2.4.1 Écart de connaissances 46 (antérieurement classé comme catégorie C,
haute priorité) ...............................................................................................................2-13
2.5 Résumé des données sur l’endurance à la fatigue et des courbes conceptuelles
de la fatigue dans l’air........................................................................................................2-14

13321451
A-11
3 FATIGUE ENVIRONNEMENTALE DANS LES RÉACTEURS À EAU LÉGÈRE ..................3-1
3.1 Introduction ................................................................................................................3-1
3.2 Essais de fatigue sous chargement cyclique conventionnel .......................................3-2
3.2.1 Écart de connaissances 19 (antérieurement classé comme catégorie C,
priorité moyenne) ............................................................................................................3-2
3.2.2 Écarts de connaissances 10, 13 et 47 ...............................................................3-4
3.2.3 Écart de connaissances 20 (antérieurement classé comme catégorie C,
priorité moyenne) ..........................................................................................................3-11
3.2.4 Écart de connaissances 7 (antérieurement classé comme catégorie C,
haute priorité) ...............................................................................................................3-13
3.2.5 Écart de connaissances 34 (antérieurement classé comme catégorie A,
haute priorité) ...............................................................................................................3-15
3.2.6 Écart de connaissances 21 (antérieurement classé comme catégorie C,
faible priorité) ................................................................................................................3-15
3.3 Essais de formes d’ondes complexes ......................................................................3-17
3.3.1 Écart de connaissances 17 (antérieurement classé comme catégorie C,
haute priorité) ...............................................................................................................3-17
3.3.2 Écart de connaissances 18 (antérieurement classé comme catégorie B,
haute priorité) ...............................................................................................................3-18
3.3.3 Écart de connaissances 11 (antérieurement classé comme catégorie A,
faible priorité) ................................................................................................................3-23
3.3.4 Écart de connaissances 14 (antérieurement classé comme catégorie A,
priorité moyenne) ..........................................................................................................3-26
3.4 Chargement non isothermique .................................................................................3-27
3.4.1 Écarts de connaissances 15 et 16 ...................................................................3-27
3.5 Essais de prototypes et de composants ...................................................................3-32
3.5.1 Écart de connaissances 33 (antérieurement classé comme catégorie A,
haute priorité) ...............................................................................................................3-32
3.5.2 Écart de connaissances 36 (antérieurement classé comme catégorie C,
haute priorité) ...............................................................................................................3-36
3.6 Résumé de la fatigue environnementale dans les réacteurs à eau légère ................3-38

4CROISSANCE DES FISSURES DE FATIGUE ENVIRONNEMENTALE ...............................4-1


4.1 Introduction ................................................................................................................4-1
4.2 Essais de fatigue sous chargement cyclique conventionnel .......................................4-2
4.2.1 Écart de connaissances 22 (antérieurement classé comme catégorie C,
priorité moyenne) ............................................................................................................4-2
4.2.2 Écart de connaissances 24 (antérieurement classé comme catégorie C,
priorité moyenne) ............................................................................................................4-4

13321451
A-12
4.2.3 Écarts de connaissances 25 et 31 .....................................................................4-5
4.2.4 Écarts de connaissances 29 et 30 ...................................................................4-10
4.2.5 Écart de connaissances 26 (antérieurement classé comme catégorie C,
priorité moyenne) ..........................................................................................................4-11
4.2.6 Écart de connaissances 27 (antérieurement classé comme catégorie C,
priorité moyenne) ..........................................................................................................4-12
4.2.7 Écart de connaissances 32 (antérieurement classé comme catégorie C,
faible priorité) ................................................................................................................4-14
4.3 Formes d’ondes complexes et chargement non isothermique ..................................4-16
4.3.1 Écart de connaissances 28 (antérieurement classé comme catégorie A,
haute priorité) ...............................................................................................................4-16
4.4 Résumé des données sur la croissance des fissures ...............................................4-21

5 MÉTHODES D’ÉVALUATION ..............................................................................................5-1


5.1 Introduction ................................................................................................................5-1
5.1.1 Résumé de la compréhension mécanistique de la fatigue et de la fatigue
environnementale ...........................................................................................................5-1
5.1.2 Révision des concepts clés ...............................................................................5-2
5.1.3 Évaluation de l’amorçage de la fatigue ..............................................................5-3
5.1.4 Évaluation de la croissance des fissures de fatigue ...........................................5-4
5.1.5 Résumé des méthodes actuelles d’évaluation de la fatigue ...............................5-4
5.2 Amorçage de la fatigue ..............................................................................................5-9
5.2.1 Écart de connaissances 2 (antérieurement non classé) .....................................5-9
5.2.2 Écart de connaissances 4 (antérieurement classé comme catégorie B,
haute priorité) ...............................................................................................................5-12
5.2.3 Écart de connaissances 1 (antérieurement classé comme catégorie C,
haute priorité) ...............................................................................................................5-13
5.2.4 Écart de connaissances 5 (antérieurement classé comme catégorie C,
haute priorité) ...............................................................................................................5-14
5.3 Amorçage de la fatigue : analyse des formes d’ondes complexes ............................5-15
5.3.1 Écart de connaissances 8 (antérieurement classé comme catégorie C,
priorité moyenne) ..........................................................................................................5-15
5.3.2 Écarts de connaissances 39, 41, 42 et 43 .......................................................5-17
5.3.3 Écarts de connaissances 6 et 44 .....................................................................5-21
5.4 Croissance de fissure de fatigue ..............................................................................5-23
5.4.1 Écart de connaissances 23 (antérieurement classé comme catégorie C,
haute priorité) ...............................................................................................................5-23

13321451
A-13
5.4.2 Écart de connaissances 28 (antérieurement classé comme catégorie A,
haute priorité) ...............................................................................................................5-25
5.5 Soudures..................................................................................................................5-28
5.5.1 Écart de connaissances 38 (antérieurement classé comme catégorie C,
priorité moyenne) ..........................................................................................................5-28
5.6 Caractéristiques géométriques .................................................................................5-30
5.6.1 Écart de connaissances 40 (antérieurement classé comme catégorie C,
priorité moyenne) ..........................................................................................................5-30
5.7 Autres écarts de connaissances...............................................................................5-31
5.7.1 Écart de connaissances 35 (antérieurement classé comme catégorie B,
haute priorité) ...............................................................................................................5-31
5.7.2 Écart de connaissances 45 (antérieurement classé comme catégorie C,
priorité moyenne) ..........................................................................................................5-32
5.8 Synthèse ..................................................................................................................5-34

6 RÉSUMÉ DES OBSERVATIONS ET MISE À JOUR PROPOSÉE DES ÉCARTS DE


CONNAISSANCES .................................................................................................................6-1
6.1 Introduction ................................................................................................................6-1
6.2 État des écarts de connaissances d’origine................................................................6-2
6.2.1 Problèmes généraux .........................................................................................6-2
6.2.2 Données sur l’endurance à la fatigue dans l’air et courbes conceptuelles
de la fatigue ....................................................................................................................6-2
6.2.3 Effets de l’environnement présent dans un réacteur à eau légère sur la
conception fondée sur la fatigue .....................................................................................6-5
6.2.4 Influence de transitoires de chargement non isothermique et complexe sur
la durée de vie en fatigue dans les réacteurs à eau légère .............................................6-8
6.2.5 Essais de prototypes et de composants ..........................................................6-10
6.2.6 Croissance de fissure de fatigue......................................................................6-10
6.2.7 Méthodes d’évaluation de la fatigue environnementale....................................6-13

7 ÉCARTS DE CONNAISSANCES RATIONALISÉS ET CONSOLIDÉS ................................7-1


7.1 Introduction ................................................................................................................7-1
7.2 Problèmes généraux ..................................................................................................7-2
7.2.1 Écarts entre les prédictions des modèles EAF et des procédures suivies
dans les centrales ...........................................................................................................7-2
7.2.2 Base de données universelle à la disposition des parties prenantes..................7-3
7.3 Données sur l’endurance à la fatigue dans l’air et courbes conceptuelles de la
fatigue .................................................................................................................................7-3

13321451
A-14
7.3.1 Variabilité des matériaux et courbes conceptuelles de la fatigue pour
chaque matériau .............................................................................................................7-3
7.3.2 Facteurs de transfert pour les courbes conceptuelles de la fatigue ....................7-4
7.3.3 Données sur les métaux soudés et évaluation ...................................................7-4
7.4 Effets de l’environnement présent dans un réacteur à eau légère sur la
conception fondée sur la fatigue ..........................................................................................7-4
7.4.1 Manque de données pour définir les équations, les paramètres et les
seuils Fen.........................................................................................................................7-4
7.4.2 Applicabilité des facteurs de transfert conceptuel de la fatigue dans l’air
aux réacteurs à eau légère .............................................................................................7-5
7.4.3 Effets de l’amplitude de la sollicitation sur Fen ....................................................7-6
7.4.4 Effets de la contrainte moyenne sur la durée de vie en fatigue et la fatigue
environnementale ...........................................................................................................7-7
7.5 Influence de transitoires de chargement non isothermique et complexe sur la
durée de vie en fatigue dans les réacteurs à eau légère......................................................7-7
7.5.1 Données sur l’endurance en fatigue environnementale sous des
chargements complexes et non isothermiques ...............................................................7-7
7.5.2 Traitement des paramètres cycliquement variables dans une équation Fen .......7-8
7.5.3 Influence des gradients de contrainte dus aux transitoires de choc
thermique........................................................................................................................7-9
7.6 Croissance des fissures de fatigue dans les réacteurs à eau légère ..........................7-9
7.6.1 Courbes de croissance des fissures de référence pour les aciers
inoxydables austénitiques ...............................................................................................7-9
7.6.2 Courbes de croissance des fissures de référence pour les aciers au
carbone et les aciers faiblement alliés ............................................................................7-9
7.6.3 Courbes de croissance des fissures de référence pour les alliages de
nickel ........................................................................................................................7-10
7.6.4 Croissance des fissures sous un chargement habituel dans une centrale .......7-10
7.7 Essais de prototypes et de composants ...................................................................7-11

8 CONCLUSIONS ...................................................................................................................8-1

9 RÉFÉRENCES .....................................................................................................................9-1

13321451
A-15
LISTE DES FIGURES

Figure 5-1 Fonction de pondération SNW ..............................................................................5-20

13321451
A-16
LISTE DES TABLEAUX

Tableau 3-1 Paramètres Fen et seuils pour les aciers ferritiques...............................................3-7


Tableau 3-2 Paramètres pour l’équation Fen pour les aciers inoxydables EFEM ......................3-8
Tableau 3-3 Paramètres Fen et seuils pour les aciers austénitiques .........................................3-9
Tableau 3-4 Paramètres Fen et seuils pour les alliages de nickel............................................3-10
Tableau 6-1 Facteurs sur la durée de vie appliqués à la courbe de fatigue moyenne ε–N
dans l’air en tenant compte des effets de divers matériaux, chargements et
paramètres environnementaux .......................................................................................6-15
Tableau 8-1 Liste des nouveaux écarts de connaissances ......................................................8-3

13321451
A-17
13321451
環境誘起疲労 (EAF) 知識ギャップ分析
EAF 知識ギャップの更新と改訂
3002013214

最終報告書 2018 年 9 月

EPRI プロジェクトマネージャー
K. Ahluwalia

本製品には、EPRI 原子力品質保証プログラム (Nuclear


Quality Assurance Program) の要件の全てまたは一部が
適用される。

ELECTRIC POWER RESEARCH INSTITUTE


3420 Hillview Avenue, Palo Alto, California 94304-1338 ▪ PO Box 10412, Palo Alto, California 94303-0813 ▪ USA
+1.800.313.3774 ▪ +1.650.855.2121 ▪ askepri@epri.com ▪ www.epri.com

13321451
A-19
免責事項及び賠償責任の制限
本文書は、ELECTRIC POWER RESEARCH INSTITUTE, INC. (米国電力研究所:EPRI) の後援または
協賛による作業の報告書として、下記の組織によって作成されたものである。EPRI、EPRI の会員、
共同主催者、下記の団体、またはこれらの代理人のいずれも、

(A) (I) 本文書内で開示された、任意の情報、器具、方法、プロセスなどの、特定の目的のための市場性


や適合性を含めた利用に関して、あるいは (II) そのような利用がいかなる関係者の知的財産を含めた私
的に有する権利を侵害もしくは妨害しないことに関して、あるいは (III) 本文書が利用者の特定の状況
に相応しいことに関して、明示あるいは黙示を問わずいかなる保証や表明も行わないこと、または、

(B) 本文書の選択や使用、あるいは、本文書内で開示された任意の情報、器具、方法、プロセスなどに
起因するいかなる損害その他の責務 (たとえ EPRI または EPRI の代理人がそのような損害の可能性に
関する通知を受けていた場合でも、いかなる間接損害をも含めて) の責任を負わないものとする。

本文書内で行われた商品名、商標、製造者などによる具体的な商品、プロセス、サービスへの言及は、
必ずしも EPRI による支持、推奨、好意を意味または示唆するものではない。

本文書は、電力研究所 (EPRI) との契約に基づいて次の団体が作成した。


Amec Foster Wheeler

本製品の技術内容は、米国連邦規制基準 10 CFR 50、補遺 B の要件を満たす EPRI 品質計画書


(QUALITY PROGRAM MANUAL) に従って作成されたものではない。本製品は、米国連邦規制基準 10
CFR PART 21 の要件の対象ではない。

注記
EPRI に関する詳しい情報については EPRI お客様相談窓口 (電話番号 +1-800-313-3774、または
電子メール askepri@epri.com) までお問い合わせいただきたい。

Electric Power Research Institute 、 EPRI 、 お よ び TOGETHERSHAPING THE FUTURE OF


ELECTRICITY は Electric Power Research Institute, Inc. の登録サービスマークである。

Copyright © 2018 Electric Power Research Institute, Inc. 不許複製。

13321451
A-20
謝辞

本文書は、Electric Power Research Institute (EPRI) との契約の下に次の団体が作成した。


Amec Foster Wheeler
Booths Park, Chelford Road
Knutsford, Cheshire
WA16 8QZ
England
主任研究員
D. Tice
レビューア
G. Quirk
N. Platts
A. Davison
著者
A. McLennan
P. Gill

本報告書は、EPRI が後援する研究について説明するものである。EPRI 職員 J. Smith と


G. Stevens が本報告書の作成に寄与した。

本刊行物は、以下の形式によって引用されるべき企業文書である。
環境誘起疲労 (EAF) 知識ギャップ分析:EAF 知識ギャップの更新と改訂 EPRI, Palo
Alto, CA:2018. 3002013214.

13321451
A-21
要約

原子力発電所の長寿命化および新設を試みる設計者は、その意図された寿命にわたる発
電所の安全な運転を実証することが求められる。これには、通常、確立済みの疲労開始
設計曲線の適用が含まれる。ただし、既存の規則および関連する指針は、環境誘起疲労
(EAF) 効果を考慮すると、一部の構成機器について非常に高い累積使用率を予測しかね
ない。さらに、予測される高計算使用率は、運転経験には反映されておらず、現在に至
るまで EAF により強化された低サイクル熱疲労に起因するとされる疲労故障は知られ
ていない。
現在利用可能な疲労規則と EAF 指針の適用可能性を理解するために、EPRI は業界の
EAF 理解におけるギャップを特定した。これは、2011 年に発行された知識ギャップ分
析報告書 (EPRI 製品 1023012) および 2012 年に発行されたロードマップ報告書 (EPRI 製
品 1026724) で文書化された。これらの文書では、不確実性が存在する技術分野を明ら
かにして、それに対処するために必要な研究開発を特定した。その結果、47 件の知識
ギャップが明らかになった。これらの知識ギャップは、原子力業界の専門家によってレ
ビューされ、さまざまなギャップに優先順位が割り当てられた。
知識ギャップ報告書が発行されて以来、世界中で数多くの重要な研究活動が行われてお
り、EAF 評価のための代替手法がいくつか提案されている。これらの活動では、以前
確認されたいくつかの知識ギャップを具体的に解決したものもある。
本報告書では、元の知識ギャップが 2011 年に作成された後に公開され利用可能となっ
た調査結果の詳細なレビューを行い、知識ギャップの最新情報を提示する。また、前回
の報告時から軽水炉環境について重要な追加の実験室データが生成されたことも確認し
ている。
なお、発電所関連の形状および荷重履歴を用いた発電所の代表的試験の必要性が確認さ
れている。その結果、現在では、関係者から成る国際的な協力グループの支援を受け、
発電所の構成機器をより代表し、発電所に似た過渡現象を再現する EAF 試験の開発お
よび実行を進めることに世界的に関心が集まっている。このような試験により、重要な
追加の洞察と、提案された評価方法を検証する手段が提供されるであろう。
既存の 47 件の知識ギャップがレビューされ、さらなる研究が必要な 17 件の未解決問題
のより短いリストとして改訂された。新たな知識ギャップリストの理論的根拠は、冗長
性と重複を排除することにより既存のギャップを合理化し、また残りのギャップを原子
力業界により特定された主要な問題に集中させることであった。
キーワード
き裂進展 EAF 環境誘起疲労
疲労評価 知識ギャップ 軽水炉

13321451
A-22
エグゼクティブサマリー

納品番号:3002013214
製品タイプ:技術報告書

製品タイトル:環境誘起疲労 (EAF) 知識ギャップ分析:EAF 知識ギャップの更新と改訂

主な対象読者: 設計者および原子力発電所運転者
二次的な対象読者: 国際原子力研究コミュニティ

主な研究課題

最初の知識ギャップ評価 (EPRI 報告書 1023012) は 2011 年に作成され、環境誘起疲労 (EAF) に対する原子


力業界の理解におけるギャップのリストを示した。同文書が公表されて以来、原子力発電所の構成機器の
疲労寿命を評価するための有効で保守的すぎない方法を見出す目的で、この劣化メカニズムをさらに理解
するために、世界中の研究機関によって多くの追加作業が行われてきた。このトピックに関して行われた
多くの研究を踏まえて、本報告書は、既存の知識ギャップ (EPRI 報告書 1023012 で以前定義されたもの)
の更新と改訂を行うために作成された。

研究概要

実験室で試験が行われた材料に対する軽水炉 (LWR) 環境の有害な影響についてはよく知られている。しか


し、完全な逆負荷の下で小規模の試験片を用いた実験室での結果が、実際の原子力発電所の構成機器にど
のように適用できるかは不明である。環境要因 (Fen) を組み込んだ現在の方法では高疲労使用率が予測され
るため、EAF は現在、新設および発電所寿命延長の両方について発電所設計にとって重大な問題を提起し
ている。EAF により問題が提示された結果、国際コミュニティによって多くの研究が行われてきた。
2011/2012 年には、業界の EAF に対する理解を評価し、業界に最大限の利益をもたらすように国際的な研
究を導くための知識ギャップのリストを提示するための作業が行われた。知識ギャップ文書では、空気お
よび LWR 環境中での疲労発生、LWR 環境中での疲労き裂進展、および EAF 評価方法に関連して 47 件の
ギャップが見つかった。2011 年以来原子力業界が EAF に対して行ってきた大量の作業を考慮して、EAF
知識ギャップを更新しレビューする必要性が明らかとなった。本報告書では、以前の知識ギャップに関す
る作業の詳細を説明し、ギャップ間の重複と冗長性を明らかにし、収集した情報に基づいてギャップを修
正する。

主な所見

元の知識ギャップが 2011 年に作成されて以来の研究の詳細なレビューが行われ、重要な追加の実験室デー


タが LWR 環境中で得られたことがわかった。この最近の研究の主な焦点は、PWR の一次冷却材環境にお
けるオーステナイト系ステンレス鋼であった。この試験の多くは、複合負荷過渡現象、熱機械的負荷、表
面仕上げなどの要因に関して、発電所の運転条件をより厳密にシミュレートすることを目的とした条件下
で行われてきた。しかし、発電所の代表的な形状および代表的な構成機器のサイズに関する試験データは
まだ限られている。これらの高度な実験的研究の観察結果は、Fen ベースモデルからの予測値と発電所構成
機器の現場での経験値との間で観察された不一致について部分的な説明を提供するものである。これらの
最近の実験室試験で観察された挙動をより良く説明する改良された予測モデルが開発されている。ただし、
発電所を代表する構成機器の形状および負荷履歴の完全な検証はまだない。

13321451
A-23
エグゼクティブサマリー

これらのデータのレビュー結果に基づき、既存の 47 件の知識ギャップを精査、改訂し、追加の研究が必要
な未解決問題のより短いリストが作成された。この改訂されたリストは次のとおりである。関連する元の
知識ギャップ (KG) は括弧内で表示されている。
A. 一般的な問題
1. 改良された EAF 評価モデルをサポートし、EAF モデルによる予測と発電所手順との間の不一致
に対処するための疲労耐性と疲労き裂進展の両方に関する追加の研究 (KG 2 および 4)
2. 適切なデータ選別を伴う利害関係者の共通 EAF データベース (KG 46)
B. 空気中の疲労耐性データと疲労設計曲線
1. 材料変動の影響に対処し、適宜材料固有の疲労設計曲線を作成するための追加作業 (KG 9)
2. 設計疲労曲線の寿命と応力に関する修正済み移動係数 (KG 3 および 12)
3. 溶接部の疲労評価方法のデータ、理解の向上、標準化 (KG 37 および 38)
C. 疲労設計に対する LWR 環境の影響
1. Fen 式の定義に関連するデータ欠陥、特に温度および歪速度閾値 (KG 1、5、10、13、20、21、
34、47)
2. LWR 環境と空気の間の移動係数の違いを組み込んだ改良型 EAF 設計法 (KG 19)
3. Fen に対する歪振幅の影響を認める改良型評価法 (KG 7)
4. 平均応力が疲労寿命と EAF に与える影響 (KG 6、11、14、44)
D. LWR 環境における疲労寿命に対する複合かつ非等温の負荷過渡現象の影響
1. 発電所構成機器への関連性がより高い複合かつ非等温の負荷下での改良型評価法 (KG 8、18、
39、40、41、42、43)
2. ステンレス鋼以外の合金への適用可能性を含む、新たな評価法をサポートするための発電所に
関連する負荷下での検証データ (KG 15、16、17)
3. 熱衝撃過渡現象に対する歪勾配の影響を取り入れるための改良型全寿命法 (KG 16)
E. LWR 環境における疲労き裂進展
1. 閾値および R 比の影響 (負の R を含む) を組み込んだ、オーステナイト系ステンレス鋼の規準き
裂進展曲線の改良、および BWR (通常の水化学および水素水化学) 条件への拡張 (KG 22、23、
24、25、31)
2. BWR 環境における炭素鋼および低合金鋼の基準き裂進展曲線 (KG 29 および 30)
3. ニッケル基合金の改良型基準き裂進展曲線 (KG 26、27、32)
4. 負荷サイクルを通して損傷の重み付けを認める、発電所に関連する荷重下におけるき裂進展の
評価法 (KG 28 および 45)
F. プロトタイプ試験および構成機器試験 (KG 33、35、36)

13321451
A-24
エグゼクティブサマリー

本文書の重要性

本稿は、元の知識ギャップ文書の発行以降に行われた研究の概要を示す。本報告書は、EAF 研究に対する
最先端の独立文書の提供を意図している。また、EAF のトピックに関するより良い共同作業と進展を促進
するために、異なる研究技法と評価法との間の関連性に焦点を当てている。

研究結果の使い方

本成果物で示されている作業内容は、個々のメンバーにとって重要な研究プログラムを正当化して焦点を
合わせる助けとするために使用できる。さらに、本文書は、実験的研究と評価法の開発との関連性を示し
ている。これにより、EPRI メンバーは、それぞれの研究の影響を最大にするための共同作業の機会をより
よく特定することができる。

学習および参加の機会
• 国際的な研究機関は、本報告書が、環境誘起疲労の研究と試験のために考えられる将来の方向性に
とって有用と考えるかもしれない。

EPRI 連絡先: Kawaljit (Al) Ahluwalia (テクニカルエグゼクティブ)、kahluwal@epri.com


Jean Smith (主任テクニカルリーダー)、jmsmith@epri.com
Gary Stevens (テクニカルエグゼクティブ)、gstevens@epri.com

プログラム:一次系腐食研究 (PSCR)、プログラム 41.01.01;沸騰水型原子炉容器及び内部構造物プロジェ


クト (BWRVIP)、プログラム 41.01.03;加圧水型原子炉材料信頼性プログラム (MRP)、プログラム
41.01.04

実施カテゴリー: 参照–技術的根拠

Together…Shaping the Future of Electricity®(力を合わせて…電力の未来を創造する)

Electric Power Research Institute


3420 Hillview Avenue, Palo Alto, California 94304-1338 • PO Box 10412, Palo Alto, California 94303-0813, USA
+1.800.313.3774 • +1.650.855.2121 • askepri@epri.com • www.epri.com
© 2018 Electric Power Research Institute (EPRI), Inc. 不許複製。Electric Power Research Institute、EPRI、および
TOGETHER...SHAPING THE FUTURE OF ELECTRICITY は Electric Power Research Institute, Inc. の登録サービスマークである。

13321451
A-25
目次

概要 ........................................................................................................................................... V

エグゼクティブサマリー.......................................................................................................... VII

1 はじめに ...............................................................................................................................1-1
1.1 背景............................................................................................................................1-1
1.2 ロードマップ仮説 ......................................................................................................1-2
1.2.1 仮説 ...................................................................................................................1-2
1.2.2 推奨事項、オプション 1 および 2 .....................................................................1-4
1.3 EAF 知識ギャップの更新と改訂の目的 ......................................................................1-4
1.4 報告書の構成 .............................................................................................................1-5

2 空気中の疲労耐性データ.......................................................................................................2-1
2.1 はじめに ....................................................................................................................2-1
2.2 疲労データの生成と平均データ曲線 ..........................................................................2-1
2.2.1 知識ギャップ 9 (以前の分類はカテゴリー C、低優先度)...................................2-1
2.2.2 知識ギャップ 37 (以前の分類はカテゴリー C、低優先度) .................................2-4
2.3 移動係数 ....................................................................................................................2-6
2.3.1 知識ギャップ 12 (以前の分類はカテゴリー C、中優先度) .................................2-7
2.3.2 知識ギャップ 3 (以前の分類はカテゴリー C、高優先度).................................2-11
2.4 疲労および環境疲労のデータベース ........................................................................2-13
2.4.1 知識ギャップ 46 (以前の分類はカテゴリー C、高優先度) ...............................2-13
2.5 空気中の疲労耐性データと設計曲線の概要..............................................................2-14

3 LWR 環境疲労 ......................................................................................................................3-1


3.1 はじめに ....................................................................................................................3-1
3.2 従来の繰返し負荷疲労試験 ........................................................................................3-2
3.2.1 知識ギャップ 19 (以前の分類はカテゴリー C、中優先度) .................................3-2

13321451
A-26
3.2.2 知識ギャップ 10、13、47 .................................................................................3-4
3.2.3 知識ギャップ 20 (以前の分類はカテゴリー C、中優先度) ...............................3-11
3.2.4 知識ギャップ 7 (以前の分類はカテゴリー C、高優先度).................................3-13
3.2.5 知識ギャップ 34 (以前の分類はカテゴリー A、高優先度) ...............................3-15
3.2.6 知識ギャップ 21 (以前の分類はカテゴリー C、低優先度) ...............................3-15
3.3 複合波形試験 ...........................................................................................................3-17
3.3.1 知識ギャップ 17 (以前の分類はカテゴリー C、高優先度) ...............................3-17
3.3.2 知識ギャップ 18 (以前の分類はカテゴリー B、高優先度) ...............................3-18
3.3.3 知識ギャップ 11 (以前の分類はカテゴリー A、低優先度) ...............................3-23
3.3.4 知識ギャップ 14 (以前の分類はカテゴリー A、中優先度) ...............................3-26
3.4 非等温装填 ...............................................................................................................3-27
3.4.1 知識ギャップ 15 および 16..............................................................................3-27
3.5 プロトタイプ試験および構成機器の試験 .................................................................3-32
3.5.1 知識ギャップ 33 (以前の分類はカテゴリー A、高優先度) ...............................3-32
3.5.2 知識ギャップ 36 (以前の分類はカテゴリー C、高優先度) ...............................3-36
3.6 LWR 環境疲労の概要 ...............................................................................................3-38

4 環境疲労き裂進展 .................................................................................................................4-1
4.1 はじめに ....................................................................................................................4-1
4.2 従来の繰返し負荷疲労試験 ........................................................................................4-2
4.2.1 知識ギャップ 22 (以前の分類はカテゴリー C、中優先度) .................................4-2
4.2.2 知識ギャップ 24 (以前の分類はカテゴリー C、中優先度) .................................4-4
4.2.3 知識ギャップ 25 および 31................................................................................4-5
4.2.4 知識ギャップ 29 および 30..............................................................................4-10
4.2.5 知識ギャップ 26 (以前の分類はカテゴリー C、中優先度) ...............................4-11
4.2.6 知識ギャップ 27 (以前の分類はカテゴリー C、中優先度) ...............................4-12
4.2.7 知識ギャップ 32 (以前の分類はカテゴリー C、低優先度) ...............................4-14
4.3 複合波形と非等温装填 .............................................................................................4-16
4.3.1 知識ギャップ 28 (以前の分類はカテゴリー A、高優先度) ...............................4-16
4.4 クラック進展データの概要 ......................................................................................4-21

5 評価法 ...................................................................................................................................5-1
5.1 はじめに ....................................................................................................................5-1

13321451
A-27
5.1.1 疲労と EAF の機構的理解の概要 .......................................................................5-1
5.1.2 主な概念のレビュー ..........................................................................................5-2
5.1.3 疲労開始評価 .....................................................................................................5-3
5.1.4 疲労き裂進展評価 ..............................................................................................5-4
5.1.5 現在の疲労評価法の概要 ...................................................................................5-4
5.2 疲労開始 ....................................................................................................................5-9
5.2.1 知識ギャップ 2 (以前の分類はなし) ..................................................................5-9
5.2.2 知識ギャップ 4 (以前の分類はカテゴリー B、高優先度) .................................5-12
5.2.3 知識ギャップ 1 (以前の分類はカテゴリー C、高優先度).................................5-13
5.2.4 知識ギャップ 5 (以前の分類はカテゴリー C、高優先度).................................5-14
5.3 疲労開始:複合波形分析 ..........................................................................................5-15
5.3.1 知識ギャップ 8 (以前の分類はカテゴリー A、中優先度) .................................5-15
5.3.2 知識ギャップ 39、41、42、43........................................................................5-17
5.3.3 知識ギャップ 6 および 44 ...............................................................................5-21
5.4 疲労き裂進展 ...........................................................................................................5-23
5.4.1 知識ギャップ 23 (以前の分類はカテゴリー C、高優先度) ...............................5-23
5.4.2 知識ギャップ 28 (以前の分類はカテゴリー A、高優先度) ...............................5-25
5.5 溶接部 ......................................................................................................................5-28
5.5.1 知識ギャップ 38 (以前の分類はカテゴリー C、中優先度) ...............................5-28
5.6 形状的特徴 ...............................................................................................................5-30
5.6.1 知識ギャップ 40 (以前の分類はカテゴリー C、中優先度) ...............................5-30
5.7 他の知識ギャップ ....................................................................................................5-31
5.7.1 知識ギャップ 35 (以前の分類はカテゴリー B、高優先度) ...............................5-31
5.7.2 知識ギャップ 45 (以前の分類はカテゴリー C、中優先度) ...............................5-32
5.8 まとめ ......................................................................................................................5-34

6 知識ギャップの観察結果と提案された更新の概要 ................................................................6-1
6.1 はじめに ....................................................................................................................6-1
6.2 元の知識ギャップの状況 ............................................................................................6-2
6.2.1 一般的な問題 .....................................................................................................6-2
6.2.2 空気中の疲労耐性データと疲労設計曲線...........................................................6-2
6.2.3 疲労設計に対する LWR 環境の影響 ..................................................................6-5

13321451
A-28
6.2.4 LWR 環境における疲労寿命に対する複合かつ非等温の負荷過渡状態の影響 ....6-8
6.2.5 プロトタイプ試験および構成機器の試験.........................................................6-10
6.2.6 疲労き裂進展 ...................................................................................................6-10
6.2.7 EAF 評価法......................................................................................................6-13

7 合理化され統合された知識ギャップ .....................................................................................7-1
7.1 はじめに ....................................................................................................................7-1
7.2 一般的な問題 .............................................................................................................7-2
7.2.1 EAF モデルによる予測と発電所手順との不一致 ...............................................7-2
7.2.2 利害関係者の共通データベース .........................................................................7-3
7.3 空気中の疲労耐性データと疲労設計曲線 ...................................................................7-3
7.3.1 材料の変動性と材料固有の疲労設計曲線...........................................................7-3
7.3.2 設計疲労曲線の移動係数 ...................................................................................7-4
7.3.3 溶接金属データと評価 .......................................................................................7-4
7.4 疲労設計に対する LWR 環境の影響 ...........................................................................7-4
7.4.1 Fen 式、パラメータおよび閾値を定義するデータの不足 ...................................7-4
7.4.2 空気疲労設計移動係数の LWR 環境への適用可能性 ..........................................7-5
7.4.3 Fen に対する歪振幅の影響 .................................................................................7-6
7.4.4 平均応力が疲労寿命と EAF に与える影響 .........................................................7-7
7.5 LWR 環境における疲労寿命に対する複合かつ非等温の負荷過渡状態の影響 ............7-7
7.5.1 複合かつ非等温の負荷下での EAF 耐性データ ..................................................7-7
7.5.2 Fen 式における周期的可変パラメータの取り扱い ..............................................7-8
7.5.3 熱衝撃過渡状態に対する歪勾配の影響 ..............................................................7-9
7.6 LWR 環境における疲労き裂進展 ...............................................................................7-9
7.6.1 オーステナイト系ステンレス鋼の基準き裂進展曲線 .........................................7-9
7.6.2 炭素および低合金鋼の基準き裂進展曲線...........................................................7-9
7.6.3 ニッケル基合金の基準き裂進展曲線................................................................7-10
7.6.4 発電所関連負荷下のき裂進展 ..........................................................................7-10
7.7 プロトタイプ試験および構成機器の試験 .................................................................7-11

8 結論 ......................................................................................................................................8-1

9 参考文献 ...............................................................................................................................9-1

13321451
A-29
図一覧

図 5-1 SNW 重み関数.............................................................................................................5-20

13321451
A-30
表一覧

表 3-1 フェライト鋼の Fen パラメータとその閾値....................................................................3-7


表 3-2 EFEM ステンレス鋼 Fen 式のパラメーター....................................................................3-8
表 3-3 オーステナイト鋼の Fen パラメータとその閾値 ............................................................3-9
表 3-4 ニッケル基合金の Fen パラメーターとその閾値 ..........................................................3-10
表 6-1 さまざまな材料、負荷、および環境パラメータの影響を考慮した平均疲労 ε-N 空
気曲線に適用される寿命の因子 .....................................................................................6-15
表 8–1 新たな知識ギャップの一覧 ...........................................................................................8-3

13321451
A-31
13321451
환경 유도 피로(EAF) 지식 격차 분석
EAF 지식 격차 업데이트 및 개정
3002013214

최종 보고서, 2018 년 9 월

EPRI 프로젝트 매니저


K. Ahluwalia

본 결과물에는 EPRI 원자력 품질보증 프로그램의 전체


또는 일부 요건이 적용됩니다.

ELECTRIC POWER RESEARCH INSTITUTE


3420 Hillview Avenue, Palo Alto, California 94304-1338 ▪ PO Box 10412, Palo Alto, California 94303-0813 ▪ USA
+1.800.313.3774 ▪ +1.650.855.2121 ▪ askepri@epri.com ▪ www.epri.com

13321451
A-33
보증 부인 및 책임 제한
본 문서는 EPRI(ELECTRIC POWER RESEARCH INSTITUTE, INC.)의 지원 또는 공동 지원을 받은 하기
기관(들)의 연구로 작성되었습니다. EPRI, EPRI 의 회원사, 공동 후원사, 하기 기관(들) 또는 이들을
대리하는 어떠한 사람도

(A) (I) 상품성 및 특정 용도 적합성을 포함하여 본 문서에 공개된 어떠한 정보, 장치, 방법, 과정 또는 유사
항목에 대해, 또는 (II) 이러한 사용이 당사자의 지적 재산권을 포함한 사적 소유권을 침해하거나 방해하지
않는다는 점에 대해, 또는 (III) 본 문서가 특정 사용자의 상황에 적합하다는 점에 대해 명시적 또는
묵시적으로 보증하거나 진술하지 않습니다.

(B) (EPRI 또는 EPRI 대리자가 해당 손해 가능성을 미리 인지한 경우라도, 간접 손해를 포함한) 본 문서


또는 정보, 장치, 방법, 과정이나 본 문서에 공개된 유사 항목의 선택 또는 사용으로 인해 발생한 어떠한
손해나 법적 책임에 대해서도 책임을 지지 않습니다.

본 문서에서 특정 상품, 과정 또는 서비스가 상품명, 상표, 제조자 등으로 언급된다고 해서 EPRI 가 반드시
이를 지지, 추천, 선호한다는 의미로 해석할 수 없습니다.

EPRI 와의 계약에 따라 본 문서를 작성한 기관은 다음과 같습니다.


Amec Foster Wheeler

본 결과물의 기술적 내용은 부록 B, 10 CFR 50 의 요건을 이행하는 EPRI 품질 프로그램 매뉴얼에 따라


작성되지 않았습니다. 본 결과물은 10 CFR PART 21 의 요건이 적용되지 않습니다.

참고
EPRI 에 대한 자세한 정보는 EPRI 고객지원센터에 +1.800.313.3774 로 전화하거나
askepri@epri.com 으로 이메일을 보내시기 바랍니다.

Electric Power Research Institute, EPRI, TOGETHERSHAPING THE FUTURE OF ELECTRICITY 는


Electric Power Research Institute, Inc.의 등록 서비스 마크입니다.

저작권 © 2018 Electric Power Research Institute, Inc. 모든 권리 보유.

13321451
A-34
감사의 글

미국전력연구소(EPRI)와의 계약에 따라 본 문서를 작성한 기관은 다음과 같습니다.


Amec Foster Wheeler
Booths Park, Chelford Road
Knutsford, Cheshire
WA16 8QZ
England
책임 연구자
D. Tice
검토자
G. Quirk
N. Platts
A. Davison
작성자
A. McLennan
P. Gill

본 보고서의 연구 내용은 EPRI 의 지원을 받아 수행한 것입니다. EPRI 직원 J. Smith 와


G. Stevens 도 본 보고서의 작성에 기여했습니다.

본 간행물은 회사 문서로서 참고 문헌으로 인용할 때는 다음과 같은 형식을 따라야


합니다.
환경 유도 피로(EAF) 지식 격차 분석: EAF 지식 격차 업데이트 및 개정. EPRI, Palo Alto,
CA: 2018. 3002013214.

13321451
A-35
초록

원자력 발전소 수명 연장 및 신규 건설을 시도하는 설계자는 의도한 수명 기간에 대한


발전소의 안전한 운전을 입증해야 합니다. 일반적으로 이 과정에는 이미 확립되어 있는
피로 발생 설계 곡선의 적용이 수반됩니다. 그러나 기존 규칙 및 관련 지침에 따라 환경
유도 피로(EAF)의 영향을 고려할 때 일부 컴포넌트의 누적 사용 계수가 매우 높을 것으로
예측할 수 있습니다. 또한 높은 측정값이 예측되는 사용 계수는 운용 경험에 반영되지
않으며 EAF 를 통해 높아진 저사이클 열 피로로 인한 피로 손상도 지금까지 알려진 바가
없습니다.
이에 EPRI 는 현재 사용할 수 있는 피로 규칙 및 EAF 지침의 적용 가능성을 파악하고자
원자력 업계의 EAF 이해도 격차를 확인해 보았습니다. 이 내용은 2011 년 발행된 지식
격차 분석 보고서(EPRI 제품 1023012) 및 2012 년 발행된 로드맵 보고서(EPRI 제품
1026724)에 기록되어 있습니다. 이들 문서에서는 불확실성이 존재하는 기술적 영역을
집중 조명하며 이를 해결하기 위해 필요한 연구 및 개발을 규명했습니다. 그 결과 원자력
업계 전문가들의 검토를 거쳐 47 개 지식 격차가 확인되었으며 전문가들은 이러한 다양한
격차에 우선 순위를 할당했습니다.
지식 격차 보고서가 발행된 이후 전 세계적으로 중요한 연구 활동이 많이 이루어졌고 EAF
평가를 대체할 방법도 여러 가지 제안되었습니다. 앞서 확인된 지식 격차의 일부를
구체적으로 다룬 활동도 있었습니다.
이 보고서에서는 2011 년 지식 격차 원안이 개발된 이후 발표된 연구 중 공개적으로
사용할 수 있는 연구 결과의 상세 검토 내용을 비롯하여 지식 격차에 대한 업데이트
정보를 제공합니다. 또한 이전 보고서 이후로 경수로(LWR) 환경에 대해 상당한 추가 실험
데이터가 생성된 점을 밝히고 있습니다.
주목해야 할 부분은 발전소 연관 형상 및 하중 이력을 사용하는 발전소 표본 테스트의
필요성이 확인되었다는 사실입니다. 그 결과, 이해관계자로 구성된 국제 협력 집단의
지원으로 발전소 컴포넌트를 보다 잘 대표하는 표본이 되며 발전소와 유사한 과도상태를
거치는 EAF 테스트의 개발 및 수행에 대한 전 세계적 관심이 매우 높아졌습니다. 이러한
테스트를 통해 상당한 추가 통찰력을 비롯하여 제안된 평가 방법의 검증 수단을 얻을 수
있을 것입니다.
기존의 47 개 지식 격차는 검토와 수정을 거쳐 추가 연구의 타당성을 입증하는 17 가지
미해결 문제로 간추린 짧은 목록이 개발되었습니다. 새로운 지식 격차 목록을 마련한
이유는 중복 및 중첩 항목을 제거하고 원자력 업계에서 식별된 주요 문제와 관련하여
여전히 존재하는 격차에 초점을 맞추어 기존 격차 목록을 간소화하기 위해서였습니다.

13321451
A-36
키워드
균열 성장 EAF 환경 유도 피로
피로 평가 지식 격차 경수로(LWR)

13321451
A-37
주요 개요

결과물 번호: 3002013214


결과물 유형: 기술 보고서

결과물 제목: 환경 유도 피로(EAF) 지식 격차 분석: EAF 지식 격차 업데이트 및 개정

1 차 대상: 설계자 및 원자력 발전소 운영자


2 차 대상: 전 세계 원자력 연구 커뮤니티

핵심 연구 과제

2011 년 생성된 지식 격차 평가(EPRI 보고서 1023012) 원안에서는 환경 유도 피로(EAF)에 대한 원자력 업계의


이해도 격차 목록이 제공되었습니다. 해당 문서가 발행된 이후로 전 세계 연구 기관들이 원자력 발전소
컴포넌트의 피로 수명 평가에 효과적이면서도 덜 보수적인 방법을 찾으려는 목적으로 상당한 추가 연구를
수행하여 열화 메커니즘에 대해 한층 더 깊이 이해할 수 있게 되었습니다. 이 주제에 대해 상당한 분량의 연구가
이루어진 점을 고려하여, 앞서 EPRI 보고서 1023012 에서 정의한 바 있는 기존의 지식 격차에 대한 업데이트
및 개선 정보를 제공하고자 본 보고서를 준비했습니다.

연구 개요

실험실 테스트를 통해 경수로(LWR) 환경이 재료에 미치는 유해한 영향은 잘 알려져 있습니다. 하지만
실험실에서 완전한 역하중 조건으로 소규모 시편을 사용하여 도출해낸 결과를 실제 원자력 발전소 컴포넌트에
적용할 방법은 분명하지 않습니다. 현재 EAF 는 원자력 발전소 설계의 신규 건설 및 발전소 수명 연장 측면에서
상당한 문제가 되고 있습니다. 환경 계수(Fen)를 포함하는 현행 방법론에 따르면 피로 사용 계수가 높을 것으로
예측되기 때문입니다. EAF 로 인해 발생하는 문제들로 인해 전 세계에서 상당한 분량의 연구가 이루어졌습니다.
2011 년과 2012 년에는 EAF 에 대한 업계의 이해도를 평가하고 업계의 편익을 극대화할 수 있도록 전 세계
연구의 길잡이가 되어주는 지식 격차 목록을 제시하는 연구가 진행되었습니다. 지식 격차 문서에서는 대기 중
및 LWR 환경에서의 피로 발생, LWR 환경에서의 피로 균열 성장, EAF 평가 방법과 관련된 47 개의 격차를
확인할 수 있었습니다. 2011 년 이후 원자력 업계에서 EAF 에 대해 수많은 연구가 이루어진 점을 고려할 때
EAF 지식 격차에 대한 업데이트 및 검토의 필요성을 확인할 수 있었습니다. 본 보고서에서는 이전의 지식
격차와 관련하여 수행된 연구에 대해 자세히 설명하고 각 격차 간 중첩 및 중복 항목을 집중 조명하며 수집된
정보를 바탕으로 격차 항목을 수정합니다.

핵심 결과

2011 년 지식 격차 원안이 개발된 이후의 연구에 대한 상세 검토를 수행한 결과, LWR 환경에서 상당한 추가
실험 데이터가 개발되었음이 밝혀졌습니다. 이러한 최신 연구에서 특히 중점을 둔 부분은 PWR 1 차 냉각재
환경의 오스테나이트 스테인리스강이었습니다. 이들 테스트의 상당 부분은 복합 하중 과도상태, 열-기계학적
하중, 표면 마감 등의 계수 측면에서 발전소 운전 조건을 더욱 면밀하게 시뮬레이션하도록 의도한 조건에서
수행되었습니다. 그러나 발전소 표본 형상 및 표본 컴포넌트 크기에 대한 테스트 데이터는 여전히
제한적입니다. 이들 고급 실험 연구의 관찰 결과를 통해 Fen 기반 모델을 사용한 예측과 발전소 컴포넌트 현장
경험 간에 차이가 발견되는 이유를 부분적으로 이해할 수 있을 것입니다. 또한 개선된 예측 모델이 개발되면서
최근의 실험 결과에서 관찰된 동작에 대해 더욱 분명한 설명을 제공하고 있습니다. 그러나 발전소 표본
컴포넌트 형상과 하중 이력에 대한 완전한 검증은 제공되지 않습니다.

13321451
A-38
주요 개요

이러한 데이터의 검토를 기반으로 기존의 47 개 지식 격차에 대한 검토와 수정을 거쳐 추가 연구의 타당성을
입증하는 미해결 문제만 간추린 짧은 목록을 개발했습니다. 이렇게 수정된 목록이 아래에 나와 있으며 지식
격차(KG) 원안의 관련 항목은 괄호 안에 표시되어 있습니다.
A. 일반적 문제
1. 피로 내구성 및 피로 균열 성장에 대한 추가 연구를 통해 EAF 평가 모델 개선을 지원하고 EAF
모델을 사용한 예측과 발전소 절차 간의 차이 문제를 해결(KG 2, 4)
2. 적절한 데이터 스크리닝이 이루어지는 전 세계 관계자들의 EAF 데이터베이스(KG 46)
B. 대기 중 및 피로 설계 곡선상의 피로 내구성 데이터
1. 재료 변동성 영향을 다루며 해당되는 경우 재료별 피로 설계 곡선을 개발하는 추가 연구(KG 9)
2. 설계 피로 곡선에 대한 수명 및 응력 관련 이동 계수 수정(KG 3, 12)
3. 용접에 대한 피로 평가 방법 관련 데이터 및 이해도 개선, 표준화(KG 37, 38)
C. LWR 환경이 피로 설계에 미치는 영향
1. Fen 방정식, 특히 온도 및 변형율 임계값 정의와 관련된 데이터 결함(KG 1, 5, 10, 13, 20, 21, 34,
47)
2. LWR 환경 및 대기 중의 이동 계수 간 차이를 포함하는 EAF 설계 방법 개선(KG19)
3. 변형 진폭이 Fen 에 미치는 영향을 파악하는 평가 방법 개선(KG 7)
4. 평균 응력이 피로 수명 및 EAF 에 미치는 영향(KG 6, 11, 14, 44)
D. 복합 및 비등온 하중 과도상태가 LWR 환경에서 피로 수명에 미치는 영향
1. 발전소 컴포넌트와 더욱 관련도 높은 복합 및 비등온 하중 조건에서의 평가 방법 개선(KG 8, 18,
39, 40, 41, 42, 43)
2. 스테인리스강 이외의 합금에 대한 적용성을 포함하여 새로운 평가 방법을 뒷받침하는 발전소
연관 하중 조건에서의 검증 데이터(KG 15, 16, 17)
3. 열충격 과도상태에 대한 변형 구배의 영향을 포함하도록 총 수명 관련 방법론 개선(KG 16)
E. LWR 환경에서의 피로균열 성장
1. 임계값 및 R 비율의 영향(음성 R 포함)을 비롯하여 BWR 에 대한 확장(정상 수질 화학 및 수소
수질 화학) 조건을 포함하도록 오스테나이트 스테인리스강에 대한 준거 균열 성장 곡선 개선(KG
22, 23, 24, 25, 31)
2. BWR 환경에서의 탄소강 및 저합금강에 대한 준거 균열 성장 곡선(KG 29, 30)
3. 니켈 합금에 대한 준거 균열 성장 곡선 개선(KG 26, 27, 32)
4. 하중 주기에 걸친 손상의 가중치를 파악할 수 있는 발전소 연관 하중 조건에서의 균열 성장 평가
방법(KG 28, 45)
F. 프로토타입 및 컴포넌트 테스트(KG 33, 35, 36)

13321451
A-39
주요 개요

중요한 이유

본 연구에서는 지식 격차 원안 문서의 발행 이후 수행된 연구를 요약하여 제공합니다. 본 보고서의 목적은 최신


EAF 연구에 대해 별도의 설명을 제공하는 것입니다. 또한 보고서를 통해 다양한 연구 기법 및 평가 방법 간의
연관성을 집중 조명하여 EAF 라는 주제에 대해 더욱 원활한 협업 및 연구 진척을 도모하고자 합니다.

결과 적용 방법

본 결과물에 제시된 연구 내용은 개별 회원에게 중요한 연구 프로그램에 초점을 맞추고 타당성을 확인하는 데
도움이 될 수 있습니다. 또한 이 문서를 통해 실험 연구와 평가 방법 개발 사이의 연관성을 알 수 있으므로 EPRI
회원이 협업 기회를 더욱 분명하게 파악하여 연구 효과를 극대화할 수 있습니다.

학습 및 참여 기회
• 본 보고서는 전 세계 연구 기관에서 환경 유도 피로 연구 및 테스트의 향후 방향 설정 시 유용할 수
있습니다.

EPRI 연락처: Kawaljit (Al) Ahluwalia, Technical Executive, kahluwal@epri.com


Jean Smith, Principal Technical Leader, jmsmith@epri.com
Gary Stevens, Technical Executive, gstevens@epri.com

프로그램: 1 차계통 부식 연구(PSCR): 프로그램 41.01.01, 비등수형 원자로 용기 및 내부구조물


프로젝트(BWRVIP): 프로그램 41.01.03, 가압경수로 재료 신뢰성 프로그램(MRP): 프로그램 41.01.04

시행 범주: 참고 자료–기술적 근거

Together...Shaping the Future of Electricity®

Electric Power Research Institute


3420 Hillview Avenue, Palo Alto, California 94304-1338 • PO Box 10412, Palo Alto, California 94303-0813 USA
+1.800.313.3774 • +1.650.855.2121 • askepri@epri.com • www.epri.com
© 2018 Electric Power Research Institute (EPRI), Inc. 모든 권리 보유. Electric Power Research Institute, EPRI 및
TOGETHER...SHAPING THE FUTURE OF ELECTRICITY 는 Electric Power Research Institute, Inc.의 등록 서비스 마크입니다.

13321451
A-40
목차

초록............................................................................................................................................ V

주요 개요.................................................................................................................................. VII

1 서론.......................................................................................................................................1-1
1.1 배경............................................................................................................................1-1
1.2 로드맵 가설 ................................................................................................................1-2
1.2.1 가설 ...................................................................................................................1-2
1.2.2 권고사항(옵션 1, 2) ............................................................................................1-4
1.3 EAF 지식 격차 업데이트 및 개정의 목적 ....................................................................1-4
1.4 보고서 구조 ................................................................................................................1-5

2 대기 중 피로 내구성 데이터....................................................................................................2-1
2.1 서론............................................................................................................................2-1
2.2 피로 데이터 생성 및 평균 데이터 곡선 ........................................................................2-1
2.2.1 지식 격차 9(이전 분류 범주 C, 우선순위 낮음) ...................................................2-1
2.2.2 지식 격차 37(이전 분류 범주 C, 우선순위 낮음) .................................................2-4
2.3 이동 계수....................................................................................................................2-6
2.3.1 지식 격차 12(이전 분류 범주 C, 우선순위 보통) .................................................2-7
2.3.2 지식 격차 3(이전 분류 범주 C, 우선순위 높음) .................................................2-11
2.4 피로 및 환경 피로 데이터베이스 ...............................................................................2-13
2.4.1 지식 격차 46(이전 분류 범주 C, 우선순위 높음) ...............................................2-13
2.5 대기 중 피로 내구성 데이터 및 설계 곡선 요약 .........................................................2-14

3 LWR 환경 피로......................................................................................................................3-1
3.1 서론............................................................................................................................3-1
3.2 기존 주기의 하중 피로 테스트 ....................................................................................3-2

xi
13321451
A-41
3.2.1 지식 격차 19(이전 분류 범주 C, 우선순위 보통) .................................................3-2
3.2.2 지식 격차 10, 13, 47 ..........................................................................................3-4
3.2.3 지식 격차 20(이전 분류 범주 C, 우선순위 보통) ...............................................3-11
3.2.4 지식 격차 7(이전 분류 범주 C, 우선순위 높음) .................................................3-13
3.2.5 지식 격차 34(이전 분류 범주 A, 우선순위 높음) ...............................................3-15
3.2.6 지식 격차 21(이전 분류 범주 C, 우선순위 낮음) ...............................................3-15
3.3 복합 파형 테스트 ......................................................................................................3-17
3.3.1 지식 격차 17(이전 분류 범주 C, 우선순위 높음) ...............................................3-17
3.3.2 지식 격차 18(이전 분류 범주 B, 우선순위 높음) ...............................................3-18
3.3.3 지식 격차 11(이전 분류 범주 A, 우선순위 낮음) ...............................................3-23
3.3.4 지식 격차 14(이전 분류 범주 A, 우선순위 보통) ...............................................3-26
3.4 비등온 하중 ..............................................................................................................3-27
3.4.1 지식 격차 15, 16 ..............................................................................................3-27
3.5 프로토타입 및 컴포넌트 테스트 ................................................................................3-32
3.5.1 지식 격차 33(이전 분류 범주 A, 우선순위 높음) ...............................................3-32
3.5.2 지식 격차 36(이전 분류 범주 C, 우선순위 높음) ...............................................3-36
3.6 LWR 환경 피로 요약 .................................................................................................3-38

4 환경 피로 균열 성장 ...............................................................................................................4-1
4.1 서론............................................................................................................................4-1
4.2 기존 주기의 하중 피로 테스트 ....................................................................................4-2
4.2.1 지식 격차 22(이전 분류 범주 C, 우선순위 보통) .................................................4-2
4.2.2 지식 격차 24(이전 분류 범주 C, 우선순위 보통) .................................................4-4
4.2.3 지식 격차 25, 31 ................................................................................................4-5
4.2.4 지식 격차 29, 30 ..............................................................................................4-10
4.2.5 지식 격차 26(이전 분류 범주 C, 우선순위 보통) ...............................................4-11
4.2.6 지식 격차 27(이전 분류 범주 C, 우선순위 보통) ...............................................4-12
4.2.7 지식 격차 32(이전 분류 범주 C, 우선순위 낮음) ...............................................4-14
4.3 복합 파형 및 비등온 하중 .........................................................................................4-16
4.3.1 지식 격차 28(이전 분류 범주 A, 우선순위 높음) ...............................................4-16
4.4 균열 성장 데이터 요약 ..............................................................................................4-21

13321451
A-42
5 평가 방법...............................................................................................................................5-1
5.1 서론............................................................................................................................5-1
5.1.1 피로 및 EAF 에 대한 기계적 이해 요약 ...............................................................5-1
5.1.2 핵심 개념 검토 ...................................................................................................5-2
5.1.3 피로 발생 평가 ...................................................................................................5-3
5.1.4 피로 균열 성장 평가 ...........................................................................................5-4
5.1.5 현행 피로 평가 방법 요약 ...................................................................................5-4
5.2 피로 발생....................................................................................................................5-9
5.2.1 지식 격차 2(이전 분류에 해당되지 않음) ............................................................5-9
5.2.2 지식 격차 4(이전 분류 범주 B, 우선순위 높음) .................................................5-12
5.2.3 지식 격차 1(이전 분류 범주 C, 우선순위 높음) .................................................5-13
5.2.4 지식 격차 5(이전 분류 범주 C, 우선순위 높음) .................................................5-14
5.3 피로 발생: 복합 파형 분석 ........................................................................................5-15
5.3.1 지식 격차 8(이전 분류 범주 A, 우선순위 보통) .................................................5-15
5.3.2 지식 격차 39, 41, 42, 43 ..................................................................................5-17
5.3.3 지식 격차 6, 44 ................................................................................................5-21
5.4 피로 균열 성장 .........................................................................................................5-23
5.4.1 지식 격차 23(이전 분류 범주 C, 우선순위 높음) ...............................................5-23
5.4.2 지식 격차 28(이전 분류 범주 A, 우선순위 높음) ...............................................5-25
5.5 용접..........................................................................................................................5-28
5.5.1 지식 격차 38(이전 분류 범주 C, 우선순위 보통) ...............................................5-28
5.6 형상 특성..................................................................................................................5-30
5.6.1 지식 격차 40(이전 분류 범주 C, 우선순위 보통) ...............................................5-30
5.7 기타 지식 격차 .........................................................................................................5-31
5.7.1 지식 격차 35(이전 분류 범주 B, 우선순위 높음) ...............................................5-31
5.7.2 지식 격차 45(이전 분류 범주 C, 우선순위 보통) ...............................................5-32
5.8 요약..........................................................................................................................5-34

6 지식 격차 관찰 결과 및 업데이트 제안 요약 ............................................................................6-1
6.1 서론............................................................................................................................6-1
6.2 지식 격차 원안 현황 ...................................................................................................6-2
6.2.1 일반적 문제 .......................................................................................................6-2

13321451
A-43
6.2.2 대기 중 및 피로 설계 곡선상의 피로 내구성 데이터 ............................................6-2
6.2.3 LWR 환경이 피로 설계에 미치는 영향 ...............................................................6-5
6.2.4 복합 및 비등온 하중 과도상태가 LWR 환경에서 피로 수명에 미치는 영향 .........6-8
6.2.5 프로토타입 및 컴포넌트 테스트 .......................................................................6-10
6.2.6 피로 균열 성장 .................................................................................................6-10
6.2.7 EAF 평가 방법 .................................................................................................6-13

7 지식 격차의 합리적 설명 및 통합............................................................................................7-1


7.1 서론............................................................................................................................7-1
7.2 일반적 문제 ................................................................................................................7-2
7.2.1 EAF 모델을 사용한 예측과 발전소 절차 간의 차이 .............................................7-2
7.2.2 전 세계 관계자들의 데이터베이스 ......................................................................7-3
7.3 대기 중 및 피로 설계 곡선상의 피로 내구성 데이터 ....................................................7-3
7.3.1 재료 변동성 및 재료별 피로 설계 곡선 ...............................................................7-3
7.3.2 설계 피로 곡선에 대한 이동 계수 .......................................................................7-4
7.3.3 용접 금속 데이터 및 평가 ...................................................................................7-4
7.4 LWR 환경이 피로 설계에 미치는 영향 ........................................................................7-4
7.4.1 Fen 방정식, 매개변수, 임계값을 정의하는 데이터 결함 .......................................7-4
7.4.2 대기 중 피로 설계 이동 계수의 LWR 환경 적용 가능성 ......................................7-5
7.4.3 변형 진폭이 Fen 에 미치는 영향 ..........................................................................7-6
7.4.4 평균 응력이 피로 수명 및 EAF 에 미치는 영향 ...................................................7-7
7.5 복합 및 비등온 하중 과도상태가 LWR 환경에서 피로 수명에 미치는 영향 ..................7-7
7.5.1 복합 및 비등온 하중 조건에서의 EAF 내구성 데이터 .........................................7-7
7.5.2 Fen 방정식에서 주기적 가변 매개변수의 처리.....................................................7-8
7.5.3 열충격 과도상태에 대한 변형 구배의 영향 .........................................................7-9
7.6 LWR 환경에서의 피로균열 성장 .................................................................................7-9
7.6.1 오스테나이트 스테인리스강에 대한 준거 균열 성장 곡선 ...................................7-9
7.6.2 탄소강 및 저합금강에 대한 준거 균열 성장 곡선 ................................................7-9
7.6.3 니켈 합금에 대한 준거 균열 성장 곡선 .............................................................7-10
7.6.4 발전소 연관 하중 조건에서의 균열 성장 ...........................................................7-10
7.7 프로토타입 및 컴포넌트 테스트 ................................................................................7-11

13321451
A-44
8 결론.......................................................................................................................................8-1

9 참고자료................................................................................................................................9-1

13321451
A-45
그림 목록

그림 5-1 SNW 가중치 함수.....................................................................................................5-20

13321451
A-46
표 목록

표 3-1 페라이트강에 대한 Fen 매개변수 및 임계값 ....................................................................3-7


표 3-2 EFEM 스테인리스강 Fen 방정식에 대한 매개변수 ..........................................................3-8
표 3-3 오스테나이트강에 대한 Fen 매개변수 및 임계값 .............................................................3-9
표 3-4 니켈 합금에 대한 Fen 매개변수 및 임계값 ....................................................................3-10
표 6-1 다양한 재료, 하중, 환경 매개변수의 영향을 나타내는 평균 피로 ε–N 대기 곡선에
적용되는 수명 관련 계수.................................................................................................6-15
표 8-1 새로운 지식 격차 목록 ...................................................................................................8-3

13321451
A-47
13321451
Análisis de deficiencias en el
conocimiento de la fatiga asistida
por el medio ambiente (EAF)
Actualización y revisión de deficiencias en el
conocimiento de la EAF
3002013214

Informe final, septiembre de 2018

Jefe de proyecto de EPRI


K. Ahluwalia

Este producto está sujeto a la totalidad o a parte de los


requisitos del Programa de garantía de calidad nuclear
de EPRI.

ELECTRIC POWER RESEARCH INSTITUTE


3420 Hillview Avenue, Palo Alto, California 94304-1338 ▪ PO Box 10412, Palo Alto, California 94303-0813 ▪ USA
+1 800.313.3774 ▪ +1 650.855.2121 ▪ askepri@epri.com ▪ www.epri.com

13321451
A-49
EXENCIÓN DE RESPONSABILIDAD SOBRE GARANTÍAS Y LIMITACIÓN
DE RESPONSABILIDADES
ESTE DOCUMENTO HA SIDO ELABORADO POR LAS ORGANIZACIONES INDICADAS A
CONTINUACIÓN EN RELACIÓN CON EL TRABAJO PATROCINADO O COPATROCINADO POR
ELECTRIC POWER RESEARCH INSTITUTE, INC. (EPRI). NI EPRI, NI NINGÚN MIEMBRO DE EPRI,
O COPATROCINADOR, O LAS ORGANIZACIONES SIGUIENTES, NI NINGUNA OTRA PERSONA
QUE ACTÚE EN NOMBRE DE CUALQUIERA DE LAS PARTES ANTERIORES:
(A) OFRECERÁN GARANTÍA NI REPRESENTACIÓN ALGUNA, EXPLÍCITA O IMPLÍCITA, (I) EN
RELACIÓN CON EL USO DE CUALQUIER INFORMACIÓN, DISPOSITIVO, MÉTODO, PROCESO O
PRODUCTO SIMILAR DESCRITO EN ESTE DOCUMENTO, INCLUIDA LA COMERCIALIZACIÓN E
IDONEIDAD PARA CUALQUIER FINALIDAD PARTICULAR, NI EN LO RELATIVO A QUE (II) DICHO
USO NO INFRINJA O NO INTERFIERA CON LOS DERECHOS PRIVADOS, INCLUIDA LA
PROPIEDAD INTELECTUAL DE CUALQUIER PARTE, O (III) QUE ESTE DOCUMENTO SEA
ADECUADO CONSIDERANDO LAS CIRCUNSTANCIAS PARTICULARES DE CUALQUIER
USUARIO; O
(B) SE RESPONSABILIZARÁN DE CUALQUIER DAÑO O RESPONSABILIDAD DE TODO TIPO
(INCLUIDOS LOS DAÑOS DERIVADOS, INCLUSO SI EPRI O CUALQUIER REPRESENTANTE DE
ESTA HUBIERA SIDO NOTIFICADO CON RESPECTO A LA POSIBILIDAD DE QUE SE
PRODUJERAN DICHOS DAÑOS) DERIVADOS DE LA ELECCIÓN O USO DE ESTE DOCUMENTO O
DE CUALQUIER OTRA INFORMACIÓN, DISPOSITIVO, MÉTODO, PROCESO O PRODUCTO
SIMILAR DESCRITO EN ESTE DOCUMENTO.
LAS REFERENCIAS REALIZADAS EN ESTE DOCUMENTO A CUALQUIER PRODUCTO
COMERCIAL, PROCESO O SERVICIOS ESPECÍFICOS POR SU NOMBRE COMERCIAL, MARCA
COMERCIAL, FABRICANTE O MEDIANTE OTRA DENOMINACIÓN NO CONSTITUYE NI IMPLICA
NECESARIAMENTE SU APROBACIÓN, RECOMENDACIÓN NI PREFERENCIA POR PARTE DE
EPRI.
LA SIGUIENTE ORGANIZACIÓN, CONTRATADA POR EPRI, HA ELABORADO ESTE INFORME:
Amec Foster Wheeler

LOS CONTENIDOS TÉCNICOS DE ESTE PRODUCTO NO HAN SIDO ELABORADOS DE ACUERDO


CON EL MANUAL DEL PROGRAMA DE CALIDAD DE EPRI, QUE CUMPLE LOS REQUISITOS DE LA
NORMA 10 CFR 50, APÉNDICE B. ESTE PRODUCTO NO ESTÁ SUJETO A LOS REQUISITOS DE
LA NORMA 10 CFR, PARTE 21.

NOTA
Para más información sobre EPRI, llame al Centro de atención al cliente de EPRI al +1 800.313.3774 o
mande un correo electrónico a: askepri@epri.com.

Electric Power Research Institute, EPRI, y TOGETHERSHAPING THE FUTURE OF ELECTRICITY


son marcas registradas del Electric Power Research Institute, Inc.
Copyright © 2018 Electric Power Research Institute (EPRI), Inc. Todos los derechos reservados.

13321451
A-50
AGRADECIMIENTOS

La siguiente organización, contratada por EPRI, ha elaborado este informe:


Amec Foster Wheeler
Booths Park, Chelford Road
Knutsford, Cheshire
WA16 8QZ
Inglaterra
Investigador principal
D. Tice
Revisores
G. Quirk
N. Platts
A. Davison
Autores
A. McLennan
P. Gill

Este informe describe la investigación patrocinada por EPRI. J. Smith and G. Stevens, miembros
del personal de EPRI, han contribuido a la elaboración de este informe.

Esta publicación es un documento corporativo que se citará del siguiente modo en la literatura:
Análisis de deficiencias en el conocimiento de la fatiga asistida por el medio ambiente (EAF):
Actualización y revisión de deficiencias en el conocimiento de la EAF. EPRI, Palo Alto, CA:
2018. 3002013214.

13321451
A-51
RESUMEN

Los diseñadores que intenten realizar una ampliación de la vida útil y nueva construcción de una
central nuclear, deben demostrar un funcionamiento seguro de esta durante toda la vida útil
prevista. Esto suele incluir la aplicación de las curvas de diseño establecidas para el inicio de la
fatiga. No obstante, las normas y guías correspondientes existentes pueden predecir factores de
uso acumulativos muy altos para algunos componentes cuando se consideran los efectos de la
fatiga asistida por el medio ambiente (EAF). Asimismo, los elevados factores de uso calculados
previstos no se reflejan en la experiencia operativa, y no ha habido fallos por fatiga conocidos
hasta la fecha que puedan atribuirse a la fatiga térmica de ciclo bajo mejorada por la EAF.
Para conocer la aplicabilidad de las normas sobre fatiga y guías de la EAF disponibles
actualmente, EPRI ha identificado deficiencias en el conocimiento de la EAF por parte del
sector. Esto se ha documentado en el informe sobre análisis de deficiencias en el conocimiento,
publicado en 2011 (producto 1023012 de EPRI) y el informe de Hoja de ruta publicado en 2012
(producto 1026724 de EPRI). Dichos documentos destacaban las áreas técnicas en las que había
incertidumbre e identificó la investigación y desarrollo necesarios para abordarlas. Esto dio
como resultado las 47 deficiencias en el conocimiento, que fueron revisadas por parte de
expertos del sector nuclear, los cuales asignaron prioridades a las distintas deficiencias.
Desde la publicación del Informe sobre deficiencias en el conocimiento, ha habido una serie de
importantes actividades de investigación en todo el mundo y se han propuesto varias
metodologías alternativas para la evaluación de la EAF. Algunas de estas actividades abordaron
específicamente algunas de las deficiencias en el conocimiento identificadas previamente.
Este informe proporciona una revisión detallada de los resultados de la investigación disponibles
al público desde que se elaboraron las deficiencias en el conocimiento en 2011 y ofrece una
actualización de dichas deficiencias. También identifica que esos importantes datos de
laboratorio adicionales se han generado para entornos de reactores de agua ligera desde el
momento del informe anterior.
Debe recordarse que se ha identificado la necesidad de un ensayo representativo de la central
usando geometrías relevantes para la misma e historiales de carga. Como resultado, en la
actualidad hay un interés considerable en todo el mundo, apoyado por un grupo de colaboración
internacional de partes interesadas, en proceder a elaborar y realizar un ensayo de la EAF que sea
más representativo de los componentes de la central y experimente transitorios tipo central.
Dicho ensayo ofrecería una importante información adicional y un medio para validad cualquier
método de evaluación propuesto.

13321451
A-52
Las 47 deficiencias en el conocimiento se revisaron y se elaboró una lista más reducida de 17
cuestiones pendientes que justifican un estudio adicional. El fundamento de la lista de
deficiencias en el conocimiento era simplificar las deficiencias existentes eliminando las
redundancias y los solapamientos, así como centrar la atención en las deficiencias restantes sobre
los principales problemas identificados por el sector nuclear.
Palabras clave
Crecimiento de las fisuras
EAF
Fatiga asistida por el medio ambiente
Evaluación de la fatiga
Deficiencias en el conocimiento
Reactor de agua ligera

13321451
A-53
RESUMEN EJECUTIVO

Número del informe de seguimiento: 3002013214


Tipo de producto: Informe técnico
Denominación del producto: Análisis de deficiencias en el conocimiento de la fatiga
asistida por el medio ambiente (EAF): Actualización y revisión de deficiencias en el
conocimiento de la EAF

PÚBLICO PRINCIPAL: Diseñadores y operadores de centrales nucleares


PÚBLICO SECUNDARIO: Comunidad internacional de investigación nuclear

FUNDAMENTO DE LA INVESTIGACIÓN
La Evaluación original de deficiencias en el conocimiento (informe 1023012 de EPRI) se publicó en 2011 y
ofrecía una lista de deficiencias en el conocimiento por parte del sector nuclear de la fatiga asistida por el
medio ambiente (EAF). Desde la publicación de dicho documento, diversas instituciones de investigación de
todo el mundo han realizado un considerable trabajo adicional para lograr un mayor conocimiento de este
mecanismo de degradación, con el objetivo de encontrar métodos eficaces y menos conservadores con los
que evaluar la vida útil por fatiga de los componentes de las centrales nucleares. A la luz de la importante
cantidad de investigación realizada al respecto, el presente informe se ha elaborado para proporcionar una
actualización y mejora de las deficiencias en el conocimiento existentes (tal y como venían definidas
previamente en el informe 1023012 de EPRI).

RESUMEN DE LA INVESTIGACIÓN
El efecto perjudicial de los entornos de los reactores de agua ligera (LWR) sobre los materiales ensayados
en el laboratorio es bien conocido. No obstante, no está claro cómo deben aplicarse los resultados de
laboratorio, que usaban muestras bajo cargas completamente invertidas, a los componentes de centrales
nucleares reales. La EAF supone actualmente un problema importante para el diseño de las centrales
nucleares de nueva construcción y las que van a ampliar su vida útil, debido a los factores de uso con alta
fatiga previstos por las metodologías actuales que incorporan factores medioambientales (Fen). Debido a los
problemas que presenta la EAF, se ha realizado una importante cantidad de investigación por parte de la
comunidad internacional. En 2011/2012, se realizó un trabajo para evaluar el conocimiento de la EAF por
parte del sector y se presentó una lista de deficiencias en el conocimiento con las que orientar la investigación
internacional para que el sector obtuviera el máximo beneficio. El documento sobre deficiencias en el
conocimiento encontró 47 deficiencias relativas al inicio de la fatiga en el aire y en los entornos de los
reactores de agua ligera, crecimiento de las fisuras por fatiga en entornos de reactores de agua ligera y
metodologías para la evaluación de la EAF. Dada la gran cantidad de trabajo realizado sobre la EAF por el
sector nuclear desde 2011, se identificó la necesidad de actualizar y revisar las deficiencias en el
conocimiento de la EAF. Este informe detalla el trabajo realizado en relación con las deficiencias en el
conocimiento anteriores, destaca el solapamiento y redundancia entre las deficiencias y revisa las
deficiencias en base a la información recopilada.

CONCLUSIONES PRINCIPALES
Se realizó una revisión detallada de la investigación desde que se elaboraron las deficiencias en el
conocimiento originales en 2011 y se encontró que se habían obtenido importantes datos de laboratorio
adicionales en los entornos de los reactores de agua ligera. El objetivo principal de este trabajo reciente fue
el acero inoxidable austenítico en el entorno del refrigerante del primario del reactor de agua a presión. Gran
parte de estos ensayos se han realizado bajo condiciones orientadas a simular con mayor precisión las

13321451
A-54
condiciones de funcionamiento en términos de factores como los transitorios de cargas complejos, cargas
termomecánicas y acabado superficial. No obstante, los datos de los ensayos sobre geometrías
representativas de la central y tamaños de componentes representativos todavía son limitados. Las
observaciones de estos estudios experimentales avanzados proporcionan una explicación parcial de la
discrepancia observada entre las predicciones de modelos basados en Fen y la experiencia de campo sobre
componentes de la central. Se han elaborado modelos predictivos mejorados que ofrecen una mejor
descripción del comportamiento observado en estos recientes ensayos de laboratorio. No obstante, no hay
disponible una completa validación de las geometrías de los componentes representativos de la central y de
los historiales de carga.
En base a la revisión de estos datos, las 47 deficiencias en el conocimiento existentes se revisaron y se
elaboró una lista más reducida de cuestiones pendientes que garantizan un estudio adicional. Esta lista
revisada es la siguiente, con las correspondientes deficiencias en el conocimiento indicadas entre paréntesis:
A. Cuestiones generales
1. Estudios adicionales sobre la resistencia a la fatiga y el crecimiento de la fisuración por fatiga
para apoyar los modelos de evaluación de la EAF mejorados y abordar las discrepancias entre
las predicciones de los modelos EAF y los procedimientos de la central (deficiencias en el
conocimiento 2 y 4)
2. Base de datos universal de la EAF de partes interesadas con un adecuado análisis de datos
(deficiencia en el conocimiento 46)
B. Datos de resistencia a la fatiga en el aire y en las curvas de diseño por fatiga
1. Trabajo adicional para abordar los efectos de la variabilidad del material y, cuando proceda,
elaborar curvas de diseño por fatiga específicas del material (deficiencia en el conocimiento 9)
2. Factores de transferencia revisados sobre la vida útil y la tensión para las curvas de diseño por
fatiga (deficiencias en el conocimiento 3 y 12)
3. Datos, mejora en el conocimiento y estandarización de los métodos de evaluación de la fatiga
para soldaduras (deficiencias en el conocimiento 37 y 38)
C. Efectos del entorno de los reactores de agua ligera en el diseño por fatiga
1. Deficiencias en los datos relativos a las ecuaciones de Fen, en especial respecto a la
temperatura y umbrales de la velocidad de los esfuerzos (deficiencias en el conocimiento 1, 5,
10, 13, 20, 21, 34 y 47)
2. Métodos de diseño de la EAF mejorados que incorporan diferencias entre los factores de
transferencia entre entornos de reactores de agua ligera y el aire (deficiencia en el conocimiento
19)
3. Métodos de evaluación mejorados que reconocen los efectos de la amplitud de los esfuerzos en
Fen (deficiencia en el conocimiento 7)
4. Efectos de la tensión media sobre la vida útil por fatiga y EAF (deficiencias en el conocimiento 6,
11, 14 y 44)
D. Influencia de los transitorios de cargas complejas y no isotérmicas sobre la vida útil por fatiga
en entornos de reactores de agua ligera
1. Metodologías de evaluación mejoradas bajo cargas complejas y no isotérmicas más relevantes
para los componentes de la central (deficiencias en el conocimiento 8, 18, 39, 40, 41, 42 y 43)
2. Datos de validación bajo cargas relevantes para la central para apoyar nuevos métodos de
evaluación, incluida la aplicabilidad de las aleaciones distintas al acero inoxidable (deficiencias
en el conocimiento 15, 16 y 17)
3. Metodologías de vida útil total mejoradas para incorporar la influencia de los gradientes de
esfuerzos para transitorios de choque térmico (deficiencia en el conocimiento 16)
E. Crecimiento de las fisuras por fatiga en entornos de reactores de agua ligera
1. Mejoras en la referencia de las curvas de crecimiento de fisuras para aceros inoxidables
austeníticos, incorporando el umbral y los efectos del coeficiente R (incluida la R negativa) y la

13321451
A-55
extensión a las condiciones de los reactores de agua en ebullición (química del agua normal y
química del agua hidrogenada) (deficiencias en el conocimiento 22, 23, 24, 25 y 31)
2. Curvas de crecimiento de fisuras de referencia para aceros al carbono y de baja aleación en
entornos de reactores de agua en ebullición (deficiencias en el conocimiento 29 y 30)
3. Curvas de crecimiento de fisuras de referencia mejoradas para aleaciones de níquel
(deficiencias en el conocimiento 26, 27 y 32)
4. Métodos de evaluación del crecimiento de fisuras bajo cargas relevantes para la central que
reconocen la ponderación del daño a través del ciclo de carga (deficiencias en el conocimiento
28 y 45)
F. Ensayos en prototipos y componentes (deficiencias en el conocimiento 33, 35 y 36)

POR QUÉ ESTO ES IMPORTANTE


Este trabajo proporciona un resumen de la investigación realizada desde la publicación del documento
original sobre deficiencias en el conocimiento. Este informe tiene por finalidad ofrecer una declaración
independiente sobre los últimos avances en la investigación de la EAF. También destaca los vínculos entre
las diferentes técnicas de investigación y métodos de evaluación para facilitar una mejor colaboración y
avance en materia de EAF.

CÓMO USAR LOS RESULTADOS


El trabajo presentado en este documento puede usarse para ayudar a justificar y orientar los programas de
investigación que son importantes para cada miembro. Asimismo, este documento muestra los vínculos entre
la investigación experimental y el desarrollo de métodos de evaluación, que podrían permitir a los miembros
de EPRI identificar mejor oportunidades de colaboración para maximizar los efectos de su investigación.

OPORTUNIDADES DE APRENDIZAJE Y PARTICIPACIÓN


• Las organizaciones internacionales de investigación pueden encontrar útil el informe para orientar la
potencial dirección futura de la investigación y ensayos sobre fatiga asistida por el medio ambiente.

CONTACTOS DE EPRI: Kawaljit (Al) Ahluwalia, Ejecutivo Técnico, kahluwal@epri.com


Jean Smith, Responsable técnico principal, jmsmith@epri.com
Gary Stevens, Ejecutivo Técnico, gstevens@epri.com

PROGRAMAS: Primary Systems Corrosion Research (PSCR), Program 41.01.01; Boiling Water Reactor
Vessel and Internals Project (BWRVIP), Program 41.01.03; Pressurized Water Reactor Materials Reliability
Program (MRP), Program 41.01.04

CATEGORÍA DE APLICACIÓN: Referencia–Base técnica

Together...Shaping the Future of Electricity®

Electric Power Research Institute


3420 Hillview Avenue, Palo Alto, California 94304-1338 • PO Box 10412, Palo Alto, California 94303-0813 USA
+1 800.313.3774 • +1 650.855.2121 • askepri@epri.com • www.epri.com
© 2018 Electric Power Research Institute (EPRI), Inc. Todos los derechos reservados. Electric Power Research Institute, EPRI, y
TOGETHER SHAPING THE FUTURE OF ELECTRICITY son marcas registradas de Electric Power Research Institute, Inc.

13321451
A-56
CONTENIDO

RESUMEN ................................................................................................................................. V

RESUMEN EJECUTIVO .......................................................................................................... VII

1 INTRODUCCIÓN ..................................................................................................................1-1
1.1 Antecedentes .............................................................................................................1-1
1.2 Hipótesis de la hoja de ruta ........................................................................................1-2
1.2.1 Hipótesis............................................................................................................1-2
1.2.2 Recomendaciones, opción 1 y 2 ........................................................................1-4
1.3 Objetivos de la actualización y revisión de las deficiencias en el conocimiento
de la EAF ............................................................................................................................1-4
1.4 Estructura del informe ................................................................................................1-5

2 DATOS DE RESISTENCIA A LA FATIGA EN EL AIRE.......................................................2-1


2.1 Introducción ...............................................................................................................2-1
2.2 Generación de datos de fatiga y curvas medias de datos ..........................................2-1
2.2.1 Deficiencia en el conocimiento 9 (anteriormente clasificada como
categoría C, de baja prioridad)........................................................................................2-1
2.2.2 Deficiencia en el conocimiento 37 (anteriormente clasificada como
categoría C, de baja prioridad)........................................................................................2-4
2.3 Factores de transferencia ...........................................................................................2-6
2.3.1 Deficiencia en el conocimiento 12 (anteriormente clasificada como
categoría C, de prioridad media) .....................................................................................2-7
2.3.2 Deficiencia en el conocimiento 3 (anteriormente clasificada como
categoría C, de alta prioridad).......................................................................................2-11
2.4 Bases de datos sobre fatiga y fatiga medioambiental ...............................................2-13
2.4.1 Deficiencia en el conocimiento 46 (anteriormente clasificada como
categoría C, de alta prioridad).......................................................................................2-13
2.5 Resumen de datos de resistencia a la fatiga y curvas de diseño en el aire ..............2-14

3 FATIGA MEDIOAMBIENTAL EN REACTORES DE AGUA LIGERA...................................3-1

13321451
A-57
3.1 Introducción ...............................................................................................................3-1
3.2 Ensayos de fatiga por cargas cíclicas convencionales ...............................................3-2
3.2.1 Deficiencia en el conocimiento 19 (anteriormente clasificada como
categoría C, de prioridad media) .....................................................................................3-2
3.2.2 Deficiencias en el conocimiento 10, 13 y 47 ......................................................3-4
3.2.3 Deficiencia en el conocimiento 20 (anteriormente clasificada como
categoría C, de prioridad media) ...................................................................................3-11
3.2.4 Deficiencia en el conocimiento 7 (anteriormente clasificada como
categoría C, de alta prioridad).......................................................................................3-13
3.2.5 Deficiencia en el conocimiento 34 (anteriormente clasificada como
categoría A, de alta prioridad) .......................................................................................3-15
3.2.6 Deficiencia en el conocimiento 21 (anteriormente clasificada como
categoría C, de baja prioridad)......................................................................................3-15
3.3 Ensayo de forma de onda compleja .........................................................................3-17
3.3.1 Deficiencia en el conocimiento 17 (anteriormente clasificada como
categoría C, de alta prioridad).......................................................................................3-17
3.3.2 Deficiencia en el conocimiento 18 (anteriormente clasificada como
categoría B, de alta prioridad) .......................................................................................3-18
3.3.3 Deficiencia en el conocimiento 11 (anteriormente clasificada como
categoría A, de baja prioridad) ......................................................................................3-23
3.3.4 Deficiencia en el conocimiento 14 (anteriormente clasificada como
categoría A, de prioridad media) ...................................................................................3-26
3.4 Carga no isotérmica .................................................................................................3-27
3.4.1 Deficiencias en el conocimiento 15 y 16 ..........................................................3-27
3.5 Ensayos en prototipos y componentes .....................................................................3-32
3.5.1 Deficiencia en el conocimiento 33 (anteriormente clasificada como
categoría A, de alta prioridad) .......................................................................................3-32
3.5.2 Deficiencia en el conocimiento 36 (anteriormente clasificada como
categoría C, de alta prioridad).......................................................................................3-36
3.6 Resumen de la fatiga medioambiental en reactores de agua ligera..........................3-38

4 CRECIMIENTO DE LAS FISURAS POR FATIGA MEDIOAMBIENTAL ..............................4-1


4.1 Introducción ...............................................................................................................4-1
4.2 Ensayos de fatiga por cargas cíclicas convencionales ...............................................4-2
4.2.1 Deficiencia en el conocimiento 22 (anteriormente clasificada como
categoría C, de prioridad media) .....................................................................................4-2
4.2.2 Deficiencia en el conocimiento 24 (anteriormente clasificada como
categoría C, de prioridad media).....................................................................................4-4
4.2.3 Deficiencias en el conocimiento 25 y 31 ............................................................4-5

13321451
A-58
4.2.4 Deficiencias en el conocimiento 29 y 30 ..........................................................4-10
4.2.5 Deficiencia en el conocimiento 26 (anteriormente clasificada como
categoría C, de prioridad media) ...................................................................................4-11
4.2.6 Deficiencia en el conocimiento 27 (anteriormente clasificada como
categoría C, de prioridad media) ...................................................................................4-12
4.2.7 Deficiencia en el conocimiento 32 (anteriormente clasificada como
categoría C, de baja prioridad)......................................................................................4-14
4.3 Formas de ondas complejas y carga no isotérmica ..................................................4-16
4.3.1 Deficiencia en el conocimiento 28 (anteriormente clasificada como
categoría A, de alta prioridad) .......................................................................................4-16
4.4 Resumen de datos sobre el crecimiento de fisuras ..................................................4-21

5 MÉTODOS DE EVALUACIÓN .............................................................................................5-1


5.1 Introducción ...............................................................................................................5-1
5.1.1 Resumen del conocimiento mecanicista de la fatiga y la EAF ...........................5-1
5.1.2 Revisión de conceptos clave .............................................................................5-2
5.1.3 Evaluación del inicio de la fatiga ........................................................................5-3
5.1.4 Evaluación del crecimiento de las fisuras por fatiga ...........................................5-4
5.1.5 Resumen de los métodos actuales de evaluación de la fatiga ...........................5-4
5.2 Inicio de la fatiga ........................................................................................................5-9
5.2.1 Deficiencia en el conocimiento 2 (no clasificada anteriormente) ........................5-9
5.2.2 Deficiencia en el conocimiento 4 (anteriormente clasificada como
categoría B, de alta prioridad) .......................................................................................5-12
5.2.3 Deficiencia en el conocimiento 1 (anteriormente clasificada como
categoría C, de alta prioridad).......................................................................................5-13
5.2.4 Deficiencia en el conocimiento 5 (anteriormente clasificada como
categoría C, de alta prioridad).......................................................................................5-14
5.3 Inicio de la fatiga: Análisis forma de onda compleja .................................................5-15
5.3.1 Deficiencia en el conocimiento 8 (anteriormente clasificada como
categoría A, de prioridad media) ...................................................................................5-15
5.3.2 Deficiencias en el conocimiento 39, 41, 42 y 43 ..............................................5-17
5.3.3 Deficiencias en el conocimiento 6 y 44 ............................................................5-21
5.4 Crecimiento de las fisuras por fatiga ........................................................................5-23
5.4.1 Deficiencia en el conocimiento 23 (anteriormente clasificada como
categoría C, de alta prioridad).......................................................................................5-23
5.4.2 Deficiencia en el conocimiento 28 (anteriormente clasificada como
categoría A, de alta prioridad) .......................................................................................5-25
5.5 Soldaduras ...............................................................................................................5-28

13321451
A-59
5.5.1 Deficiencia en el conocimiento 38 (anteriormente clasificada como
categoría C, de prioridad media) ...................................................................................5-28
5.6 Características geométricas .....................................................................................5-30
5.6.1 Deficiencia en el conocimiento 40 (anteriormente clasificada como
categoría C, de prioridad media) ...................................................................................5-30
5.7 Otras deficiencias en el conocimiento ......................................................................5-31
5.7.1 Deficiencia en el conocimiento 35 (anteriormente clasificada como
categoría B, de alta prioridad) .......................................................................................5-31
5.7.2 Deficiencia en el conocimiento 45 (anteriormente clasificada como
categoría C, de prioridad media) ...................................................................................5-32
5.8 Resumen..................................................................................................................5-34

6 RESUMEN DE LAS OBSERVACIONES Y ACTUALIZACIÓN PROPUESTA DE LAS


DEFICIENCIAS EN EL CONOCIMIENTO ...............................................................................6-1
6.1 Introducción ...............................................................................................................6-1
6.2 Estado de las deficiencias en el conocimiento originales ...........................................6-2
6.2.1 Cuestiones generales ........................................................................................6-2
6.2.2 Datos de resistencia a la fatiga en el aire y curvas de diseño por fatiga ............6-2
6.2.3 Efectos del entorno de los reactores de agua ligera en el diseño por fatiga.......6-5
6.2.4 Influencia de los transitorios de cargas complejas y no isotérmicas sobre
la vida útil por fatiga en entornos de reactores de agua ligera ........................................6-8
6.2.5 Ensayos en prototipos y componentes ............................................................6-10
6.2.6 Crecimiento de las fisuras por fatiga ................................................................6-10
6.2.7 Métodos de evaluación de la EAF ...................................................................6-13

7 JUSTIFICACIÓN Y CONSOLIDACIÓN DE DEFICIENCIAS EN EL CONOCIMIENTO ........7-1


7.1 Introducción ...............................................................................................................7-1
7.2 Cuestiones generales ................................................................................................7-2
7.2.1 Discrepancias entre las predicciones de los modelos EAF y los
procedimientos de la central ...........................................................................................7-2
7.2.2 Base de datos universal de las partes interesadas ............................................7-3
7.3 Datos de resistencia a la fatiga en el aire y curvas de diseño por fatiga .....................7-3
7.3.1 Variabilidad del material y curvas de diseño por fatiga específicas del
material ..........................................................................................................................7-3
7.3.2 Factores de transferencia para curvas de diseño por fatiga ...............................7-4
7.3.3 Datos y evaluación del metal de soldadura........................................................7-4
7.4 Efectos del entorno de los reactores de agua ligera en el diseño por fatiga ...............7-4

13321451
A-60
7.4.1 Deficiencias en los datos que definen las ecuaciones de Fen, parámetros y
umbrales .........................................................................................................................7-4
7.4.2 Aplicabilidad de los factores de transferencia en el diseño por fatiga en el
aire a los entornos de reactores de agua ligera ..............................................................7-5
7.4.3 Efectos de la amplitud de los esfuerzos en Fen ..................................................7-6
7.4.4 Efectos de la tensión media sobre la vida útil por fatiga y EAF ..........................7-7
7.5 Influencia de los transitorios de cargas complejas y no isotérmicas sobre la vida
útil por fatiga en entornos de reactores de agua ligera ........................................................7-7
7.5.1 Datos de resistencia de EAF bajo cargas complejas y no isotérmicas ...............7-7
7.5.2 Tratamiento de los parámetros variables cíclicamente en la ecuación Fen .........7-8
7.5.3 Influencia de los gradientes de esfuerzos en transitorios de choque
térmico ..........................................................................................................................7-9
7.6 Crecimiento de las fisuras por fatiga en entornos de reactores de agua ligera ...........7-9
7.6.1 Curvas de crecimiento de fisuras de referencia para aceros inoxidables
austeníticos ....................................................................................................................7-9
7.6.2 Curvas de crecimiento de fisuras de referencia para aceros al carbono y
de baja aleación..............................................................................................................7-9
7.6.3 Curvas de crecimiento de fisuras de referencia para aleaciones de níquel ......7-10
7.6.4 Crecimiento de fisuras bajo cargas relevantes para la central .........................7-10
7.7 Ensayos en prototipos y componentes .....................................................................7-11

8 CONCLUSIONES .................................................................................................................8-1

9 REFERENCIAS ....................................................................................................................9-1

13321451
A-61
LISTADO DE FIGURAS

Figura 5-1 Función de ponderación SNW ..............................................................................5-20

13321451
A-62
LISTADO DE TABLAS

Tabla 3-1 Parámetros de Fen y sus valores umbral para aceros ferríticos ................................3-7
Tabla 3-2 Parámetros de la ecuación Fen de acero inoxidable EFEM .......................................3-8
Tabla 3-3 Parámetros de Fen y sus valores umbral para aceros austeníticos. ..........................3-9
Tabla 3-4 Parámetros de Fen y sus valores umbral para aleaciones de níquel. ......................3-10
Tabla 6-1 Factores sobre la vida útil aplicados a la curva de aire de fatiga media ε–N
para tener en cuenta los efectos de varios materiales, cargas y parámetros
medioambientales. .........................................................................................................6-15
Tabla 8-1 Listado de nuevas deficiencias en el conocimiento ..................................................8-3

13321451
A-63
13321451
13321451
Export Control Restrictions The Electric Power Research Institute, Inc. (EPRI, www.epri.com)
Access to and use of EPRI Intellectual Property is granted with the spe- conducts research and development relating to the generation, delivery
cific understanding and requirement that responsibility for ensuring full and use of electricity for the benefit of the public. An independent,
compliance with all applicable U.S. and foreign export laws and regu- nonprofit organization, EPRI brings together its scientists and engineers
lations is being undertaken by you and your company. This includes as well as experts from academia and industry to help address
an obligation to ensure that any individual receiving access hereunder challenges in electricity, including reliability, efficiency, affordability,
who is not a U.S. citizen or permanent U.S. resident is permitted access health, safety and the environment. EPRI members represent 90% of the
under applicable U.S. and foreign export laws and regulations. In the electric utility revenue in the United States with international participation
event you are uncertain whether you or your company may lawfully in 35 countries. EPRI’s principal offices and laboratories are located in
obtain access to this EPRI Intellectual Property, you acknowledge that it Palo Alto, Calif.; Charlotte, N.C.; Knoxville, Tenn.; and Lenox, Mass.
is your obligation to consult with your company’s legal counsel to deter-
Together...Shaping the Future of Electricity
mine whether this access is lawful. Although EPRI may make available
on a case-by-case basis an informal assessment of the applicable U.S.
export classification for specific EPRI Intellectual Property, you and your
company acknowledge that this assessment is solely for informational
purposes and not for reliance purposes. You and your company ac-
knowledge that it is still the obligation of you and your company to make
your own assessment of the applicable U.S. export classification and
ensure compliance accordingly. You and your company understand and
acknowledge your obligations to make a prompt report to EPRI and the
appropriate authorities regarding any access to or use of EPRI Intellec-
tual Property hereunder that may be in violation of applicable U.S. or
foreign export laws or regulations.

Programs:
Boiling Water Reactor Vessel and Internals Project
Pressurized Water Reactor Materials Reliability Program
Primary Systems Corrosion Research (PSCR)

© 2018 Electric Power Research Institute (EPRI), Inc. All rights reserved. Electric Power
Research Institute, EPRI, and TOGETHER...SHAPING THE FUTURE OF ELECTRICITY are
registered service marks of the Electric Power Research Institute, Inc.

3002013214

Electric Power Research Institute


3420 Hillview Avenue, Palo Alto, California 94304-1338 • PO Box 10412, Palo Alto, California 94303-0813 USA
800.313.3774 • 650.855.2121 • askepri@epri.com • www.epri.com
13321451

You might also like