You are on page 1of 114

Technical Basis for ASME Code Case N-830,

Revision 1 (MRP-418)
Direct Use of Master Curve Fracture Toughness Curve for Pressure-Retaining
Materials of Class 1 Vessels, Section XI

2017 TECHNICAL REPORT


Technical Basis for ASME Code
Case N-830, Revision 1 (MRP-418)
Direct Use of Master Curve Fracture Toughness
Curve for Pressure-Retaining Materials of Class 1
Vessels, Section XI
3002010332

Final Report, October 2017

EPRI Project Manager


T. Hardin

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
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 ORGANIZATIONS PREPARED THIS REPORT:
Phoenix Engineering Associates, Inc.#
U.S. Nuclear Regulatory Commission##
Structural Integrity Associates, Inc.#

# Under contract to EPRI.


## This work was authored in part by a U.S. Government employee in the scope of his/her

employment. EPRI disclaims all interest in the U.S. Government’s contribution.

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 © 2017 Electric Power Research Institute, Inc. All rights reserved.
ACKNOWLEDGMENTS

The following organizations prepared this report:


Phoenix Engineering Associates, Inc.
119 Glidden Hill Road
Unity, NH 03743
Principal Investigator
M. Erickson
U.S. Nuclear Regulatory Commission, Office of Nuclear Regulatory Research (RES)
Washington, DC 20555-0001
Principal Investigator
M. Kirk*
Structural Integrity Associates, Inc.
11515 Vanstory Drive, Suite 125
Huntersville, NC 28078
Principal Investigator
G. Stevens

*
The statements, findings, conclusions and recommendations expressed in this report are those of
the authors and do not necessarily reflect the views of the U.S. Nuclear Regulatory Commission.
This report describes research co-sponsored by EPRI.

This publication is a corporate document that should be cited in the literature in the following
manner:
Technical Basis for ASME Code Case N-830, Revision 1 (MRP-418): Direct Use of Master
Curve Fracture Toughness Curve for Pressure-Retaining Materials of Class 1 Vessels, Section
XI. EPRI, Palo Alto, CA: 2017. 3002010332.
iii
This document was prepared as part of the ASME Code, Section XI Working Group on Flaw
Evaluation. The authors would like to thank the volunteer members of the ASME Code, Section
XI Working Group on Flaw Evaluation for their valuable input, feedback, and review of this
report, as well as their help and participation in solving the sample problems associated with this
effort.
Working Group on Flaw Evaluation Members Who Contributed to This Report:
• Russell Cipolla Intertek
• Yil Kim GE POWER
• Mark Kirk & Mike Benson U.S. NRC
• Darrell Lee BWXT
• Cheng Liu & Steven Xu Kinectrics
• Do Jun Shim Structural Integrity Associates, Inc.

iv
ABSTRACT

Section XI of the ASME Boiler and Pressure Vessel Code (ASME Code) provides KIc and KIa
fracture toughness models for ferritic steels. These models are based on linear elastic fracture
mechanics (LEFM) methods, and were initially developed in the 1970s for incorporation into the
ASME Code. The models have remained largely unchanged since their original incorporation
into the code. Since the publication of the technical bases documents for the fracture toughness
equations contained in Section XI, considerable advancements to the state of theoretical and
practical knowledge have occurred, particularly with respect to the amount of available fracture
toughness data. The Electric Power Research Institute (EPRI) now has a fracture toughness
database containing well over 9,000 fracture toughness values ranging across specimen sizes, test
temperatures and strain rates. As part of the pressurized thermal shock (PTS) re-evaluation
program, the U.S. Nuclear Regulatory Commission (NRC) and the industry used this database to
develop an integrated model that predicts the mean trends and scatter of the fracture toughness
behavior of ferritic steels throughout the temperature range from the lower shelf to the upper
shelf fracture regions. This integrated model includes the transition fracture toughness Master
Curve approach that describes the temperature dependence and scatter in KJc in the lower
transition temperature region, a new model for describing the temperature dependence and
scatter of JIc on the upper shelf, and includes identification of a temperature at which the KJc
curve transitions to upper shelf behavior, marking the upper limit of applicability for the KJc
transition curve. This collection of models was used by the NRC to establish the index
temperature screening limits adopted in the Alternate PTS Rule documented in Title 10 to the
U.S. Code of Federal Regulations (CFR), Part 50.61a (10CFR50.61a).
The ASME Section XI Working Group on Flaw Evaluation (WGFE) has an ongoing effort
intended to implement the KJc Master Curve (MC) into Section XI of the ASME Code. This
effort began with indirect implementation of the MC through use of a transition reference
temperature, RTT0, defined by using the KJc T0 value to replace RTNDT for indexing the ASME KIc
curve. In Revision 0 of Code Case N-830, direct use of the MC was defined as an alternative to
using the ASME KIc curve. Revision 1 to Code Case N-830 (N-830-1) incorporates the complete
and self-consistent suite of fracture toughness models developed over the last decade to
completely describe the temperature dependence, scatter, and interdependencies between all the
fracture toughness metrics (i.e., KJc, KIa, JIc, J0.1, and J-R) from the lower shelf through the upper
shelf regimes. This report describes the technical basis for Code Case N-830-1.

v
Keywords
Master curve
Fracture toughness model
T0 fracture toughness reference temperature
RTNDT fracture toughness reference temperature
ASME Section XI Appendix A flaw evaluation procedures

vi
EXECUTIVE SUMMARY

Deliverable Number: 3002010332


Product Type: Technical Report
Product Title: Technical Basis for ASME Code Case N-830, Revision 1 (MRP-418):
Direct Use of Master Curve Fracture Toughness Curve for Pressure-Retaining Materials
of Class 1 Vessels, Section XI

PRIMARY AUDIENCE: ASME Boiler & Pressure Vessel Code, Section XI, Committees
SECONDARY AUDIENCE: Engineers using Master Curve fracture toughness for vessel integrity evaluations

KEY RESEARCH QUESTION


Currently, the ASME Boiler & Pressure Vessel Code, Section XI methods for evaluation of vessel integrity are
based on methodology developed in the 1970s for conservatively representing fracture toughness without
actually measuring fracture toughness. It is desirable for the Code to provide a modern suite of best estimate
fracture toughness models that provide a complete description of fracture toughness crack initiation and arrest
behavior from lower shelf, through transition, to ductile upper shelf regimes for all ferritic steels.

RESEARCH OVERVIEW
The ASME Section XI Working Group on Flaw Evaluation (WGFE) has an ongoing effort intended to
implement the KJc Master Curve (MC) into Section XI of the ASME Code. This effort began with indirect
implementation of the MC through use of a transition reference temperature, RTT0, defined by using the KJc
T0 value to replace RTNDT for indexing the ASME KIc curve. In Revision 0 of Code Case (CC) N-830, direct
use of the MC was defined as an alternative to using the ASME KIc curve. The proposed Revision 1 to CC N-
830 (N-830-1) incorporates the complete and self-consistent suite of fracture toughness models developed
over the last decade to completely describe the temperature dependence, scatter, and interdependencies
between all the fracture toughness metrics (i.e., KJc, KIa, JIc, J0.1, and J-R) from the lower shelf through the
upper shelf regimes. This report describes the technical basis for Code Case N-830-1. This document was
prepared by a small task group to provide information to the Working Group on Flaw Evaluation to support
finalization and decision-making on CC N-830-1.

KEY FINDINGS
• The technical bases for the fracture toughness models contained in ASME CC N-830-1 are presented
in this report. The suite of best estimate fracture toughness models provides a complete description
of fracture toughness crack initiation and arrest behavior from lower shelf, through transition, to ductile
upper shelf regimes for all ferritic steels.
• The best estimate models used for CC N-830-1 are based on updated techniques and available data,
sound physical bases, and extensive empirical evaluations that collectively promote confidence in their
use for flaw assessment following Nonmandatory Appendix A of ASME Section XI and similar
methods.
• These models are appropriate for use in both deterministic and probabilistic assessments, as each
model describes the full distribution in values about the mean for any temperature and material
condition.

vii
EXECUTIVE SUMMARY

• Equations that allow an analyst to determine any percentile value of interest for any of the fracture
toughness parameters KJc, KIa, JIc, J0.1, and J-R are presented for each fracture toughness model.
Specific values of these parameters may be used in deterministic assessments, or the entire
distributions may be sampled for use in probabilistic assessments.

WHY THIS MATTERS


The fracture toughness models presented in this report provide a consistent, best-estimate representation of
ferritic steel fracture toughness behavior, including uncertainties, to allow quantitative fracture toughness
assessments that ensure the safety of nuclear (and other) ferritic components.

HOW TO APPLY RESULTS


Equations are presented for each fracture toughness model that allow an analyst to determine any percentile
value of interest for any of the fracture toughness parameters KJc, KIa, JIc, J0.1, and J-R. Specific values of
these parameters may be used in deterministic assessments, or the entire distributions may be sampled for
use in probabilistic assessments.

LEARNING AND ENGAGEMENT OPPORTUNITIES


Regulatory authorities considering the approval of use of Master Curve technologies for integrity evaluations
may also be interested in this report.

EPRI CONTACTS: Timothy C. Hardin, Technical Executive, thardin@epri.com

PROGRAM: Materials Reliability Program, 41.01.04

IMPLEMENTATION CATEGORY: Reference; Early R&D

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
© 2017 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.
NOMENCLATURE

Category Symbol Unit Description

KIc MPa√m plane strain fracture initiation toughness

K-equivalent of the value of J measured at cleavage


KJc MPa√m
crack initiation
KIa MPa√m fracture toughness measured at cleavage crack arrest
Different Fracture
ductile crack initiation toughness measured according
Toughness Metrics JIc kJ/m2
to ASTM E1820
ductile crack initiation toughness after “x” mm of ductile
Jx kJ/m2
crack extension

variation of ductile fracture toughness with stable crack


J-R kJ/m2
extension
temperature at which the KJc Master Curve 1 has a
T0 °C
median value of 100 MPa√m
Index temperature at which the KIa master curve has a
TKIa °C
Temperatures median value of 100 MPa√m
temperature at which the median KJc Master Curve
TUS °C
crosses the mean JIc upper shelf master curve

Temperature T °C temperature at the crack-tip


p dimensionless percentile for the lower bounding curves
Parameters to
Define Statistical T-statistic multiplier for the lower bounding curves.
Mp dimensionless Standard normal distribution with a mean of zero and
Bounding Curve
standard deviation of 1.0
KJcp MPa√m value of KJc at percentile p
Used in the KIc or KJc
Kmin MPa√m 20 MPa√m
equations
Ko MPa√m value of KJc at the 63.2nd percentile

Used in the KIa


KIap MPa√m value of KIa at percentile p
equation

1
Master Curve is only capitalized when referring to the KJc Master Curve. In all other cases, “master” is used as an
adjective to describe the type of curve-fit.

ix
Category Symbol Unit Description

JIcp kJ/m2 value of JIc at percentile p


JXp kJ/m2 value of JX at percentile p

leading coefficient of an exponential fit to the J-R


C kJ/m2
curve. value of C in the equation J = C(Δa)n

slope of an exponential fit to the J-R curve. The value


n dimensionless
of n in the equation J = C(Δa)n
σΔJIc kJ/m2 standard deviation of JIc
JIcmean kJ/m2 mean value of JIc

Used in the J-R and Jc(US) kJ/m2 value of Jc at TUS


JX equations
ΔJIc(US) kJ/m2 value of ΔJIc at TUS

E GPa Young’s modulus


A kJ/m2 fitting parameter in σΔJIc equation

B 1/°C fitting parameter in σΔJIc equation

P dimensionless percentile in σΔJIc equation

P1 dimensionless percentile in σΔJIc equation

P2 dimensionless percentile in σΔJIc equation

ν dimensionless Poisson’s ratio, a value of 0.3 can be used

x
CONTENTS

ABSTRACT ................................................................................................................................v

EXECUTIVE SUMMARY ..........................................................................................................vii

NOMENCLATURE ....................................................................................................................ix

1 INTRODUCTION ..................................................................................................................1-1
1.1 ASME Section XI, Appendix A Approach .......................................................................1-2
1.1.1 Summary of ASME Section XI, Appendix A Flaw Evaluation Procedures ..........1-2
1.1.2 Treatment of Uncertainties in Appendix A ..........................................................1-4
1.1.3 Technical Basis for the Appendix A Methodology ..............................................1-5
1.1.4 Issues with Appendix A Methodology ................................................................1-8
1.2 Objectives of Proposed Code Case N-830-1 ...............................................................1-10

2 OVERVIEW OF CODE CASE N-830-1 .................................................................................2-1


2.1 Introduction....................................................................................................................2-1
2.2 CC N-830-1 Contents ....................................................................................................2-2
2.2.1 Inquiry ...............................................................................................................2-2
2.2.2 Reply .................................................................................................................2-2
2.2.3 Discussion of N-830-1 .......................................................................................2-2

3 FRACTURE TOUGHNESS MODELS IN CC N-830-1 ..........................................................3-1


3.1 Cleavage Crack Initiation Toughness, KJc ......................................................................3-1
3.1.1 Description of the KJc Model ..............................................................................3-2
3.1.2 Basic Form ........................................................................................................3-2
3.1.3 Distribution ........................................................................................................3-2
3.1.4 Theoretical Basis ...............................................................................................3-3
3.1.5 Empirical Basis ..................................................................................................3-4
3.1.6 Model Validation ................................................................................................3-5
3.1.7 Limits of Applicability .........................................................................................3-6

xi
3.2 Cleavage Crack Arrest Fracture Toughness, KIa ............................................................3-7
3.2.1 Description of Model ..........................................................................................3-7
3.2.2 Basic Form ........................................................................................................3-7
3.2.3 Distribution ........................................................................................................3-7
3.2.4 Theoretical Basis ...............................................................................................3-8
3.2.5 Model Validation ................................................................................................3-9
3.2.6 Limits of Applicability .........................................................................................3-9
3.3 Ductile Crack Initiation Fracture Toughness, JIc .............................................................3-9
3.3.1 Description of Model ........................................................................................3-10
3.3.2 Basic Form ......................................................................................................3-10
3.3.3 Distribution ......................................................................................................3-10
3.3.4 Theoretical Basis .............................................................................................3-11
3.3.5 Empirical Basis ................................................................................................3-12
3.3.6 Model Validation ..............................................................................................3-13
3.3.7 Limits of Applicability .......................................................................................3-15

4 FRACTURE TOUGHNESS LINKAGE MODELS IN CC N-830-1..........................................4-1


4.1 The Relationship Between Cleavage Crack Initiation (KJc) and Upper Shelf (JIc);
TUS.......................................................................................................................................4-1
4.1.1 Mathematical Form of the Model .......................................................................4-1
4.1.2 Physical Basis ...................................................................................................4-2
4.1.3 Empirical Basis ..................................................................................................4-3
4.1.4 Model Validation ................................................................................................4-3
4.1.5 Limits on Validity of the Model ...........................................................................4-4
4.2 The Relationship Between Cleavage Crack Initiation (KJc) and Arrest (KIa) ....................4-4
4.2.1 Mathematical Form of Model .............................................................................4-4
4.2.2 Physical Basis for the Model ..............................................................................4-5
4.2.3 Empirical Basis ..................................................................................................4-6
4.2.4 Model Validation ................................................................................................4-7
4.2.5 Limits on Validity of the Model ...........................................................................4-7
4.3 The Relationship Between Upper Shelf (JIc) Crack Initiation and Upper Shelf
Crack Growth (J-R) .............................................................................................................4-7
4.3.1 Mathematical Form of the Model .......................................................................4-8
4.3.2 Empirical Basis ..................................................................................................4-8
4.3.3 Model Validation ..............................................................................................4-10
4.3.4 Limits on Validity of Model ...............................................................................4-10

xii
5 IMPLICATIONS OF PROPOSED CHANGES .......................................................................5-1
5.1 Introduction....................................................................................................................5-1
5.2 Sources of Uncertainties in Fracture Mechanics Analyses.............................................5-1
5.2.1 Flaw Size Uncertainty ........................................................................................5-2
5.2.2 Stress and Stress Intensity Factor Uncertainty ..................................................5-2
5.2.3 Fracture Toughness Uncertainty........................................................................5-3
5.3 Treatment of Uncertainties ............................................................................................5-3
5.3.1 Treatment of Uncertainty Due to Flaw Size and Location ..................................5-3
5.3.2 Treatment of Uncertainty Due to Stress .............................................................5-6
5.3.3 Treatment of Uncertainty on Fracture Toughness ..............................................5-8
5.4 CC N-830 ....................................................................................................................5-11
5.5 Code Case N-830-1 Uncertainty Treatment .................................................................5-12
5.6 Summary .....................................................................................................................5-14

6 POTENTIAL CODE/REGULATORY APPLICATIONS OF CC N-830-1 ................................6-1


6.1 Introduction....................................................................................................................6-1
6.2 Past Use of the Wallin Master Curve .............................................................................6-2
6.2.1 Within the ASME Code ......................................................................................6-2
6.2.2 NRC Applications ..............................................................................................6-2
6.3 Currently Foreseen Uses of the CC N-830-1 Suite of Fracture Toughness Models .......6-4

7 SAMPLE PROBLEMS AND RESULTS ................................................................................7-1


7.1 Introduction....................................................................................................................7-1
7.2 The Sample Problem .....................................................................................................7-1
7.2.1 Allowable Toughness Values .............................................................................7-2
7.2.2 Allowable Flaw Size Values ...............................................................................7-3

8 SUMMARY AND CONCLUSIONS........................................................................................8-1

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

A SAMPLE PROBLEM 1 STATEMENT ................................................................................. A-1

B SAMPLE PROBLEM 2 STATEMENT ................................................................................. B-1


Sample Problem 2 Statement ............................................................................................ B-1

xiii
C DRAFT CC N-830-1 (VERSION USED FOR SAMPLE PROBLEM 2)................................. C-1
Direct Use of Fracture Toughness for Flaw Evaluations of Pressure Boundary
Materials in Class 1 Ferritic Steel Components .................................................................. C-1
Section XI, Division 1 ..................................................................................................... C-1
-1000 Scope............................................................................................................. C-1
-2000 Reference Temperature ................................................................................. C-2
-3000 Toughness Variability ..................................................................................... C-3
-4000 Toughness Curves ......................................................................................... C-4
-4100 Cleavage Crack Initiation toughness, KJc ................................................... C-4
-4200 Cleavage Crack Arrest Toughness, KIa ...................................................... C-5
-4300 Ductile Crack Initiation Toughness, JIc ....................................................... C-5
-4400 Ductile Crack Extension Toughness, J-R and JX........................................ C-6
-5000 Applicability Limits .......................................................................................... C-7
-5100 Ductile Crack Extension Range ................................................................. C-7
-5200 Lower Temperature Limits on KJc and KIa................................................... C-7
-5300 Upper Temperature Limits on JIc, J-R, and Jx............................................. C-7
-5400 Intermediate Temperature Limits ............................................................... C-7
-6000 Units Conversions .......................................................................................... C-8
-7000 Nomenclature ................................................................................................. C-8

xiv
LIST OF FIGURES

Figure 1-1 Appendix A Flaw Evaluation Procedure to Evaluate the Continued


Serviceability of Ferritic Components. ..............................................................................1-4
Figure 1-2 KIc Curve (top) and KIa [KID] Curve (bottom) Referenced to RTNDT ([6]). ...................1-7
Figure 3-1 Comparison of the temperature dependence exhibited by the JIc data for the
EURO Forge with the model proposed in [65] (i.e., Eqn. (3-15) with uncertainty
bounds based on Eqn (3-16)). ........................................................................................3-14
Figure 3-2 Comparison of the revised JIc model, Eqns. (3-9) and (3-10), with JIc data
from steels having three different upper shelf toughness (JIc(288)) levels [65]. ..................3-16
Figure 4-1 Relationship between TUS and T0 [11, 14]. ..............................................................4-2
Figure 4-2 Schematic illustrating the relationship between the transition and upper shelf
toughness, and defining TUS as the intersection of the Wallin MC and the upper
shelf MC. ..........................................................................................................................4-3
Figure 4-3 Illustration of variation in the temperature separation between the KIa and KJc
master curves as a function of T0 [73]...............................................................................4-5
Figure 4-4 Illustration of the effects of strain rate increase on yield strength elevation for
materials having different degrees of prior strain hardening [61]. ......................................4-6
Figure 5-1 Schematic illustration of physical causes for systematic over-estimation of
flaw size using UT. ...........................................................................................................5-4
Figure 5-2 Cumulative probability distribution function showing the relationship between
RTNDT and T0. ...................................................................................................................5-9
Figure 5-3 Plot of KJc, 1% MC bound, 99% MC bound, the RTNDT-indexed KIc curve, and
the RTT0-indexed KIc curve. ..............................................................................................5-9
Figure 5-4 Plot of KJc, 1% MC bound, 99% MC bound, the RTNDT-indexed KIc curve
divided by √2, and the RTT0-indexed KIc curve divided by √2 (for emergency/faulted
operating conditions). .....................................................................................................5-10
Figure 5-5 Plot of KJc, 1% MC bound, 99% MC bound, the RTNDT-indexed KIc curve
divided by √10, and the RTT0-indexed KIc curve divided by √10 (for normal/upset
operating conditions). .....................................................................................................5-10
Figure 5-6 Plot of KJc, 1% MC bound, 99% MC bound, the CC N-830 5% MC bound, the
CC N-830 5% MC divided by √2, and the CC N-830 5% MC divided by √10 (for
emergency/faulted and normal operating conditions, respectively). ................................5-11
Figure 5-7 Plot of KJc, with the 1% MC bound, 99% MC bound, the CC N-830-1 1% MC
bound, the CC N-830 5% MC bound divided by √2, and the Appendix A RTT0-
indexed KIc curve divided by √2 (all for emergency/faulted operating conditions). .........5-13

xv
Figure 5-8 Plot of KJc, 1% MC bound, 99% MC bound, the 0.5% MC bound (CC N-830-
1), the 5% MC bound divided by √10 (CC N-830), and the RTT0-indexed KIc curve
divided by √10 (Appendix A) for normal/upset operating conditions. ...............................5-14
Figure A-1 Sample Problem Properties ................................................................................... A-2
Figure C-1 Illustration of Intermediate Temperature Limits when 5th Percentile Bounding
Curves are used .............................................................................................................. C-8

xvi
LIST OF TABLES

Table 4-1 RMSD values for different product forms. .................................................................4-8


Table 4-2 Composition of the J-R curve database....................................................................4-9
Table 6-1 Summary of unirradiated RTT0 value for various Linde 80 weld wire heats. .............6-4
Table 7-1 Material Properties for use in Appendix A and Proposed Code Case N-830-1
Sample Problem 2. ...........................................................................................................7-2
Table B-1 Material Properties for use in the Sample Problem 2 .............................................. B-2
Table B-2 Table for Presentation of Results of the Phase II Sample Problem ......................... B-3
Table C-1 Values of p and Mp Corresponding to Different Bounding Toughness Curves......... C-3
Table C-2 RMSD values for different product forms. ............................................................... C-7
Table C-3 Unit Conversion Coefficients................................................................................... C-8
Table C-4 Symbols ................................................................................................................. C-9
Table C-5 Definitions............................................................................................................. C-10

xvii
1
INTRODUCTION

Historically, the safety of nuclear power plant pressure-retaining components has been
demonstrated using the rules of the ASME Boiler and Pressure Vessel Code (ASME Code, or
Code). Section III of the ASME Code provides Rules for Construction of Nuclear Facility
Components, and Section XI provides Rules for In-Service Inspection of Nuclear Plant
Components. Both sections of the Code provide methods for assessing stresses and moments
contributing to the forces available to drive crack growth in components containing postulated or
detected flaws. The Code primarily makes use of linear elastic fracture mechanics (LEFM)
methods to calculate stress intensity factors, and has fracture toughness models based on
empirical data to estimate material resistance to crack extension. Much of the current Code is
based on LEFM models of material behavior in the presence of flaws that were developed more
than 40 years ago at a time when drop-weight tests [1] and Charpy V-notch (CVN) impact tests
[2] were the accepted standards used to estimate metrics that correlate with fracture toughness,
such as the nil-ductility temperature (NDT) or ductile-to-brittle transition temperature (DBTT).
The ferritic steels used to fabricate nuclear power plant reactor pressure vessels (RPVs) were
selected to have sufficient strength and toughness to provide adequate safety margins against
overload failure and catastrophic crack extension at all operating temperatures and conditions.
To ensure adequate toughness, the RPV steels selected for power plant construction in the 1960s
and 1970s, were chosen to have a DBTT well below the expected operating conditions of the
plant. The CVN and the drop-weight tests were among the most commonly used test methods
for characterizing the DBTT temperature of these steels at that time. However, these tests do not
directly provide the specimen-independent measures of fracture toughness required to support an
ASME Code analysis. These test results can only be correlated to the measure of the material’s
resistance to crack extension. Linear-elastic plane-strain fracture toughness testing, as prescribed
by ASTM Standard E399 [3], was developed to provide a direct measure of a material’s
resistance to crack extension using a measure of the critical stress intensity factor required for
crack extension, KIc. Such a value allows for more direct comparison to the crack driving force
in stress intensity factor calculations.
Linear-elastic plane-strain fracture toughness testing for RPV materials often requires large
specimens to ensure that validity criteria for small scale yielding are met, and the test specimens
and procedures are often expensive. As such, testing an adequate number of specimens to fully
define a fracture toughness transition curve and reference temperature is expensive. Because of
this, the nil-ductility test (used to define NDT) and the CVN test, both of which use smaller test
specimens and simpler test procedures compared to those required for valid KIc determination,
became the dominant methods for characterizing material toughness transition temperature,
RTNDT, defined as the reference temperature for nil ductility transition to signify the reference
temperature below which a material exhibits limited ductility in the presence of a notch.
Calculation of RTNDT from a combination of data from NDT and CVN testing is described in
Paragraph NB-2331 of Section III of the ASME Code [4]. The prevalence of NDT and CVN

1-1
Introduction

data, combined with work performed to correlate these values with KIc [5, 6], resulted in an
RTNDT-referenced KIc curve that was adopted into the ASME Code, Section XI, Appendix A
flaw assessment procedures [7].
There is uncertainty inherent in both the RTNDT and KIc values determined for a specific material.
This uncertainty is caused by the natural material inhomogeneity that controls fracture behavior,
and the uncertainties surrounding modeling assumptions, test procedures, and analytical methods
used for determining these values. These uncertainties can be treated explicitly by quantifying
the uncertainties in these values (defining their distributions) and then either taking a lower
bound value, or assigning a factor that is applied to the best-estimate value that directly accounts
for the uncertainties. If the distributions in the data are not well established, the uncertainties can
be treated implicitly by making conservative assumptions about the operating conditions or using
conservative models of material behavior. Explicit treatments of uncertainties are preferred, as
they are more transparent, their impact more easily understood, and they can more readily be
changed as knowledge and information are expanded. The method employed in ASME Code,
Section XI, Appendix A uses both implicit and explicit treatments of uncertainties, which
obscures accurate representation of material behavior and increases the difficulty of taking
advantage of increased knowledge of material properties.

1.1 ASME Section XI, Appendix A Approach


ASME Code, Section XI, Nonmandatory Appendix A, “Analysis of Flaws,” [7] provides
analytical procedures for use in determining the acceptability of flaws for continued service that
are detected during inspection and that exceed the flaw acceptance standards of IWB-3500. The
procedures are based on LEFM principles and apply to ferritic components with wall thicknesses
of 100 mm (4 inches) or greater, and having simple geometries and stress distributions.
Appendix A is limited to ferritic steels having a minimum yield strength of 350 MPa (50 ksi) or
less, and provides procedures for three areas of flaw assessment: (1) characterization of the flaw
size, shape, and location for use in the LEFM analysis, (2) methods for performing crack driving
force (stress intensity factor) calculations, and (3) methods for determining allowable material
properties (fracture resistance) to be used in the analyses.

1.1.1 Summary of ASME Section XI, Appendix A Flaw Evaluation Procedures


Appendix A, developed based on the work described in References [5, 6] describes a method that
can be used to determine whether a ferritic steel component with a detected flaw that exceeds the
IWB-3500 flaw acceptance criteria is acceptable for continued use. The methods involve the
following steps [7]:
1. Determine the actual flaw configuration in accordance with IWA-3000.
2. Characterize the flaw in accordance with IWB-3610.
3. Resolve the flaw into a simple shape that can be readily analyzed.
4. Determine the stresses at the location of the observed flaw for normal, emergency, and
faulted conditions.
5. Calculate stress intensity factors for each condition.
6. Determine the necessary material properties, including the effects of irradiation, if applicable.

1-2
Introduction

7. Determine the following critical flaw parameters:


a) af = expected end-of-life flaw size
b) ac = minimum critical flaw size for normal conditions
c) ai = minimum critical initiation flaw size for emergency and faulted
conditions.
8. Using the critical flaw parameters, af, ac and ai, apply the flaw evaluation criteria of IWB-
3600 to determine whether the observed flaw is acceptable for continued service.
The methods described in Appendix A are based in LEFM, and therefore only apply to
conditions for which the ferritic steel of interest exhibits lower transition fracture toughness
behavior. For conditions in which the ferritic steel of interest exhibits upper shelf toughness
behavior, Appendix A is not applicable. For vessels in these situations the methodology
described in Nonmandatory Appendix K, “Assessment of Reactor Vessels with Low Upper Shelf
Charpy Impact Levels,” [8] may be used. The Code does not provide any guidance to identify
when this transition in behavior occurs.
Figure 1-1 contains a diagram showing the steps of the Appendix A process that are described in
1-8 above. The steps are separated into the major components of the process, as depicted by the
Flaw Evaluation, Crack Driving Force Calculation, and Resistance Calculation shaded areas.
Flaw evaluation procedures are described in Article A-2000 of Appendix A, and the procedures
for calculating the stress intensity factors for use in assessing flaw growth and acceptability are
described in Article A-3000. The procedure for determining the appropriate material property
for use in comparing to the stress intensity factor is described in Article A-4000. Of particular
note, Article A-4000 specifies the lower bounding fracture toughness curve, KIc, as a function of
the difference between the metal temperature, T, and the material RTNDT determined in
accordance with Paragraph NB-2331 of Section III of the ASME Code [4]. Where appropriate,
RTNDT is adjusted to account for the effects of neutron irradiation embrittlement. Use of RTNDT
as a reference temperature value for the KIc curve provides a level of implicit conservatism that
has proven to be sufficiently conservative with respect to expected material toughness behavior
[5]. However, the level of conservatism is unknown and varies for different materials because
neither RTNDT nor KIc provide accurate representations of material fracture toughness behavior.

1-3
Introduction

Figure 1-1
Appendix A Flaw Evaluation Procedure to Evaluate the Continued Serviceability of Ferritic
Components.

1.1.2 Treatment of Uncertainties in Appendix A


RTNDT is not a true measure of material fracture behavior but only an indication of the
temperature below which a material exhibits little to no ductility. Because RTNDT is determined
using a combination of nil-ductility drop-weight and Charpy v-notch data, there is a lot of scatter
inherent in RTNDT values, and inconsistency in how well a material’s actual ductile-to-brittle
transition temperature (DBTT) is represented by RTNDT. The ASME NB-2331 methods for
determining RTNDT were developed to provide conservative estimates of the DBTT.
The use of RTNDT to reference the KIc curve provides a level of implicit conservatism believed
sufficient to very conservatively bound expected material fracture toughness behavior [5]. It is
difficult to quantify the conservatism provided by RTNDT because neither RTNDT nor the KIc lower
bound curve provide accurate representations of material fracture toughness behavior. This
inaccuracy results in an inconsistent treatment of uncertainties as the level of conservatism (and
accounting for uncertainties) varies with each material.
A relatively recent change to Appendix A includes use of a best-estimate T0 value, if it is
available for the material of interest, to calculate an RTT0 value to be used as the reference
temperature adjustment for the KIc curve instead of RTNDT. While T0 is considered an accurate
representation of a material’s fracture toughness behavior, the ASME Code adds a “margin” by
defining RTT0 as:
RTT0= T0 + 19.4°C or RTT0= T0 + 35°F Eq. 1-1

1-4
Introduction

to ensure a very conservative estimate of material resistance to crack growth.


Further conservatism is added in the ASME approach when comparing the KIc values to the
crack driving forces based on the criteria required by Paragraph IWB-3612 of Section XI. The
allowable stress intensity factor criteria impose additional structural factors depending on the
applicable Service Level: KIc must exceed KI√10 for normal operating conditions, and must
exceed KI√2 for postulated emergency or faulted conditions. The bases for the use of the factors
of √2 and √10 are described in Reference [6]. This explicit use of these structural factors further
adds additional conservatism with unquantified uncertainty to the Appendix A approach.
Although many changes have been implemented in Appendix A since it was first published, the
RTNDT-referenced KIc curve still provides a conservative method for characterizing material
resistance to crack extension. The technical bases for Appendix A are documented in the
Welding Research Council (WRC) Bulletin 175 [5] and EPRI Report NP-719-SR [6]. Together,
these two documents define flaw characterization methods, material fracture toughness curves,
and crack driving force calculation procedures currently contained in Appendix A.

1.1.3 Technical Basis for the Appendix A Methodology


In 1971, the Pressure Vessel Research Committee (PVRC) of the WRC undertook the task to
review research and make recommendations on toughness requirements for ferritic materials in
nuclear power plant components to both the ASME Boiler and Pressure Vessel Committee and to
the Atomic Energy Commission. Specifically, these recommendations, based on knowledge that
was current at the time, were intended to provide ferritic-material-toughness requirements for
pressure retaining components of the reactor coolant pressure boundary operating below 370°C
(700°F). The goal was that these criteria, together with the stress limits allowed by the ASME
Code, would permit the establishment of safe operating procedures for nuclear reactor
components under normal, upset, and test conditions.
The requirements and recommendations arising from WRC 175 [5] that pertain to Appendix A,
Article A-4000, “Material Properties,” included the following:
• A lower bound, temperature-dependent KIa curve was defined based on a curve drawn below
all the KId (the stress intensity factor under dynamic loading) and KIa data available at the
time, referenced to the drop-weight nil-ductility temperature, TNDT (KIc data was too high to
have any impact on the lower bounding curve) as shown in Figure 1-2. The equation to
describe the temperature dependence of KIa is given by:

KIa = 13.675 exp [0.0261 (T- TNDT)] + 29.4 (in units of MPa√m, oC)
KIa = 1.223 exp {0.0145[T- (TNDT +160)]} + 26.8 (in units of ksi√in, °F) Eq. 1-2

This KIa curve was then termed the reference toughness KIR curve and was indexed using TNDT to
eliminate the need for performing expensive KIc tests.
• To ensure that the transition temperature used to reference the KIa curve was well below the
upper shelf temperature for the material of interest, a criterion was described that combined
nil-ductility test results to establish TNDT, and CVN tests to define the temperature at least
33°C (60°F) above TNDT at which Charpy specimens exhibited at least 0.89 mm (35 mils =
0.35 in.) of lateral expansion. These two results were combined to define RTNDT as T0.89(mm) –

1-5
Introduction

33°C (T35(mils) - 60°F), such that RTNDT was defined as the higher of TNDT or T0.89(mm) – 33°C
(T35(mils) - 60°F). An alternate requirement involving both 0.89 mm (35 mils) of lateral
expansion and a minimum CVN energy of 68 J (50 ft-lb) was also suggested, i.e., T68J – 33°C
(T50(ft-lb) - 60°F). The minimum of the CVN energy and the lateral expansion criteria were
recommended to eliminate materials that might have a low transition temperature or very low
upper shelf energies from consideration using the KIR-RTNDT procedure.
• A very conservative defect size that included a depth of one-quarter of the wall thickness
(¼t), a length of six times the depth (or 1.5t), a sharp crack tip, and an orientation
perpendicular to the maximum stress direction was recommended.
• A safety factor was recommended for application to the crack driving force stresses along
with a flaw size safety margin by recommending a reference flaw size considerably larger
than the actual or anticipated maximum flaw size.
• Procedures for calculating the allowable loading were presented in an Appendix to WRC 175
[5] that involved primary membrane stresses due to pressure and secondary thermal stresses
caused by thermal gradients near the crack tip.
• Additional safety factors on loading beyond safety factors between 1.0 and 2.0 applied to
stresses were not recommended as these were believed to be outside the scope of the PVRC.
The recommendations presented in WRC 175 [5] were modified by Marston, et al. [6] of the
newly formed ASME Section XI Working Group on Flaw Evaluation before they were
implemented into Appendix A. The modifications included using RTNDT to index the KIa curve
instead of TNDT. A bounding KIc curve was defined by drawing a curve beneath all the available
static plane strain fracture toughness data referenced to RTNDT for the same materials, in a
manner similar to that used to define the KIa curve, defining the temperature dependence of KIc
as:
KIc = 36.5 + 22.783 exp [0.036 (T - RTNDT)] (in units of MPa√m, °C)
KIc = 33.2 + 20.734 exp [0.02 (T - RTNDT)] (in units of ksi√in, °F) Eq. 1-3

The KIa and KIc curves are shown in Figure 1-2.


While no upper shelf toughness was defined in WRC 175, the EPRI report, or Appendix A,
Marston, et al. recommended that the calculations made using the newly proposed J-integral and
equivalent energy methods showed upper shelf toughness exceeded 220 MPa√m even for
irradiated materials [7, 8]. Based on this, a value of 220 MPa√m has been used by many analysts
to define the upper limit of applicability for the KIc curve.

1-6
Introduction

Figure 1-2
KIc Curve (top) and KIa [KID] Curve (bottom) Referenced to RTNDT ([6]).

1-7
Introduction

Because there have been no changes made to these material toughness curves since they were
originally published in the 1970s, they retain their inherent conservatisms. These two curves are
both similarly affected by degradation due to irradiation through ∆RTNDT. Because both curves
are referenced to T-RTNDT, and since irradiation embrittlement is characterized by the
temperature at which the CVN energy is 41 J (30 ft-lb), ∆T41J, the separation between KIc and KIa
does not change with irradiation. The upper limit of applicability for the linear elastic KIc curve
of 220 MPa√m (200 ksi√in), which is sometimes assumed by analysts, also does not change with
irradiation. However, upper shelf Charpy energy values falling below 68 J (50 ft-lb) 2 require an
Equivalent Margins Assessment (EMA) by Appendix G to 10 CFR Part 50 [9]. An EMA can be
performed using USNRC RG 1.161 [10], or using elastic-plastic fracture mechanics (EPFM) as
described in Appendix K of the ASME, Section XI Code, or using other similar methods that
have been developed, such as the Owner’s Groups evaluations representing different nuclear
steam supply system (NSSS) vendors.

1.1.4 Issues with Appendix A Methodology


Issues arise in use of the Appendix A method in two main areas:
1. Unquantified uncertainties that are treated both implicitly and explicitly, which result in a
very conservative bias in the characterization of material resistance to fracture.
2. Use of models that do not consistently represent material behavior resulting in varying, and
unknown, degrees of conservative bias in different situations.
The conservatisms inherent in the Appendix A approach arise, in large part, due to the continued
use of RTNDT to reference the KIc and KIa curves, as well as the shape (i.e., temperature
dependence) of the curves themselves. While correlations have been established between KIc
and CVN behavior versus temperature and irradiation, conservatisms remain and additional
margins are typically added to account for material inhomogeneity, uncertainty in material
property information available for specific materials, and uncertainties in the use of CVN to
provide a measure of material fracture toughness [5, 6]. Structural factors are also added to the
estimate of material fracture toughness to account for uncertainties in the knowledge of the
stresses at the crack tip used to calculate the crack driving force for crack growth.
Considerable advancements to the state-of-knowledge, both theoretical and practical, have
occurred since WRC 175 and NP-719-SR were published, particularly with regards to the
amount of available data. These data were used in recent studies to develop a set of integrated,
best estimate models that predict the mean trends and scatter of fracture toughness of ferritic
steels throughout the temperature range between lower shelf behavior, through DBTT to upper
shelf behavior [11, 12]. Comparisons of these new models to the Appendix A methodology
reveal areas where Appendix A is inconsistent with trends predicted by the large amount of
ferritic steel fracture toughness data now available. Some of these areas include:
1. On the lower shelf, the low-temperature asymptote of the Appendix A KIc curve does not
represent a lower bound to all available data resulting in a non-conservative bias. The ASME
model over-estimates the lower shelf fracture toughness at temperatures ≈60°C or more
below RTT0 for all values of RTT0. For un-irradiated materials, such low temperatures cannot
2
Parameters in the historical discussion were presented in both Metric and English units. All subsequent discussion
will be in terms of metric units.

1-8
Introduction

be achieved during normal operations. However, as radiation embrittlement causes the


material transition temperature to approach regulatory limits (e.g., the PTS limits of 132 and
149°C in 10CFR50.61 [13]), a temperature 60°C below these values may be within the
achievable temperature range during a cool-down event.
2. The temperature dependence of the Appendix A KIc curve does not accurately reflect the
temperature dependence of transition toughness data at all temperatures. This results in a
conservative bias, particularly in the lower transition temperature region.
3. On the upper shelf, the KIc limit of applicability of 220 MPa√m exceeds available data,
especially after consideration of irradiation effects. This could result in a non-conservative
bias in estimates of fracture toughness. Above RTT0 of 0°C, 220 MPa√m exceeds the upper
shelf fracture toughness of most RPV steels by a considerable amount, suggesting a practical
limit on KIc should be informed by the upper shelf fracture toughness. The upper shelf of
many ferritic materials falls below 220 MPa√m, even when J0.1 3 (the value of J at 0.1 inch
crack extension, or 2.54 mm of crack extension) is used as the characterizing parameter. The
temperature at which the mean KJc curve equals the mean JIc curve, TUS, can be used to define
the upper limit of applicability for the KIc curve based on supporting data [11, 14].
4. The separation between the KIc and KIa curves depends on the amount of irradiation
embrittlement, a functionality not captured by the Appendix A equations. Recently
developed data-based models show that, as RTT0 increases, the KIc and KIa curves converge.
This convergence is not a feature of the Appendix A curves, which maintain a constant
temperature separation. The result of using a constant temperature separation to represent the
actual material behavior is that the Appendix A model is overly pessimistic for high values of
RTT0, indicative of highly irradiated material, and the model over-estimates KIa at low values
of RTT0.
5. The temperature above which upper shelf behavior can be expected depends on the amount
of irradiation embrittlement, a functionality not captured in the Appendix A equations.
With the development of best-estimate, probabilistic models of material fracture toughness
behavior and a clearer understanding of the mechanisms driving crack extension in ferritic steels,
the conservatisms inherent in the Appendix A fracture toughness models are no longer necessary.
With increased knowledge, extensive additions of data, and better test methodologies, the
uncertainties for these models may be quantified and set appropriately to reflect the uncertainties
for Charpy, Nil-Ductility and LEFM-based models. In many cases, measurements of both
unirradiated and irradiated KJc fracture toughness are available for various limiting RPV
materials, providing a means by which direct comparisons between crack driving force, KI, and
the material resistance, KJc, can be made. This knowledge provides a technical and rational basis
for the reduction, or elimination, of the conservatism and unnecessary margins that may prohibit
continued plant operation. Test methodologies are continuing to improve with better,
standardized ASTM testing procedures to determine more reliable KJc values from miniature
specimens, thereby enabling use of surveillance CVN specimen for determining irradiated T0

3
Throughout this report we adopt the common nomenclature used in the international literature of J0.1 (representing
the value of J at 0.1 inch of crack extension, and equivalently, representing the value of J at 2.54 mm of crack
extension). All formulas for Jx (J at x mm of crack extension) are in metric units with x representing the amount of
crack extension in millimeters.

1-9
Introduction

values. All these advances provide for superior modelling and a sound technical basis for
incorporating more accurate, T0-based models into Appendix A of the ASME Code.

1.2 Objectives of Proposed Code Case N-830-1


The objective of Code Case N-830-1 is to implement an integrated suite of best-estimate fracture
toughness models that can all be determined from a knowledge of a material-specific T0 value for
use in component flaw evaluations. These models are appropriate for use in both deterministic
and probabilistic assessments, as each model describes the full distribution in values about the
mean for any temperature and material condition. Specific goals for CC N-830-1 implementation
include:
1. To ensure that material properties are accurately represented by the latest available best-
estimate models, that uncertainties in fracture toughness are well quantified, characterized,
and explicitly treated either by use of bounding values or well-understood margins,
2. To ensure that fracture toughness models are appropriately linked to consistently account for
the effects of hardening and irradiation,
3. To take advantage of varying degrees of knowledge regarding material properties. For
example, measured T0 values for a given material should result in lower margins, and T0
values established through correlations should result in larger margins, and
4. To enable predictions and estimates of all toughness values for any level of embrittlement
from knowledge of a single value, T0.
This report describes the technical bases for new fracture toughness models contained in Code
Case N-830-1 that satisfy the above objectives. This document was prepared by a small task
group to provide information to the Working Group on Flaw Evaluation to support finalization
and decision-making on CC N-830-1. Chapter 2 provides a summary of the information
contained in CC N-830-1. Chapters 3 and 4 provide information supporting all of the models
contained in CC N-830-1 including the data supporting the empirical derivations, the physical
basis for the trends observed in the models, and work performed in validation of the models.
Chapter 5 discusses sources and treatment of uncertainty in Appendix A calculations including
comparisons between treatment of uncertainty in the current Appendix A, the original CC N-830
and, the proposed CC N-830-1. Chapter 6 discusses applications for CC N-830-1, and Chapter 7
describes results of a several sample problems worked by the WGFE in support of CC N-830-1
development. This report ends with a summary of information supporting CC N-830-1 in
Chapter 8.

1-10
2
OVERVIEW OF CODE CASE N-830-1

2.1 Introduction
With standardization of the test methodology for obtaining the T0 fracture toughness reference
temperature in ASTM E1921 [15] the stage was set for implementation of T0 into the ASME
Code. This implementation occurred via the adoption of two Code Cases: N-629 in Section XI
and N-631 in Section III [16, 17]. These Code Cases proposed use of a T0-based reference
temperature for use in indexing the ASME’s KIc curve in Section XI and Section III Code
applications. RTT0 is a T0-based transition toughness reference temperature defined by Equation
(1-1).
Code Case (CC) N-830 was approved by ASME in 2014, and was the first direct implementation
of the KJc Master Curve (MC) into the ASME Code [18]. The CC made use of the 5th percentile
lower bound of the Wallin Master Curve as an alternative to the ASME KIc curve to characterize
material resistance to fracture in flaw evaluations. Since that time, work has progressed within
the ASME Section XI Working Group on Flaw Evaluation (WGFE) to expand and improve the
original CC methods.
To take advantage of the best-estimate fracture toughness models recently developed and linked
to T0, CC N-830 was modified to include a suite of self-consistent fracture toughness models
describing material fracture toughness behavior from lower shelf, through transition, to upper
shelf behavior [19]. These models include linkage models that describe the inter-relationships
controlling changes in toughness behavior for all toughness parameters with irradiation. The
proposed Revision 1 of CC N-830 incorporates a complete suite of best-estimate models that
completely describe the temperature dependence, scatter, and interdependencies (such as those
resulting from irradiation or other hardening mechanisms) between all fracture toughness metrics
(i.e., KJc, KIa, JIc, J0.1, and J-R). By incorporating both a statistical characterization of fracture
toughness, and the ability to estimate a toughness curve for any percentile bound, CC N-830-1
provides a consistent basis for the conduct of both conventional deterministic flaw evaluations,
as well as probabilistic evaluations. Additionally, both transition and upper shelf toughness
properties are defined in a consistent manner in one document to provide the analyst an easy
means to determine what fracture behavior (i.e., transition or upper shelf) can be expected for
any condition.

2-1
Overview of Code Case N-830-1

2.2 CC N-830-1 Contents

2.2.1 Inquiry
The inquiry for CC N-830-1 is as follows:
“What current best-estimate (alternative) fracture toughness models and
relationships may be used for flaw evaluations performed in accordance with
Nonmandatory Appendix A and/or Nonmandatory Appendix K in lieu of the
current requirements of these Appendices for the values of KIc, KIa, JIc, J0.1,
and J-R?”

2.2.2 Reply
The initial portion of the reply for CC N-830-1 is as follows:

“It is the opinion of the Committee that the fracture toughness models based on
the Master Curve Method in accordance with ASTM E-1921 may be used in lieu
of the current requirements of Nonmandatory Appendices A or K when
determining values for KIc, KIa, JIc, J0.1, and J-R using the procedures and
equations given below.”

CC N-830-1 uses a T0 value measured in accordance with ASTM Standard E1921, “Standard
Test Method for the Determination of Reference Temperature, To, for Ferritic Steels in the
Transition Range” [15]. Using T0, it is possible to estimate the variation of fracture toughness
with temperature across the entire range of interest to operating vessels for Class 1 ferritic
reactor pressure vessel (RPV) materials. This estimate can be used as an alternative to:
• The crack initiation fracture toughness curve, KIc, of Nonmandatory Appendix A, Subarticle
A-4200 for pressure retaining materials other than bolting, and
• The crack arrest fracture toughness curve, KIa, of Nonmandatory Appendix A, Subarticle
A-4200 for pressure retaining materials other than bolting, and
• The J-integral fracture resistance for the material at a ductile flaw extension of 0.1-in.
(2.5 mm), J0.1, of Nonmandatory Appendix K for pressure retaining materials other than
bolting.
• The J-integral fracture resistance for the material and its variation with ductile flaw
extension, Δa, J-R, of Nonmandatory Appendix K for pressure retaining materials other than
bolting.
The remaining content of the reply to CC N-830-1 defines all these relationships.

2.2.3 Discussion of N-830-1


CC N-830-1 defines best-estimate models of fracture toughness for implementation into the
ASME Code. The CC provides definitions for temperature dependencies and distributions for
best-estimate models that describe fracture toughness behavior from the lower shelf, through the
transition region, and to the upper shelf. In addition, the linkage models that enable consistent
representation of fracture toughness behavior across all temperatures and conditions are defined.

2-2
Overview of Code Case N-830-1

These models are based on large databases of measured fracture toughness values such that
uncertainties in values are well understood, characterized, quantified and validated. Full
distributions are defined for each fracture toughness parameter to enable consistent and explicit
treatment of uncertainties in both deterministic and probabilistic assessments.
CC N-830-1 defines bounding values for percentiles of interest for use in deterministic
evaluations. The linkage models ensure consistent data bounds for all material hardening
conditions provided the same percentiles are selected for the transition and upper shelf toughness
models. The distributions defined for all parameters provide the information required to sample
across the expected range of values at each temperature for probabilistic assessments, with the
linkage models ensuring that these distributions change synchronously, as material conditions
change.
The materials fracture toughness models presented in CC N-830-1 are meant to be used in lieu of
those described in Article 4000 of the current Appendix A for describing material resistance to
fracture. Lower bounding values of the distributions are recommended for use in deterministic
flaw analyses. Use of the same percentile bounding value for all toughness curves from lower
shelf through upper shelf, coupled with elimination of structural factors, ensures consistent
representation of material behavior for all fracture modes. While these recommendations for CC
N-830-1 apply explicitly only to the fracture toughness parameter, we argue in Chapter 5 that
explicit factors applied to stress and flaw size are not needed due to the conservatisms inherent to
non-destructive flaw sizing and the analytical determination of stresses.
To ensure that the uncertainties inherent in all aspects of flaw analysis are treated appropriately
and consistently, the WGFE plans to develop explicit partial structural factors to apply to each
parameter (flaw characterization, driving force analysis and material resistance) in the flaw
analysis to more accurately reflect the uncertainties in that specific parameter. The
recommendations contained in CC N-830-1 for appropriate bounding values to use for material
fracture toughness only account for the uncertainties in the fracture toughness parameter.
The best estimate fracture toughness models, linkage models, and technical bases for the models
contained in CC N-830-1 are presented in Chapters 3 and 4. The model temperature dependence
and distributions are described, along with a summary of their development and validation,
including limitations on their use.

2-3
3
FRACTURE TOUGHNESS MODELS IN CC N-830-1

There are three basic fracture toughness models presented in CC N-830-1; KJc, and JIc to describe
the initiation fracture toughness in the transition region and on the upper shelf, and KIa to
describe the crack arrest fracture toughness. Measured values of KJc and KIa are used to define
indexing temperatures (T0 and TKIa, respectively) that are material specific. These indexing
temperatures were used to normalize the KJc and KIa curves to establish a single temperature
dependence for each curve. The JIc curve was normalized to establish the temperature
dependence of JIc for all ferritic steels based on the JIc value at 288 °C, JIc288. Each of these
models was empirically derived from large databases of toughness values, but the model forms
were informed from a mechanistic understanding of the fracture process that provides a
theoretical underpinning to identify empirical trends. The assumption used for all models was
that applied energy absorption by dislocation motion prior to fracture is the mechanism
controlling the temperature dependence of the fracture toughness. Therefore, these models are
applicable only to ferritic steels, and only in temperature regions where deformation is
dislocation-dominated.

3.1 Cleavage Crack Initiation Toughness, KJc


Wallin, working in collaboration with Sarrio and Törrönen, began to publish papers that became
the basis for what is now referred to as the Master Curve as part of his doctoral research work in
1984. This work includes two components: a statistical model of cleavage fracture, and a
temperature dependency of fracture toughness common to all ferritic steels [20, 21]. The
concept includes:
• a weakest-link failure model that uses a 3-parameter Weibull function to describe the
distribution of fracture toughness values at a fixed temperature,
• a temperature-dependence described by an inverse Peierls-Nabarro relationship, and
• a methodology to account for the effect of crack front length (size effects) on fracture
toughness.
Wallin observed that the temperature dependence of fracture toughness is not sensitive to steel
alloying, heat treatment, or irradiation [22]. This observation led to the concept of a universal
temperature-dependent curve shape for all ferritic steels. Several investigators have empirically
assessed the validity of the universal curve shape for both unirradiated and irradiated nuclear
RPV steels, with favorable results [23, 24]. These research and development activities led to
publication of an ASTM Standard (E1921-97) to estimate the MC index temperature (T0) [15],
and, later, to adoption of Code Case N-629 and N-631 in Section XI and Section III of the
ASME Code that uses T0 to establish an index temperature (RTT0) for the KIc and KIR curves [16].
Note that Code Case N-629 has been replaced by Code Case N-851 to include the proper
relationship to TKIa [25]. Code Case N-851 also has been incorporated into Section XI of the

3-1
Fracture Toughness Models in CC N-830-1

Code in both Appendices A and G, and Code Case N-631 is in the process of being included in
Section III in NB-2300 [26]. Research, and test and evaluation programs, aimed at enhancing the
wealth of information collected on the applicability of the MC and the MC methodology, provide
support for direct implementation of the MC and T0 for use in assessing the fracture safety of
critical RPV components.

3.1.1 Description of the KJc Model


The model developed by Wallin, et al. [20] was first presented as an expression for the
probability of failure, Pf, of a cracked specimen using the simple Weibull form:
𝑩𝑩 𝑲𝑲 −𝑲𝑲 𝟒𝟒
𝑷𝑷𝒇𝒇 = 𝟏𝟏 − 𝒆𝒆𝒆𝒆𝒆𝒆 �− 𝑩𝑩 �𝑲𝑲 𝑰𝑰 −𝑲𝑲𝒎𝒎𝒎𝒎𝒎𝒎 � � Eq. 3-1
𝟎𝟎 𝟎𝟎 𝒎𝒎𝒎𝒎𝒎𝒎

where B0 and K0 are normalization constants, KI is measured toughness and Kmin is taken as
20 MPa√m. B0 can be set to any desired specimen reference thickness but is usually taken as
25.4 mm. K0 is the temperature dependent scale parameter taken as the 63.2% probability of
fracture for a specimen of thickness B0. The shape parameter is defined by the exponent of 4.

3.1.2 Basic Form


Wallin first suggested a temperature dependence common to all ferritic steels in work published
from his Ph.D. dissertation [21]. Further work demonstrated that the alloying, heat treatment and
irradiation conditions characteristic of a particular ferritic material only influences the position of
the transition fracture toughness curve on the temperature axis, while the variation of cleavage
fracture toughness with temperature follows a common form irrespective of these factors [22].
Definition of the temperature dependence of the MC was based on analysis of irradiated and
unirradiated KJc data from Welds 72W and 73W tested as part of the Heavy Section Steel Test
(HSST) Irradiation Program at the Oak Ridge National Laboratory (ORNL) [23]. Normalizing
the data to a 25.4-mm-long crack front, the temperature dependent scale parameter, K0, was
defined by:
𝑲𝑲𝟎𝟎 = 𝟑𝟑𝟑𝟑 + 𝟕𝟕𝟕𝟕 ⋅ 𝒆𝒆𝒆𝒆𝒆𝒆[𝟎𝟎. 𝟎𝟎𝟎𝟎𝟎𝟎(𝑻𝑻 − 𝑻𝑻𝒐𝒐 )] Eq. 3-2

where T is the test temperature and T0 is the temperature at which the measure KJc value is 100
MPa√m. The median cleavage crack initiation toughness, KJc(median), curve was then defined as:
𝑲𝑲𝑱𝑱𝑱𝑱(𝒎𝒎𝒎𝒎𝒎𝒎𝒎𝒎𝒎𝒎𝒎𝒎) = 𝟑𝟑𝟑𝟑 + 𝟕𝟕𝟕𝟕 ⋅ 𝒆𝒆𝒆𝒆𝒆𝒆[𝟎𝟎. 𝟎𝟎𝟎𝟎𝟎𝟎(𝑻𝑻 − 𝑻𝑻𝒐𝒐 )] Eq. 3-3

3.1.3 Distribution
The distribution of data at any given temperature follows a Weibull distribution with a slope of
four and Kmin equal to 20 MPa√m [20, 23], as shown in Eqn. (3-1). The cleavage crack initiation
toughness, KJc, curve can be defined at any percentile, p, as follows:
𝒑𝒑
𝑲𝑲𝑱𝑱𝑱𝑱 = 𝟐𝟐𝟐𝟐 + (𝑲𝑲𝒐𝒐 − 𝟐𝟐𝟐𝟐){−𝒍𝒍𝒍𝒍(𝟏𝟏 − 𝒑𝒑)}𝟏𝟏/𝟒𝟒 Eq. 3-4

Eqn. (3-4) can be used to produce both lower and upper bound curves. For example, using a
value of 0.05 for p would produce a 5% lower bound curve, while using a value of p equal to
0.95 would produce a 95% upper bound curve. The scale factor, K0, is given by Eqn. (3-2).
There is no effect of product form or irradiation on Eqn. (3-4).

3-2
Fracture Toughness Models in CC N-830-1

Eqn. (3-4) assumes a crack front length of 25.4 mm (1 in.) in laboratory test specimens with
straight crack fronts. While an adjustment to Eqn. (3-4) that accounts for different crack front
lengths in laboratory test specimens was developed, currently there is insufficient basis to
recommend a generic equation that applies to non-straight cracks fronts (e.g., surface breaking
cracks, fully embedded cracks, etc.) that are of interest in structural analyses. CC N-830-1
therefore uses Eqn. (3-4) unless the user can demonstrate that a crack front length other than
25.4 mm (1 in.) is appropriate to the structural situation of interest.

3.1.4 Theoretical Basis


Based on dislocation mechanics considerations, Zerilli and Armstrong (Z-A) [27] described the
constitutive behavior of metals using an equation that divided the flow behavior due to loading
into dislocation mechanisms that were thermally activated, and those that were not thermally
activated. Short range barriers to dislocation motion are those described by length scales on the
order of the atomic spacing of the metal that can be affected by changes in temperature and thus
lattice atom vibration. These were included in the thermally activated terms in the constitutive
equation. Long range barriers to dislocation motion, i.e., those barriers that have inter-barrier
spacings that are orders of magnitude greater than the atomic spacing of the lattice structure, are
not affected by changes in temperature, so they were included in the non-thermally activated
constitutive equation terms. The temperature dependence of the flow stress, as derived by Z-A,
was controlled by the short-range barriers to dislocation motion with the temperature dependence
described by that of the Peierls-Nabarro stress. For body-centered cubic (BCC) metals (e.g., all
ferritic steels), the only short-range barriers to dislocation motion are the lattice atoms
themselves. All other metallurgical features, including grain boundaries, other dislocations,
point defects, precipitates and inclusions, are all considered to be long range barriers, so they do
not affect any control of the temperature-dependent behavior of ferritic steels.
Following the work by Z-A, Natishan, et al. [28-30] demonstrated that the temperature
dependence of the fracture toughness in the fracture mode transition region also depends only on
the short-range barriers to dislocation motion established by the lattice structure of the material
(BCC for ferritic steels). Other microstructural features that vary with steel composition, heat
treatment, and irradiation include grain size/boundaries, point defects, inclusions, precipitates,
and dislocation substructures; these features only influence the position of the transition curve on
the temperature axis (i.e., T0 as determined by E1921-97), but not the shape of the curve. This
understanding suggests that the myriad of metallurgical factors that can influence absolute
strength and toughness values exert no control over the form of the variation of toughness with
temperature in fracture mode transition. Moreover, this understanding provides a theoretical
basis to establish, a priori, those steels to which the MC and T0-linked equations should apply,
and those to which they should not. On this basis, the MC and T0-linked models are particularly
good at describing the temperature-dependence of the fracture toughness of all steels having an
iron BCC lattice structure (e.g., pearlitic steels, ferritic steels, bainitic steels, and tempered
martensitic steels) from lower shelf, through transition and upper shelf behavior, including arrest
toughness behavior. Conversely, these temperature-dependent fracture toughness models should
not be applied to un-tempered martensitic steels, which have a body-centered tetragonal (BCT)
lattice structure, or to austenite, which has a face-centered cubic (FCC) structure.
The statistical model of cleavage fracture proposed by Wallin, Sarrio and Törrönen (WST) [21]
was based on an understanding of the weakest link nature of the fracture mechanisms describing

3-3
Fracture Toughness Models in CC N-830-1

cleavage fracture dating to the 1950s. The idea that cleavage fracture of ferritic steel occurs
when a critical tensile stress is exceeded at a critical particle evolves from the work of McMahon
and Cohen, Curry and Knott, and Smith, among others [31-37]. In the 1970s, Ritchie, Knott, and
Rice (RKR), and Curry and Knott incorporated these observations into models that predict,
respectively, how toughness changes with temperature, and the scatter of fracture toughness at a
single temperature [32-34, 38, 39]. The WST model begins with the notion, most commonly
attributed to RKR, that cleavage fracture will initiate and propagate to failure when a critical
opening mode stress is exceeded over some critical distance ahead of the crack tip. WST
combined an RKR-type model with Curry and Knott’s idea that cleavage fracture is controlled
by a “statistical competition between crack nuclei of varying sizes and frequencies in the rapidly
changing stress gradient ahead of a {sharp} crack tip” [38]. The most significant contribution of
the WST model is not the introduction of a new understanding of cleavage fracture, but rather
the important generalizations WST made concerning the cleavage fracture behavior of all ferritic
steels.

3.1.5 Empirical Basis


As discussed previously, the empirical basis of the MC temperature-dependence derivation was
based on data obtained from the ORNL HSST Irradiation Program [23]. Within the ORNL
program, there were two very large datasets containing both unirradiated and irradiated
toughness values for welds 72W and 73W. These very large datasets presented an excellent
opportunity to study the effects of irradiation embrittlement on the shape of the transition
fracture toughness curve with respect to temperature. Even when large shifts in toughness were
observed with irradiation, the shape of the temperature-dependence curves remained the same,
confirming the theory that the equation describing the temperature-dependence of transition
fracture toughness was common to all ferritic steels regardless of their irradiated condition.
In 1984, Wallin [20] demonstrated that the distribution of cleavage fracture toughness values at a
single temperature is well represented by a three-parameter Weibull distribution having two
parameters fixed: a minimum value (Kmin) of 20 MPa√m, and a shape parameter (b) of four.
Wallin showed that this distribution applies to ferritic materials that can be considered to have a
random distribution of cleavage initiation sites spread homogeneously throughout the material.
The only material dependent quantity needed to establish the distribution of cleavage fracture
toughness values at a single temperature is the third parameter of the Weibull distribution, the
location parameter, which Wallin called K0.
The Weibull shape parameter of four is theoretically based. It depends only on the assumption
of small scale yielding in the presence of a sharp crack and a homogeneous distribution of
potential cleavage initiators spread throughout the ferrite matrix. Wallin drew data from nine
literature sources (several dozen data points in all, including both RPV and non-RPV steels, both
base materials and welds) to provide empirical evidence that supported a Weibull shape
parameter of four [20].
Wallin observed that some experimental data sets were not represented well by a two-parameter
Weibull distribution with a shape parameter of four, despite the fact that the data was in good
conformance with the underlying assumptions [20]. Referencing the WST work [21], he argued
that the primary reason for this departure from theoretical expectation was that:

3-4
Fracture Toughness Models in CC N-830-1

… the two-parameter form of the Weibull distribution is used, which assumes no


limiting KIc value beneath which cleavage crack propagation becomes impossible.
The existence of a limiting value is, however, physically reasonable. This can
also be shown with the WST-model. The WST-model predicts a Kmin between 5
and 15 MPa√m below which crack propagation is impossible in more than one
grain, causing blunted microcracks.
In [22], Wallin used the same literature database described previously to demonstrate that a
reasonable value of Kmin lies between 10 and 20 MPa√m. In his 2011 textbook, Wallin states
[40]:
The assumption of Kmin = 10 MPa√m improves the compatibility between
experiments and theory, but the best compatibility is obtained when assuming Kmin
= 20 MPa√m. It was thus concluded that realistic values for Kmin, in the case of
normal structural steels would be of the order of 10-30 MPa√m. Since a reliable
experimental estimation of Kmin is not possible [from limited data] … in the
standard MC procedure it was adopted a constant value of Kmin = 20 MPa√m.
ASTM E1921 adopted the value of 20 MPa√m.
A consequence of the work in [20, 21] is that the effect of specimen size (i.e., crack front length)
on fracture toughness was found to scale with the ¼-power of thickness [41-43]. This derives
directly from the Weibull shape parameter of four, so it is a theoretical expectation dependent on
the same conditions as the shape parameter (i.e., the assumption of small scale yielding in the
presence of a sharp crack and a homogeneous distribution of potential cleavage initiators
throughout the ferrite matrix).

3.1.6 Model Validation


Extensive work has occurred since the late 1990s to validate the three fundamental aspects of the
MC:
• Distribution of KJc values following a three-parameter Weibull distribution having two
parameters fixed (shape parameter of four and a Kmin equal to 20 MPa√m).
• A “size,” or crack front length, effect having an exponent equal to the ¼-power of thickness.
• A temperature dependency having an exponential slope of 0.019.
Some of the more extensive efforts are listed below:
• In the late 1990s and early 2000s, the U.S. nuclear power industry undertook an extensive
effort to review the MC for potential use in the ASME Code. That effort resulted in the RTT0
Code Cases [16, 17, 44]. As part of this activity, an extensive empirical database was
compiled from the literature and was used to empirically evaluate the three fundamentals
aspects of the MC. This work, and closely related efforts (e.g., the Kewaunee plant
submittal), are reported in [45-49].
• In 2009, the NRC published its own evaluation, using a similar methodology to that used by
the industry, but expanded to include more data from the literature [24].

3-5
Fracture Toughness Models in CC N-830-1

• Between 2000 and 2005, a group of laboratories working under European Commission
funding performed extensive KJc characterization of a single RPV-grade forging to validate
the MC. Key papers from this work include [50, 51].
• In the early 2000s, the NRC sponsored a study at the University of California at Santa
Barbara focused specifically on the size-effect aspect of the MC. This study featured an
extensive KJc characterization using ex-vessel (Shoreham) materials [52, 53].
Across the board these efforts found no substantial deviations from the Master Curve model as
originally proposed by Wallin [20-22] and as represented within E1921.
Over the past 15 years, various publications have suggested possible further refinements of the
MC concept that are useful in specific situations. These include the following:
• On temperature dependence, References [54] and [55] provide information on how the
exponential slope of 0.019 is affected by various factors. There is considerable dispersion in
the data, but even so, a tendency to a reduction of the value of 0.019 with increasing
embrittlement can be seen. Reference [55] provides a formula to estimate this effect.
• On Kmin, a method proposed in Reference [56] enables estimation of a data-set specific value
for Kmin. However, large amounts of KJc data are needed to use the procedure, and the value
of Kmin of 20 MPa√m is still seen as a practicable estimate.
• Methods have been developed when it is suspect that a particular KJc dataset may not have
been obtained from a homogeneous population of cleavage crack initiators. These methods
are particularly useful in application to T0 values measured on welds [57].
• Some empirical evidence demonstrates that, as embrittlement occurs and the crack initiation
and crack arrest distributions merge, the constant Weibull shape parameter of four becomes
less accurate. EricksonKirk and co-workers [58, 59] postulated that the cause of this
behavior might be stable micro-arrests in high embrittlement materials.
All the foregoing reports were reviewed and form the basis for the limits on the applicability of
CC-N-830-1, as appropriate. As additional information comes to light, more refined models may
be developed for implementation into a subsequent revision of CC N-830-1.

3.1.7 Limits of Applicability


Based on the development and validation work described above, the following limitations for
using the KJc model are contained in CC N-830-1:
• The MC KJc model cannot be applied to non-ferritic steels (e.g., not to hardened martensite,
or austenitic steels).
• The MC KJc model may not be applicable to specimens loaded to very high strain rates. At
higher strain rates, there may be a slight effect of increasing strain rate that results in an
increase in the slope of the transition toughness curve such that toughness becomes less
sensitive to temperature. But even at the rates included in the EPRI database, the effect is
very slight [11].

3-6
Fracture Toughness Models in CC N-830-1

• The MC KJc model cannot be used at temperatures above which crack extension occurs by
dislocation motion, void initiation growth, and coalescence (i.e., upper limit of applicability
of the MC).
• The MC KJc model may not be applicable at temperatures below which deformation occurs
predominantly by twinning (T0-160oC).

3.2 Cleavage Crack Arrest Fracture Toughness, KIa


Wallin et al. [60] also developed a model for the temperature dependence and distribution of
crack arrest fracture toughness, KIa. Consistent with the MC treatment, Wallin selected the
temperature at which the mean measured KIa value is 100 MPa√m for use as the temperature to
index data from different heats of steel together to form a single crack arrest transition curve.
Wallin called this index temperature TKIa [60].

3.2.1 Description of Model


In questioning the ASME-specified separation between KIc and KIa, Wallin, et al. [60] analyzed
nine sets of KIa data using a method similar to that used in developing the KJc MC. The authors
started with the assumption that the KIa data followed the same temperature-dependence as was
observed for the KJc MC (Eqn. (3-3)), and defined TKIa as the temperature at which the mean KIa
value equals 100 MPa√m. They further reasoned that because crack arrest is not a weakest-link
mechanism, there shouldn’t be a size effect since the scatter in data is controlled by the matrix
properties and not the distribution of crack initiating particles. Using TKIa to normalize the data
from the nine datasets, Wallin confirmed that the assumed temperature-dependence fit the data
well.

3.2.2 Basic Form


The mean temperature dependence of KIa follows the form given for the initiation MC, and is
given by [60]:
𝑲𝑲𝒎𝒎𝒎𝒎𝒎𝒎𝒎𝒎
𝑰𝑰𝑰𝑰 = 𝟑𝟑𝟑𝟑 + 𝟕𝟕𝟕𝟕 ⋅ 𝒆𝒆𝒆𝒆𝒆𝒆[𝟎𝟎. 𝟎𝟎𝟎𝟎𝟎𝟎(𝑻𝑻 − 𝑻𝑻𝑲𝑲𝑲𝑲𝑲𝑲 )] Eq. 3-5

where TKIa is defined relative to T0 using:


𝑻𝑻𝑲𝑲𝑲𝑲𝑲𝑲 = 𝑻𝑻𝟎𝟎 + 𝟒𝟒𝟒𝟒. 𝟗𝟗𝟗𝟗𝟗𝟗𝟗𝟗𝟗𝟗[−𝟎𝟎. 𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝑻𝑻𝟎𝟎 ] Eq. 3-6

Assessment of additional data combined with the nine datasets used by Wallin further confirmed
the temperature dependence described by Eqn. (3-5) [61, 62]. Moreover, the data suggest that,
similar to the initiation MC, the temperature dependence of KIa is not affected strongly by
irradiation [62].

3.2.3 Distribution
Wallin observed that the scatter in KIa data was less than that observed for KJc. He assumed a
log-normal distribution so that the proportional scatter in KIa was constant, matching the
empirical evidence. A log normal distribution with a variance equal to 18% of the mean value
was found to match the data well [60]. Using Wallin’s log normal distribution the crack arrest
toughness curve at percentile, p, is defined as:

3-7
Fracture Toughness Models in CC N-830-1

𝒑𝒑
for lower bound curves: 𝑲𝑲𝑰𝑰𝑰𝑰 = 𝑲𝑲𝒎𝒎𝒎𝒎𝒎𝒎𝒎𝒎
𝑰𝑰𝑰𝑰 �𝟏𝟏 − 𝟎𝟎. 𝟏𝟏𝟏𝟏𝑴𝑴𝒑𝒑 � Eq. 3-7

(𝟏𝟏−𝒑𝒑)
𝑲𝑲𝑰𝑰𝑰𝑰 = 𝑲𝑲𝒎𝒎𝒎𝒎𝒎𝒎𝒎𝒎 �𝟏𝟏 + 𝟎𝟎. 𝟏𝟏𝟏𝟏𝑴𝑴𝒑𝒑 � Eq. 3-8
for upper bound curves: 𝑰𝑰𝑰𝑰

Using Eqn. (3-7), a value for p of 0.05 will produce a 5% lower bound curve. A 95% upper
bound curve is similarly defined using Eqn. (3-8). When using either of these equations, the
value of p cannot exceed 0.5 (0 < p < 0.5). There is no effect of component thickness, crack
front length, or product form on Eqns. (3-7) and (3-8).

3.2.4 Theoretical Basis


In 2002, Kirk, et al [61] presented a physically-based mechanism for crack arrest to support the
crack arrest toughness model of ferritic steels developed by Wallin [60]. They present a detailed
discussion based in dislocation mechanics that demonstrates that the empirical trends observed
by Wallin are anticipated physically.
The mechanism controlling the temperature-dependence of crack arrest toughness is the same as
that controlling the temperature dependence of initiation toughness, i.e., dislocation motion
through the matrix to absorb the applied energy. Crack initiation occurs when energy absorption
by dislocation motion can no longer occur and energy is then available to drive crack initiation
and growth. Crack arrest occurs when dislocations once again become mobile in a material and
can move to the crack tip to absorb the energy of the propagating crack. Because of this, both
KIc and KIa data are expected to exhibit the same temperature dependence. This temperature
dependence results from the temperature dependence of the Peierls-Nabarro stresses required to
move dislocations through the ferritic matrix. The temperature dependence of both toughness
values is controlled by the atomic arrangement, or crystal structure of the material.
Consequently, the temperature dependence of KIc and KIa is expected to be common to all ferritic
steels.
Wallin observed a significant reduction in the scatter inherent in KIa data over that observed in
KIc data. The cause of this reduction is based on the difference in the distribution of dislocation-
trapping barriers that affect each property. Crack initiation occurs when dislocations accumulate
at non-coherent particles (i.e., carbides, grain boundaries, twin boundaries, etc.) and produce
enough strain to elevate the local stress at the barrier high enough to fracture the barrier or cause
its decohesion from the matrix. These non-coherent particles are large with respect to size and
inter-particle spacing relative to the dislocation trapping sites responsible for crack arrest
toughness. The non-coherent particles responsible for initiation are of sub-micron size (i.e., 1/10
micron), and their spacing is on the same order. The dislocation-trapping defects responsible for
crack arrest (i.e., vacancy clusters, interstitial clusters, coherent and semi-coherent particles, and
other dislocations) are of a much smaller size (nanometer) and have a spacing on the same scale.
The possible variation in local stress state over the microstructural distances that control crack
arrest is also much smaller than that possible over the microstructural distances that control crack
initiation. This is due to the high strain rate in front of the moving crack tip, which constrains
the development of a large strain field.
The considerably smaller size and spacing of the defects responsible for crack arrest, relative to
those responsible for crack initiation, also suggests that crack arrest toughness should not be
greatly influenced by the length of the crack front for all crack front lengths of practical concern

3-8
Fracture Toughness Models in CC N-830-1

in RPV applications. The size effect observed in crack initiation toughness is due to the weakest
link nature of the initiation event, but crack arrest is not a weakest-link phenomena.

3.2.5 Model Validation


After the initial development of the KIa curve, Wallin continued his analysis with 53 sets of KIa
data. Most of the data consisted of RPV steels, including plates, forgings and welds, both
irradiated and unirradiated, but other steels (non-RPV) were included as well. The data sets
covered a wide range of yield strengths (280 MPa to 1,082 MPa). To ensure that TKIa for each
dataset could readily be obtained, it was desired that each dataset contained at least ten
specimens [60], although several datasets contained fewer specimens.
TKIa was determined for each dataset and used to normalize the datasets. The normalized data
were then compared to the temperature dependence and log normal distribution assumed in the
initial model development, and were found to match well, verifying the initial assumptions. The
data were then used to develop a correlation between T0 and TKIa, as described in Chapter 4.
Other validation efforts have subsequently been conducted, including that of Hein, et al. [62] in
the CARINA and CARISMA projects in which RPV materials consisting of six base materials
(plates and forgings), seven welds, and weld heat affected zone (HAZ) materials were tested.
The data, including both KJc and KIa, were used to confirm the efficacy of the MC approach for
assessing German nuclear power plant safety. The materials were tested in both the irradiated
and unirradiated conditions to assess the effects of embrittlement on the characterization of
toughness behavior. Compact crack arrest and duplex crack arrest specimens were tested
following the ASTM E1221-10 standard test method [63], and KIa values used to determine TKIa
for each data set. The results of this program showed that the KIa master curve, with the
temperature dependence described by Eqn. (3-5) and log normal distribution defined by Eqns.
(3-7) and (3-8), well represented all the KIa data sets tested in the CARINA project.

3.2.6 Limits of Applicability


Based on the development and validation work described above, the following limitations for
using the KIa model are contained in CC N-830-1:
• The KIa model cannot be applied to non-ferritic steels (e.g., not to hardened martensite, or
austenitic steels).
• The KIa model cannot be used at temperatures above which crack extension occurs by
dislocation motion, void initiation growth, and coalescence (i.e., the upper limit of
applicability of the MC).
• The KIa model may not be applicable at temperatures below which deformation occurs
predominantly by twinning (T0-160oC), but it is believed that cracks arrest does not occur at
such low temperatures.

3.3 Ductile Crack Initiation Fracture Toughness, JIc


To identify the upper limit of applicability of the KJc MC, EricksonKirk et al. [11, 62] developed
a model describing the temperature dependence of upper shelf fracture toughness (JIc) that
pertained to all ferritic steels. This upper shelf model was used to define the temperature region

3-9
Fracture Toughness Models in CC N-830-1

in which fracture transitioned from cleavage-dominated to ductile-crack-extension-dominated,


thereby defining a point beyond which the MC should no longer be used.
Both ductile fracture toughness and flow stress measure the ability of a material to absorb energy
by dislocation motion. This similarity in mechanism provided EricksonKirk the justification to
look to the temperature dependence of the flow stress to define the temperature dependence of
the upper shelf fracture toughness. The JIc model development was based on the Zerilli-
Armstrong (Z-A) constitutive equation describing the temperature dependence of the flow stress
that was common to ferritic steels [27]. The equation describing the temperature dependence of
JIc was found to be a simple scalar multiple of the temperature dependence predicted by Z-A for
flow stress. By assuming this temperature-dependence, individual datasets of JIc data versus test
temperature were fit, and then each dataset was normalized by the mean JIc at a single
temperature. The JIc at 288o C was selected as the reference value to use in normalizing the data
for model development [11, 64].

3.3.1 Description of Model


Similar to the Wallin MC, the JIc equation has two components; a temperature dependence
common to all ferritic steels, and a distribution that defines the expected scatter in JIc at any
given temperature.

3.3.2 Basic Form


The equation describing the temperature dependence of the mean value of JIc, is defined in CC
N-830-1 as:

𝑱𝑱𝒎𝒎𝒎𝒎𝒎𝒎𝒎𝒎
𝑰𝑰𝑰𝑰 = 𝟏𝟏. 𝟕𝟕𝟕𝟕{𝟏𝟏𝟏𝟏𝟏𝟏𝟏𝟏 ∙ 𝐞𝐞𝐞𝐞𝐞𝐞[−𝟎𝟎. 𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎(𝑻𝑻 + 𝟐𝟐𝟕𝟕𝟑𝟑. 𝟏𝟏𝟏𝟏)] − 𝟑𝟑. 𝟑𝟑𝟑𝟑𝟑𝟑} + 𝑱𝑱𝒄𝒄(𝑼𝑼𝑼𝑼) − 𝚫𝚫𝑱𝑱𝑰𝑰𝑰𝑰(𝑼𝑼𝑼𝑼) Eq. 3-9(a)

Where T is the temperature in oC and the reference JIc value is taken at 288o C. Jc(US) and ∆JIc(US)
are given by:
𝟏𝟏−𝝊𝝊𝟐𝟐
𝑱𝑱𝒄𝒄(𝑼𝑼𝑼𝑼) = {𝟑𝟑𝟑𝟑 + 𝟕𝟕𝟕𝟕 × 𝐞𝐞𝐞𝐞𝐞𝐞[𝟎𝟎. 𝟎𝟎𝟎𝟎𝟎𝟎(𝟒𝟒𝟒𝟒. 𝟖𝟖𝟖𝟖𝟖𝟖 − 𝟎𝟎. 𝟐𝟐𝟐𝟐𝟐𝟐𝟐𝟐𝑻𝑻𝒐𝒐 )]}𝟐𝟐 Eq. 3-9(b)
𝑬𝑬𝑼𝑼𝑼𝑼

𝚫𝚫𝑱𝑱𝑰𝑰𝑰𝑰(𝑼𝑼𝑼𝑼) = 𝟏𝟏. 𝟕𝟕𝟕𝟕{𝟏𝟏𝟏𝟏𝟏𝟏𝟏𝟏 ∙ 𝐞𝐞𝐞𝐞𝐞𝐞[−𝟎𝟎. 𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎(𝑻𝑻𝑼𝑼𝑼𝑼 + 𝟐𝟐𝟐𝟐𝟐𝟐. 𝟏𝟏𝟏𝟏)] − 𝟑𝟑. 𝟑𝟑𝟑𝟑𝟑𝟑} Eq. 3-9(c)

{𝟐𝟐𝟐𝟐𝟐𝟐𝟐𝟐𝟐𝟐𝟐𝟐−𝟕𝟕𝟕𝟕.𝟒𝟒𝑻𝑻𝑼𝑼𝑼𝑼 }
𝑬𝑬𝑼𝑼𝑼𝑼 = Eq. 3-9(d)
𝟏𝟏𝟏𝟏𝟏𝟏𝟏𝟏

𝑻𝑻𝑼𝑼𝑼𝑼 = 𝟒𝟒𝟒𝟒. 𝟖𝟖𝟖𝟖𝟖𝟖 + 𝟎𝟎. 𝟕𝟕𝟕𝟕𝟕𝟕𝟕𝟕𝑻𝑻𝒐𝒐 Eq. 3-9(e)

There is no effect of component thickness or crack front length on Eqn. (3-9.).

3.3.3 Distribution
The distribution on JIc is a function of both temperature and prior hardening, as defined by the
mean value of JIc at 288o C. Based on the work presented by Kirk, et al. [65], the standard
deviation for JIc, σΔJIc , is defined in CC N-830-1 as:

𝝈𝝈𝜟𝜟𝜟𝜟𝜟𝜟𝜟𝜟 = 𝑨𝑨 ⋅ 𝒆𝒆𝒆𝒆𝒆𝒆[𝑩𝑩(𝑻𝑻 − 𝟐𝟐𝟐𝟐𝟐𝟐)] Eq. 3-10(a)

3-10
Fracture Toughness Models in CC N-830-1

𝑨𝑨 = 𝟗𝟗. 𝟎𝟎𝟎𝟎 ⋅ 𝒆𝒆𝒆𝒆𝒆𝒆(𝟏𝟏. 𝟏𝟏𝟏𝟏 ⋅ 𝑷𝑷) Eq. 3-10(b)

𝑩𝑩 = 𝐌𝐌𝐌𝐌𝐌𝐌{𝟎𝟎, (𝟎𝟎. 𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎 ⋅ 𝑷𝑷 − 𝟎𝟎. 𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎)} Eq. 3-10(c)

𝑷𝑷 = 𝐌𝐌𝐌𝐌𝐌𝐌{𝟏𝟏, 𝐌𝐌𝐌𝐌𝐌𝐌[𝟎𝟎, 𝐌𝐌𝐌𝐌𝐌𝐌(𝑷𝑷𝟏𝟏 , 𝑷𝑷𝟐𝟐 )]} Eq. 3-10(d)

𝑱𝑱𝑰𝑰𝑰𝑰(𝟐𝟐𝟐𝟐𝟐𝟐)
𝑷𝑷𝟏𝟏 = − 𝟎𝟎. 𝟒𝟒𝟒𝟒 Eq. 3-10(e)
𝟏𝟏𝟏𝟏𝟏𝟏

𝑱𝑱𝑰𝑰𝑰𝑰(𝟐𝟐𝟐𝟐𝟐𝟐)
𝑷𝑷𝟐𝟐 = + 𝟎𝟎. 𝟓𝟓𝟓𝟓 Eq. 3-10(f)
𝟖𝟖𝟖𝟖𝟖𝟖

The ductile crack initiation toughness curve at percentile, p, or (1-p), is defined as follows:
𝒑𝒑
𝑱𝑱𝑰𝑰𝑰𝑰 = 𝑱𝑱𝒎𝒎𝒎𝒎𝒎𝒎𝒎𝒎 − 𝝈𝝈𝚫𝚫𝑱𝑱𝑰𝑰𝑰𝑰 𝑴𝑴𝒑𝒑 Eq. 3-11(a)
for lower bound curves 𝑰𝑰𝑰𝑰

(𝟏𝟏−𝒑𝒑)
𝑱𝑱𝑰𝑰𝑰𝑰 = 𝑱𝑱𝒎𝒎𝒎𝒎𝒎𝒎𝒎𝒎 + 𝝈𝝈𝚫𝚫𝑱𝑱𝑰𝑰𝑰𝑰 𝑴𝑴𝒑𝒑 Eq. 3-11(b)
for upper bound curves: 𝑰𝑰𝑰𝑰

As an example, a value for p of 0.05 would produce a 5% lower bound curve using Eqn. (3-11a),
and a 95% upper bound curve using Eqn. (3-11b). When using Eqns. (3-11a) and (3-11b), the
value of p should not exceed 0.5 (0 < p < 0.5).

3.3.4 Theoretical Basis


In developing the model for upper shelf fracture toughness behavior, EricksonKirk [11, 64]
assumed that, since the upper shelf fracture toughness mechanism was dislocation-controlled
(void nucleation, growth and coalescence), it should follow the same, or similar, temperature
dependence as the flow stress. Starting with the Zerilli-Armstrong constitutive equation for the
flow stress in ARMCO iron [27], the Z-A equation for BCC-material flow stress is given by:
−𝟏𝟏� 𝟏𝟏�
𝝈𝝈𝒁𝒁𝒁𝒁 = 〈𝝈𝝈𝑮𝑮 + 𝒌𝒌𝒅𝒅 𝟐𝟐 + 𝑲𝑲𝜺𝜺 𝟐𝟐 〉 + �𝑩𝑩𝟎𝟎 𝒆𝒆−(𝜷𝜷𝟎𝟎 −𝜷𝜷𝟏𝟏 𝒍𝒍𝒍𝒍𝜺𝜺̇ )𝑻𝑻 � Eq. 3-12

where:
σG = the prior hardening term
kd-1/2 = the Hall-Petch grain boundary hardening term
Kε1/2 = the strain hardening term
B0 and β0 = material constants
𝛽𝛽1 𝑙𝑙𝑙𝑙𝜖𝜖̇ = term that accounts for dynamic strain aging in BCC materials
T = temperature (K)
All three of the terms in the <brackets> are athermal terms, and the terms in [square brackets] are
the thermal terms.
Only the thermal terms were used in characterizing the upper shelf fracture behavior. JIc is
characterized using quasi-static test rates, so the dynamic strain aging term was set to a constant

3-11
Fracture Toughness Models in CC N-830-1

using 𝜀𝜀̇ equal to 0.0004/sec. To develop an empirical fit to identify the constants, the data was
characterized relative to a reference JIc value, arbitrarily chosen as the JIc at 288o C, and
compared to the difference in flow stress between that predicted by the Z-A equation and the
flow stress measured at a reference temperature:

∆𝑱𝑱𝑰𝑰𝑰𝑰 ≡ 𝑱𝑱𝑰𝑰𝑰𝑰 (𝑻𝑻) − 𝑱𝑱𝑰𝑰𝑰𝑰(𝟐𝟐𝟐𝟐𝟐𝟐) ≡ 𝜶𝜶�∆𝝈𝝈𝒇𝒇𝒍𝒍𝒍𝒍𝒍𝒍 � Eq. 3-13

Where α has units of mm to convert from stress units (MPa) inside the {brackets} to J units of
kJ/m2.
Substituting the Z-A thermal flow equation for ∆𝜎𝜎𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓 , and taking the reference temperature as
288oC for both JIc and σflow, gives the temperature-dependence equation for JIc as:

∆𝑱𝑱𝑰𝑰𝑰𝑰 ≡ 𝑱𝑱𝑰𝑰𝑰𝑰 (𝑻𝑻) − 𝑱𝑱𝑰𝑰𝑰𝑰(𝟐𝟐𝟐𝟐𝟐𝟐) = 𝜶𝜶�𝑩𝑩𝟎𝟎 𝒆𝒆𝒆𝒆𝒆𝒆[−𝜷𝜷𝟎𝟎 𝑻𝑻 + 𝜷𝜷𝟏𝟏 𝑻𝑻𝑻𝑻𝑻𝑻(𝜺𝜺̇ )] − 𝝈𝝈𝒇𝒇𝒇𝒇𝒇𝒇𝒇𝒇(𝟐𝟐𝟐𝟐𝟐𝟐) � Eq. 3-14

This Z-A dislocation mechanics-based equation was used as the basis for the empirical fitting
performed to develop the JIc versus temperature equation as described in detail in References
[11, 64]. The empirical basis for the model is described in the following section of this report.
The expected scatter in the JIc data at a given temperature can be understood in the context of
dislocation mechanics. Ductile fracture occurs by accumulation of dislocations at defects in the
material upon loading. When a critical dislocation density is reached, voids initiate, followed by
growth of the void and coalescence with other voids to form a crack. The energy absorbed by
dislocation motion leading to ductile crack initiation in a specimen, and therefore the energy
defining the upper shelf fracture toughness, is controlled by the strain history and defect density
of the material prior to loading, as well as the ease with which dislocations move within the
material. Dislocation motion is controlled by the temperature-dependent Peierls-Nabarro stress,
which in turn is controlled by the short-range obstacles to dislocation motion provided by the
lattice atoms (BCC crystal structure in the case of ferritic steels). The total energy absorbed
prior to crack initiation is controlled by the long-range barriers to dislocation motion, i.e., the
defects in the material (vacancies/interstitials, other dislocations and precipitates/particles) [11].
The higher the initial defect density in the material, the less energy is absorbed by movement of
dislocations upon loading before the critical dislocation density is obtained for void initiation and
growth. As a result, materials that have a higher yield strength and hardness have a lower JIc at
any given test temperature and show less scatter in the data compared to specimens that have a
lower yield strength and hardness. The microstructural features controlling hardness and yield
strength will not affect the temperature-dependence of dislocation motion for the reasons stated
in the KJc Theoretical Basis Section, but they will affect the scatter in the JIc data at any given
temperature. Based on this understanding of the dislocation-based, ductile fracture process, it is
expected that scatter in JIc data varies with temperature and the mean value of JIc at any given
temperature.

3.3.5 Empirical Basis


A database of over 1,000 JIc values was compiled from the literature [64] and used to develop an
empirical fit to the theoretically-derived model form described by Eqn. (3-14). The data
consisted primarily of RPV steels (ASTM A533B and A508 and their welds) in the unirradiated
and irradiated conditions, as well as some high strength, low alloy ferritic steels used by the U.S.

3-12
Fracture Toughness Models in CC N-830-1

Navy. The data were filtered to ensure that only data obtained at quasi-static loading rates were
used, that each dataset had at least five JIc values, and that each dataset contained JIc values
obtained from at least two different temperatures. This filtering resulted in a total of 809 JIc
values that were appropriate to use in the development of the upper shelf fracture toughness
model.
Starting with Eqn. (3-14), an iterative process was used to determine the constants in Eqn. (3-14)
[64]. A reference temperature, Tref, was arbitrarily set to 288oC, as there was a significant
amount of data at this test temperature. Both ∆σflow and ∆JIc are zero at Tref, resulting in σT (the
value of the thermal part of the flow stress at 288 oC) equal to 3.3 MPa using starting values for
the constants in the Z-A equation for ferritic steels of B0 = 1,000 MPa, β0 = 0.0074 K-1, and
β1 = 0.004 K-1 [66], and taking α to be 2.1 mm based on a preliminary analysis of a small
dataset. The equation was fit to individual datasets using a least squares method, thereby
establishing a value of JIc at 288oC for every data set. The value of α was then adjusted to
minimize the sum of squares residuals between each measured ∆JIc value (∆JIc = JIc(T) – JIc(288))
and each predicted value of ∆JIc (∆JIc=α{B0exp[-β0T+β1Tln𝜀𝜀̇]-σflow(288)}), which required 15
iterations to converge. Once α was minimized for all datasets, the 809 values were considered
together, and a final iteration was performed to adjust the constants to define a best estimate
model to represent all ferritic steels. Inserting the constants determined from the least squares
fitting method into Eqn. (3-14) gives the temperature dependence of the JIc MC:
𝒌𝒌𝒌𝒌
∆𝑱𝑱𝑰𝑰𝑰𝑰 ≡ 𝑱𝑱𝑰𝑰𝑰𝑰 (𝑻𝑻) − 𝑱𝑱𝑰𝑰𝑰𝑰(𝟐𝟐𝟐𝟐𝟐𝟐) = 𝟏𝟏. 𝟕𝟕𝟕𝟕{𝟏𝟏𝟏𝟏𝟏𝟏𝟏𝟏𝟏𝟏𝟏𝟏𝟏𝟏[−𝟎𝟎. 𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎 + 𝟎𝟎. 𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎(𝜺𝜺̇ )] − 𝟑𝟑. 𝟑𝟑} � � 𝟐𝟐 � Eq. 3-15
𝒎𝒎

The original work to develop Eqn. (3-15) did not consider characterization of the scatter in the
data at any one temperature [11]. Follow-on work by Kirk, et al. [63] using the same data
demonstrated the scatter in ∆JIc to be temperature-dependent, with the standard deviation given
by:

σ ∆J = 51.2 ⋅ e −0.0056T {σ∆JIc in kJ/m2, temperature in °C}


Ic
Eq. 3-16

3.3.6 Model Validation


The JIc models given by Eqns. (3-15) and (3-16) well represent both the temperature dependence
and the scatter of the JIc data to which they were fit. While the findings from [63] are promising,
questions remained concerning the possible influence of the ∆JIc normalization procedure on the
outcome of the analysis. A more detailed analysis of the data used in the original model
development was performed to better calibrate the fitting parameters, and this recalibrated JIc
model was used to assess the appropriateness of the models in describing the behavior of a more
recent, larger set of JIc data from a RPV forging (ASTM A508 Class 3 steel called the EURO
Forge) [65]. The EURO Forge is a well-characterized material that was extensively tested and
documented [67-71]. The data in this validation and recalibration effort included 45 JIc data
points (defined using ASTM E1820-05 [72]) obtained at test temperatures ranging between -20
and +288oC, with 6 to 8 specimens tested at each temperature to quantify the scatter.
Figure 3-1 shows the EURO Forge data relative to the JIc temperature dependence and scatter
model described in Eqns. (3-15) and (3-16). Except for the dataset at 175oC, the models describe
the data reasonably well. Further assessment of the specimen tested at 175oC provided no

3-13
Fracture Toughness Models in CC N-830-1

explanations for this outlier data set. A more quantitative assessment of the scatter in the EURO
Forge specimens was performed to confirm the observations made in plotting the data relative to
the model predictions. The outcome of that assessment was a finding that the scatter in the
EURO Forge data was higher than predicted by the model, and that the scatter had an athermal
component that had not been observed previously. This athermal component of the scatter was
found to scale in proportion to the mean upper shelf toughness of the dataset, i.e., higher
toughness materials (such as the EURO Forge material) exhibited higher scatter than did lower
toughness materials [65]. This aspect of the scatter may not have been observed in the upper
shelf model development study due to the limited amount of JIc values at higher temperatures in
the original database.

Figure 3-1
Comparison of the temperature dependence exhibited by the JIc data for the EURO Forge
with the model proposed in [65] (i.e., Eqn. (3-15) with uncertainty bounds based on Eqn
(3-16)).

A detailed quantitative assessment of the scatter in the EURO Forge data, combined with the
data used in the original model development, was performed relative to both the temperature and
mean value of JIc for each dataset [65]. The TUS model [11] (described in Chapter 4) was used to
censor data at low upper shelf temperatures that may have cleavage components to the fracture
mode, and the JIc data was divided into five bins based on the percentile of the total distribution
of JIc(288) for all datasets the average JIc(288) value for each dataset fell into. Datasets for which
the JIc(288) value fell between the zero and 20th percentiles were placed into the first bin, between
the 20th and 40th percentiles were placed in the second bin, and so on (the JIc(288) value of 283
kJ/m2 of the EURO Forge placed it at the 83rd percentile or the fifth bin). The standard deviation
for each bin was calculated, and this showed a trend of increasing standard deviation with
increasing JIc(288). Based on this analysis, a fit of the form:

3-14
Fracture Toughness Models in CC N-830-1

σ ∆J = A ⋅ e (B⋅T )
ˆ
Ic
Eq. 3-17

was identified to represent the standard deviation in JIc. In this equation, A and B are fitting
parameters, and 𝑇𝑇� is the test temperature expressed relative to 288oC, which was the reference
temperature used to define the bins. Fitting this equation to the data in each of the five bins
resulted in definitions of the A and B as shown in Eqn. (3-10).
The distribution model described by Eqn. (3-10) was developed using all the available JIc data
(the original 91 datasets and the additional EURO Forge 45 JIc values). In Figure 3-2, the JIc
temperature dependence model from Eqn. (3-9) is combined with the JIc scatter model of Eqn.
(3-10) and compared to three data sets having considerably different upper shelf toughness
levels. This comparison demonstrates the ability of the combined model from Eqns. (3-9) and
(3-10) to represent the temperature dependence and scatter of a wide range of toughness
conditions. Further validation efforts await additional JIc data.

3.3.7 Limits of Applicability


Based on the development and validation work described above, the following limitations for
using the JIc model are contained in CC N-830-1:
• The use of the models described in Eqns. (3-9) through (3-11) is limited to the ranges of
data used in their development and validation. As such, these models should only be used to
predict fracture toughness on the upper shelf and should not be applied at temperatures for
which cleavage fracture is the expected fracture mode.
• The JIc, upper shelf model only pertains to ferritic steel materials and should not be used to
predict upper shelf behavior of non-ferritic materials (e.g., martensite, austenite or other
alloys).
• Data was not available at temperatures above 300oC, so extrapolation of the model to higher
temperatures is questionable.
• This model is not intended to predict ductile fracture toughness at high (dynamic) strain
rates, nor should it be used in temperature and strain rate conditions where dynamic strain
aging occurs as these may change the temperature dependence of upper shelf toughness for
the material.

3-15
Fracture Toughness Models in CC N-830-1

Figure 3-2
Comparison of the revised JIc model, Eqns. (3-9) and (3-10), with JIc data from steels
having three different upper shelf toughness (JIc(288)) levels [65].

3-16
4
FRACTURE TOUGHNESS LINKAGE MODELS IN CC N-
830-1

While the models summarized in Chapter 3 describe the temperature dependence and scatter
inherent to different measures of the fracture toughness of ferritic steels, they do not describe the
interrelationship of the different toughness measures. While general relationships have long
been recognized (e.g., steels with low transition temperatures tend to have high toughness on the
upper shelf), it is only in the last 10 to 15 years that systematic trends common to all ferritic
steels have been noted.
In this chapter, three models that link the KJc, KIa, and JIc toughness models described in Chapter
3 are summarized. These linkage models include the relationship between cleavage crack
initiation (KJc) and upper shelf (JIc) fracture toughness data, the relationship between cleavage
crack initiation (KJc) and cleavage crack arrest (KIa) fracture toughness data, and the relationship
between upper shelf (JIc) crack initiation and upper shelf crack growth (J-R) fracture toughness
data. These three linkage models, taken together, provide the relationships that link all
toughness values to a single reference value, T0.

4.1 The Relationship Between Cleavage Crack Initiation (KJc) and Upper
Shelf (JIc); TUS
Conventionally, the transition fracture toughness and upper shelf fracture toughness of ferritic
steels have been viewed as either separate properties or as properties between which only
general/qualitative relationships exist. Information presented in 2004 by EricksonKirk [11] and
further developed in 2006 by EricksonKirk, et al. [14] demonstrated the opposite to be true. The
EricksonKirk studies [11, 14] showed transition fracture toughness and upper shelf fracture
toughness to be directly related because the microstructural features responsible for both the
temperature dependence of fracture toughness and for the magnitude of fracture toughness at any
given temperature are the same in both transition and on the upper shelf. Data from several
dozen steels demonstrated a consistent linear relationship between T0 and the temperature at
which KJc (converted to Jc) and JIc are equal, termed TUS, over a range of T0 values exceeding
300°C (-180°C < T0 < +140°C).

4.1.1 Mathematical Form of the Model


Figure 4-1 shows the data from Reference [14]. These define the following relationship between
T0 and TUS:

𝑻𝑻𝑼𝑼𝑼𝑼 = 𝟒𝟒𝟒𝟒. 𝟖𝟖𝟖𝟖𝟖𝟖 + 𝟎𝟎. 𝟕𝟕𝟕𝟕𝟕𝟕𝟕𝟕𝑻𝑻𝟎𝟎 Eq. 4-1

Here both T0 and TUS are expressed in degrees Celsius.

4-1
Fracture Toughness Linkage Models in CC N-830-1

200

150 TUS = 0.7985To + 48.843


R2 = 0.9812
100

TUS [ C] 50
o

Weld
0 Plate
Forging
-50 HSLA
Mild Steel
-100 All
Linear (All)

-150
-200 -150 -100 -50 0 50 100 150 200

T o [oC]

Figure 4-1
Relationship between TUS and T0 [11, 14]

4.1.2 Physical Basis


EricksonKirk, et al. performed their studies motivated by the observation of “master curves” that
consistently describe the temperature dependence and scatter of fracture toughness in both
fracture mode transition [20, 21] and on the ductile upper shelf [11, 64]. They noted that the
interatomic spacing in the BCC ferritic steel matrix was responsible for the temperature
dependence of both the transition and upper shelf fracture toughness, and the size and
distribution of barriers to dislocation motion (second phase particles, point defects and other
dislocations) control the magnitude of fracture toughness at any given temperature for both
transition and upper shelf behavior. For example, steels exhibiting a low T0 (high KJc at any
given temperature) also exhibit a high JIc on the upper shelf because these materials exhibit a
fine/homogeneous distribution of defects (that is, many small, second phase particles, grain
boundaries, etc.). Such steels can accommodate considerable uniform dislocation motion prior
to sufficient accumulation of dislocations at any defect, which results in either cracking of the
second-phase particle (in transition) or decohesion of the second phase particle and subsequent
void growth (on the upper shelf). Because considerable uniform dislocation motion is possible,
higher toughness values occur at any temperature whether in transition or on the upper shelf.
Conversely, steels that have a high T0 also exhibit low upper shelf JIc values because these steels
contain a larger, less homogeneous distribution of defects (that is, larger grain sizes, larger
second-phase particles, higher original dislocation density, etc.). Such steels cannot
accommodate as much uniform dislocation motion prior to sufficient accumulation of
dislocations at any defect, which results in either cracking of the second-phase particle (in
transition) or the decohesion of the second phase particle and subsequent void growth (on the
upper shelf). Higher initial defect density results in less energy absorbed by dislocation motion
upon load application prior to accumulation of sufficient strain to cause fracture at any given
temperature, whether it is in the transition region or on the upper shelf.

4-2
Fracture Toughness Linkage Models in CC N-830-1

4.1.3 Empirical Basis


The TUS model is supported by data from 38 ferritic steels for which sufficient KJc and JIc data
were available for defining, respectively, the KJc MC and the upper shelf, JIc, master curve.
These materials are predominately nuclear grade base and weld metals that were tested in both
the unirradiated and irradiated conditions. Additionally, some data for mild steels and for a
copper precipitation hardened high-strength low alloy (HSLA) steel supported the model. A
total of 47 data sets existed for different material and irradiation conditions [14].
The model was developed as illustrated in Figure 4-2. The KJc model was used to define the MC,
and the JIc model was used to define the upper shelf master curve using Eqns. (3-3) and (3-9),
respectively. The intersection point of these two curves was called TUS [11, 14]. TUS can be
defined for any percentile of toughness values by using the same percentiles in Eqns. (3-4) and
(3-11) for KJc and JIc, respectively.
TUS was defined for individual datasets containing both KJc and JIc data. The T0 value determined
for a given KJc dataset was used to locate the median KJc MC (converted to Jc using Eqn. (4) in
[14]). The best least squares fit of the upper shelf master curve [64] through measured JIc data
was used to define the average JIc curve. The intersection of the median Jc curve with the
average JIc curve was used to define the temperature, TUS, for the given dataset. Once TUS was
established for each dataset, a fit to the plot of TUS versus T0 was used to define the relationship
described by Eqn. (4-1).
Fracture Toughness

100 MPa√m
TUS

To

Temperature
Figure 4-2
Schematic illustrating the relationship between the transition and upper shelf toughness,
and defining TUS as the intersection of the Wallin MC and the upper shelf MC.

4.1.4 Model Validation


Considerable data was used in the development of the TUS model [11, 14] that included RPV
steels and low alloy, naval steels, spanning T0 values over a broad range from −180oC to +140oC.
This data included both irradiated and non-irradiated plates, forgings and welds, with a variety of
chemical compositions and heat treatments. This model has strong theoretical and empirical
support and shows a very clear trend with little scatter.
Comparisons of the TUS model predictions to individual KJc/JIc datasets show that the accuracy
with which upper shelf behavior is predicted from T0 depends primarily on the accuracy with

4-3
Fracture Toughness Linkage Models in CC N-830-1

which T0 and TUS can be determined from measured data. This accuracy is strongly influenced
by the size of individual KJc and JIc data sets and not by the material variability. The KJc MC is
very steep at the transition from cleavage to upper shelf behavior, so predictions of upper shelf
behavior from T0 are very sensitive to very small variations, or uncertainties, in T0. This was
confirmed by EricksonkKirk, et al. [14] in their investigation that studied the effects of dataset
size (i.e., how many KJc and JIc values were available for determining T0 and TUS) on the fitting
error of Eqn. (4-1). They found that the error in the fit was inversely proportional to the number
of both KJc and JIc values in the dataset; the higher the number of values, the smaller the fit error.
The conclusion was that the uncertainty in the TUS model is not a material-dependent effect, but
rather is due to epistemic uncertainties on measured T0 and TUS determinations.

4.1.5 Limits on Validity of the Model


Empirically, the TUS model applies over a wide range of T0 values, from -180°C to +140°C, with
very little scatter, as shown in Figure (4-1). The model is based on a variety of ferritic steels,
both irradiated and un-irradiated nuclear RPV steels (including different product forms and “low
upper shelf” welds), mild steels, and the higher-strength copper precipitation hardened steels
(ASTM A710 and HSLA-100) used by the U.S. Navy in surface ship fabrication. An error
analysis of the data [14] demonstrated that material-dependent effects are not responsible for the
uncertainty in the relationship between T0 and TUS shown in Figure (4-1). This evidence,
combined with the physical basis information summarized in Chapter 3, suggests that Eqn. (4-1)
should apply without exception to any ferritic steel.
Eqn. (4-1) only applies to ferritic steels. Extrapolation beyond the data used to develop the
model (-180oC < T0 < +140oC) should be validated prior to use of Eqn. (4-1) in any analysis.

4.2 The Relationship Between Cleavage Crack Initiation (KJc) and Arrest
(KIa)
It is generally recognized that steels with higher amounts of hardening (that is, a higher T0 value)
tend to have less separation between the cleavage crack initiation (KJc) and cleavage crack arrest
(KIa) curves. This relationship was first shown by Wallin and Rintamma in 1998 [60] using a
large quantity of data. Further work on this concept, which added more data and provided a
physical basis for the observation, was reported in 2002 and 2014 by Kirk, et al. [61, 73]. Kirk’s
2014 work included a large amount of data that was used to refine the models presented in the
1998 and 2002 studies [60, 61]. However, it should be noted that even though the 2014 work
included considerably more data than that reported in either 1998 or 2002, the relationship did
not change significantly.

4.2.1 Mathematical Form of Model


Figure 4-3 shows the data from Kirk’s 2014 investigation [73]. These define the relationship
between TKIa and T0 as shown in Eqn. (3-6) and repeated here:
𝑻𝑻𝑲𝑲𝑲𝑲𝑲𝑲 = 𝑻𝑻𝒐𝒐 + 𝟒𝟒𝟒𝟒. 𝟗𝟗𝟗𝟗𝟗𝟗𝟗𝟗𝟗𝟗[−𝟎𝟎. 𝟎𝟎𝟎𝟎𝟔𝟔𝟔𝟔𝟔𝟔𝑻𝑻𝒐𝒐 ] Eq. 4-2

4-4
Fracture Toughness Linkage Models in CC N-830-1

160

120 Base

TKIa - To [°C]
Weld
HAZ
80
High Yield
Low Yield
40 Original Fit
All Data Fit
RPV Data Fit
0
-200 -100 0 100 200
To [°C]

Figure 4-3
Illustration of variation in the temperature separation between the KIa and KJc master
curves as a function of T0 [73].
Eqn. (4-2) is represented by the dashed black line labeled “All Data Fit.”

4.2.2 Physical Basis for the Model


Materials experience higher strain rates prior to crack arrest compared to strain rates experienced
prior to quasi-static crack initiation. High strain rates elevate the energy required to move
dislocations past barriers, resulting in an increase in the apparent yield stress of the steel in a
manner similar to that produced by prior strain hardening. As shown in Figure 4-4, Kirk, et al.
[61] used the idea of a universal hardening curve for all ferritic steels to explain how the effects
of strain elevation caused by the dynamic strain rate associated with crack arrest (defined as ∆εo)
produces a progressively diminishing elevation in the yield strength as prior hardening (εo)
increases. As shown on the figure, incremental increases in strain hardening (∆ε) produce
diminishing amounts of increase in yield strength. This suggests that the dynamic strain rates
associated with crack arrest will have a smaller impact on crack arrest toughness of high
hardness material than they will on materials with lower yield strengths/hardness. Since T0
scales with yield strength, this theory of progressively diminishing effects of hardening on yield
strength elevation (whether due to strain hardening or strain rate hardening) suggests a physical
basis for the observed trend of a progressively diminishing separation between the crack
initiation and crack arrest curves for steels with higher T0 values. The invariance of the true
stress at the maximum load that follows directly from the universal hardening curve suggests
that, in the limit of very high strength ferritic materials, the crack initiation and crack arrest
transition curves should approach each other (i.e., TKIa ≈ T0). This trend is demonstrated by the
available data.

4-5
Fracture Toughness Linkage Models in CC N-830-1

Figure 4-4
Illustration of the effects of strain rate increase on yield strength elevation for materials
having different degrees of prior strain hardening [61].

4.2.3 Empirical Basis


The TKIa/T0 linkage model is supported by data from 64 ferritic steels for which sufficient KJc and
KIa data were available to define, respectively, the KJc MC defined by Eqn. (3-3), and the crack
arrest MC defined by Eqn. (3-6) [71]. These materials are predominately nuclear grade base and
weld metals that were tested in both the unirradiated and irradiated conditions. Additionally,
some data for mild and high strength steels supported the model. The yield strengths of the
steels tested ranged from 280 MPa to 1,082 MPa, with the majority of strengths between 450
MPa and 650 MPa. Details on the data used in development of the TKIa/T0 linkage model are
provided in the next section, with further details provided in References [60, 62].

4-6
Fracture Toughness Linkage Models in CC N-830-1

4.2.4 Model Validation


Validation of the TKIa/T0 linkage model presented in Eqn. (4-2) was performed by Kirk, et al.
[73] using the data from the CARINA project [62], as well as data from the original TKIa/T0
model development study [61]. In their study, Kirk, et al. binned the data for regression analysis
to fit the original data, the CARINA data, followed by fitting all the data using a model of the
form:
𝑻𝑻𝑲𝑲𝑰𝑰𝑰𝑰 − 𝑻𝑻𝟎𝟎 = 𝑨𝑨 × 𝒆𝒆𝒆𝒆𝒆𝒆[𝑩𝑩 × 𝑻𝑻𝟎𝟎 ] Eq. 4-3

where A and B are fitting parameters and TKIa and T0 are in oC. The authors compared the
results to the original fit [61], including a statistical assessment relative to fit bias, trend with T0,
and uncertainty, or scatter, about the mean to assess the log-normal distribution fit the data.
The conclusions of the statistical assessment [73] were as follows:
• the original TKIa/T0 relationship [60, 61] well-represented each of the datasets (well within
the margin of error),
• there were no trends exhibited by any of the data that clearly differed from trends exhibited
by all the data, and
• limiting the data to only RPV steels did not change the fit. The trends observed for only the
RPV data were the same as trends observed for all or the data.
This statistical assessment validates the use of Eqn. (4-2) for use in assessment of all ferritic
steels.

4.2.5 Limits on Validity of the Model


Empirically, the TKIa/T0 linkage model applies over a wide range of T0 values, from -150oC to
+175°C. The model is based on a variety of ferritic steels, both irradiated and un-irradiated
nuclear RPV steels (including different product forms and “low upper shelf” welds), as well as
non-RPV steels having yield strengths that are much higher and lower than characteristic of RPV
steels.
The TKIa/T0 linkage model applies only to ferritic steels with KJc tested at quasi-static strain rates.
The model should not be used outside the ranges of the temperatures and yield strengths of the
data used in the model development and validation.

4.3 The Relationship Between Upper Shelf (JIc) Crack Initiation and Upper
Shelf Crack Growth (J-R)
All other models used in Code Case N-830-1, and summarized above, predict toughness values,
or link different toughness values, as a function of T0. However, prior to efforts undertaken
within the WGFE to develop Code Case N-830-1, all J-R curve models were formulated as a
function of Charpy upper shelf energy (USE) instead of T0 [74-76]. Therefore, in 2015 Kirk,
Erickson, and Stevens [77] undertook development of a model to predict the J-R curve from
information on JIc and the product form of the material. JIc, and its temperature dependence are
predicted from T0 as shown in Eqns. (3-9). Combining the models in Eqns. (3-9) with those
presented earlier in this section provides a means to estimate J-R curves from T0.

4-7
Fracture Toughness Linkage Models in CC N-830-1

4.3.1 Mathematical Form of the Model


J-R curves are represented by the following two-parameter power-law curve:

𝑱𝑱 = 𝑪𝑪 × ∆𝒂𝒂𝒏𝒏 Eq. 4-4(a)

Where C and n are parameters fit to the J vs. ∆a data for a particular tested specimen. The
database used in Reference [74] showed strong correlations between JIc and both the C and n
fitting parameters. The following model was developed from the data summarized in Table 4-1
[77].

𝑱𝑱𝒎𝒎𝒎𝒎𝒎𝒎𝒎𝒎
𝑿𝑿 = 𝑪𝑪 × ∆𝒂𝒂𝒏𝒏 Eq. 4-4(b)

𝑪𝑪 = 𝟏𝟏. 𝟔𝟔 × 𝑱𝑱𝒎𝒎𝒎𝒎𝒎𝒎𝒎𝒎
𝑰𝑰𝑰𝑰 Eq. 4-4(c)

𝒏𝒏 = 𝟎𝟎. 𝟎𝟎𝟎𝟎𝟎𝟎 × 𝑪𝑪𝟎𝟎.𝟑𝟑𝟑𝟑 Eq. 4-4(d)

𝑿𝑿 = ∆𝒂𝒂 Eq. 4-4(e)

In these equations, “X” is used as a subscript for compactness and signifies a particular amount
of ductile crack extension, or Δa, in mm. Values of J at a particular ductile crack extension (for
example, the value of J at 2.54 mm (0.1 inch)) can be determined by using these equations.
Also, the entire J-R curve can be produced by solving these equations for a range of Δa values.
Lower and upper bounds on J-R can also be predicted using the following formula:
𝒑𝒑
𝑱𝑱𝑿𝑿 = 𝒆𝒆𝒆𝒆𝒆𝒆�𝒍𝒍𝒍𝒍[𝑱𝑱𝒎𝒎𝒎𝒎𝒎𝒎𝒎𝒎 ] − 𝑴𝑴𝒑𝒑 × 𝑹𝑹𝑹𝑹𝑹𝑹𝑹𝑹� Eq. 4-5(a)
for lower bound curves: 𝑿𝑿

(𝟏𝟏−𝒑𝒑)
= 𝒆𝒆𝒆𝒆𝒆𝒆�𝒍𝒍𝒍𝒍[𝑱𝑱𝒎𝒎𝒎𝒎𝒎𝒎𝒎𝒎 ] + 𝑴𝑴𝒑𝒑 × 𝑹𝑹𝑴𝑴𝑴𝑴𝑴𝑴� Eq. 4-5(b)
for upper bound curves: 𝑱𝑱𝑿𝑿 𝑿𝑿

The value of p should not exceed 0.5. Values of root mean square deviation (RMSD) for
different product forms appear in Table 4-1. There is no effect of component thickness or crack
front length on Eqns. (4-4) and (4-5).
Table 4-1
RMSD values for different product forms.

Product Form RMSD

Forging (A508) 0.112


Plate (A533B) 0.138

RPV Welds 0.131

Linde 80 RPV Welds 0.206

It is noted that the values of the coefficients contained in Eqns. 4-4(c) and 4-4(d) and the RMSD
values in Table 4-1 differ from those presented in [77]. Analyses subsequent to those described
in [77] identified errors which have been corrected here, as explained in [101].

4-8
Fracture Toughness Linkage Models in CC N-830-1

4.3.2 Empirical Basis


The goal of ASME CC N-830-1 is to define a suite of best estimate fracture toughness models
for predicting fracture behavior across a broad range of temperatures. All the models described
in this report were available prior to initiating the revision to CC N-830. The only missing
information was a model for determining J-R behavior from T0. Such a model allows for
prediction of J at any level of crack growth based on knowledge of T0 value. To keep
development of the model as simple as possible, a straight, empirical approach was taken to fit
the trends observed in the data.
The database compiled by Eason [74], and shown in Table 4-2, was used as the basis for the
empirical model development Reference [77]. This included data from 403 J-R curves from
nuclear grade base and weld metals that were tested in unirradiated and irradiated conditions
using the procedures in ASTM 1820 [72]. Both two-parameter and three-parameters fits to the
data were made to determine which provided the most accurate representation of the 403 sets of
data. The forms of these fitting equations are given by:

𝑱𝑱 = 𝑪𝑪 × ∆𝒂𝒂𝒏𝒏 Eq. 4-6(a)


𝒏𝒏𝟐𝟐 )
𝑱𝑱 = 𝑫𝑫 × ∆𝒂𝒂(𝒏𝒏𝟏𝟏×∆𝒂𝒂 Eq. 4-6(b)

where C, D, n, n1 and n2 are fitting parameters used to fit individual sets of J-∆a data. Both
equations were determined to fit the data well for small amounts of crack extension, but the three
fitting-parameter form shown in Eqn. (4-6b) provided a more accurate representation of the data
at larger crack extensions. Based on these findings, Kirk, et al. developed the J-R/T0 model
using Eqn. (4-6b).
Correlations were identified between JIc and the J-R curve fitting parameter C, and between the
J-R curve fitting parameters C and n. The goal was to develop a correlation between J-R and T0
since it was already established that JIc is correlated to T0.
Best fit parameters were developed using the two-parameter and three parameter models, with
and without variability due to product form. Based on comparisons of the model predictions to
measured data relative to bias and scatter, the two-parameter model, which accounts only for the
effects of product form in predicting the scatter about the mean, was selected for incorporation
into CC N-830-1. This is the model defined by Eqns. (4-4) and (4-5).
Table 4-2
Composition of the J-R curve database.

# of Unirradiated # of Irradiated
Product Form Totals
Specimens Specimens

Forging (A508) 55 7 62

Plate (A533B) 62 18 80

RPV Welds 30 29 59

Linde 80 RPV Welds 113 89 202

Totals 260 143 403

4-9
Fracture Toughness Linkage Models in CC N-830-1

4.3.3 Model Validation


An independent validation effort has not yet been conducted on this model. However, the
prediction capabilities of the model described by Eqns. (4-4) and (4-5) were compared to the
predictive capabilities of other, candidate models for predicting the measured J-R curves
contained in the database [77]. The two-parameter model performed similarly to the three-
parameter model, despite the fact that the three-parameter model showed a better representation
of individual specimen J-R curves at large amounts of crack extension. Because of this similar
predictive performance, the complexity of the three-parameter model could not be justified. The
two-parameter, JIc-based model performed better than the USE-based, J-R curve predictive
models to which it was compared.

4.3.4 Limits on Validity of Model


Eqns. (4-4) and (4-5) can be used to predict J-∆a over a crack extension range of 0.5 to 10 mm
for temperatures up to 300°C, which reflects the range of crack growth and test temperatures
used in the model development. While the equations may have applicability to a broader range
of ferritic steels, their applicability has only been demonstrated to-date for nuclear grade ferritic
steels and their weldments. This model is strictly an empirically-derived model. In the future,
additional development work may be undertaken to assess the applicability of this model, and/or
its expansion, to other grades of RPV steel (i.e., A302B) and to ferritic piping steels. The
underlying physics of this model have not yet been investigated.

4-10
5
IMPLICATIONS OF PROPOSED CHANGES

5.1 Introduction
The changes adopted in Code Case (CC) N-830-1 are an alternative to the RTNDT-indexed KIc
fracture toughness curve prescribed in Section XI, Nonmandatory Appendix A with curves that
better reflect the scatter and temperature dependence observed in material fracture toughness
from lower shelf, through transition to upper shelf behavior, including crack arrest. The RTNDT-
indexed KIc curve bounds most, but not all, of the available fracture toughness data [12] because
it does not fully account for the uncertainty inherent in using the Charpy-based RTNDT to index a
fracture toughness curve, and the KIc curve was developed based on a limited fracture toughness
dataset. Using the updated CC N-830-1 fracture toughness models eliminates the implicit
conservatism inherent in the RTNDT-indexed KIc curve method, and replaces it with a full
distribution of toughness curves that more consistently and accurately represents fracture
toughness behavior and enables selection of a bounding value that explicitly and quantitatively
accounts for uncertainties. This chapter discusses the sources and treatments of uncertainties in
the Appendix A flaw evaluation methodology, in the original CC N-830 method adopted in
2014, and in the new CC N-830-1 method and compares the material fracture toughness curves
contained in each their ability to represent measured fracture toughness data.

5.2 Sources of Uncertainties in Fracture Mechanics Analyses


Fitness-for-Service (FFS) procedures typically contain methods to account for uncertainties that
arise due to incompleteness of models, lack of knowledge about the inputs and material behavior,
and random variability of material properties. In probabilistic methods, uncertainties are
typically differentiated by those that are due to random variability (aleatory uncertainty), and
those that arise due to a lack of knowledge (epistemic uncertainty). This differentiation allows
for implementation of sampling methods to properly account for the effects of each type of
uncertainty on the outcome, resulting in refinement of outcome estimation. In deterministic
methods, these uncertainties are not differentiated, and are rarely quantified. Uncertainties in
deterministic methods are typically treated using excess conservatisms (bounding models, safety
factors, etc.), but the lack of characterization of the uncertainties makes it difficult to quantify the
conservative bias of the results and thus, the level of safety provided by the deterministic
method. Experience over time with deterministic assessments and their results contribute to a
level of confidence that the factors applied are sufficiently conservative but not overly
operationally restrictive. Despite the inability to directly compare deterministic results to a risk-
informed safety metric, the veracity of newer probabilistic models is often judged relative to
deterministic methods due to the greater level of experience with deterministic methods even
though probabilistic methods provide a more transparent description of the physical reality being
modeled.

5-1
Implications of Proposed Changes

In this chapter, the sources and treatment of uncertainties defined by relevant flaw evaluation
codes are discussed with regard to the implications of the changes in CC N-830-1, and compared
to provide a foundation for the proposed treatment of uncertainties in a future revision to CC N-
830-1. This discussion is included to help place context on the new fracture toughness models
incorporated into CC N-830-1 and their intended use.
The typical uncertainties associated with FFS assessments are divided into three parts for this
discussion: (1) uncertainty due to flaw size, (2) uncertainty due to applied stress, and (3)
uncertainty due to the material resistance to crack extension characterized by the material
fracture toughness. These three sources of uncertainties are described in the following sections,
followed by a discussion of how these uncertainties are treated in various flaw evaluation codes.

5.2.1 Flaw Size Uncertainty


For the nuclear-grade ferritic materials that comprise the reactor pressure vessel (RPV) primary
pressure boundary, flaw size and location is typically determined using nondestructive
examination (NDE) volumetric methods. These methods, when used in the field, have a
considerable uncertainty associated with them [78] from a probability of detection perspective as
well as from a sizing and location accuracy perspective. These uncertainties are dependent upon
several factors, including NDE delivery technique, component accessibility, inspector
experience, tool positioning and placement, and material surface finish and homogeneity.
The U.S. nuclear industry has conducted extensive investigations and demonstrations of NDE
techniques appropriate for volumetric inspection of RPVs. These efforts have included the
design, fabrication, and inspection of mockups containing realistic simulations of the degradation
mechanisms of concern. These investigations resulted in the development of evaluation factors,
which are numerical values, related to the uncertainties inherent in delivering and executing an
NDE technique in a RPV. These evaluation factors are combined with the actual results of an
inspection to form input into a fracture-mechanics assessment of the component’s serviceability.

5.2.2 Stress and Stress Intensity Factor Uncertainty


Calculation of stresses due to applied loadings depend on knowledge of component geometry,
loading conditions and time variability of loading, method of calculating stress, the accuracy of
the fracture mechanics models used to determine crack tip stress intensity factors, and the
accuracy of crack growth rate models used to determine the end-of-evaluation-period flaw size.
Flaw size and location are also important factors in determining the stress at the flaw location as
stresses vary throughout a component based on the flaw’s proximity to applied loads, geometric
or material discontinuities and free surfaces, including other flaws.
Article A-3000 of Section XI, Nonmandatory Appendix A contains a procedure for calculating
the stress intensity factor for planar flaws. The procedure depends on knowledge of the applied
stresses from all forms of loading including pressure, thermal gradients, welding, and cladding.
While pressure is generally well-known and carefully controlled, weld residual stress (WRS),
cladding-induced stresses, and thermal stresses have significant variability, so they are
significant contributors to the uncertainties in stress characterization. There is also an
uncertainty contribution from the analysis methods chosen by the individual analyst, particularly
in the assumptions made concerning the geometric idealization of the component, model
refinement (i.e., finite element methods, hand calculations, etc.), discretization of stresses

5-2
Implications of Proposed Changes

through the component thickness (whether the stresses are fit using a polynomial or linear stress
distribution, interpolated), non-linear behavior (e.g., plasticity), etc. These analyst-dependent
factors can significantly alter the calculated applied stresses, and therefore lead to variations in
the estimate of critical flaw size. This is particularly true when the crack driving force, KI-applied,
calculated from the stresses, flaw size and location are compared to the allowable toughness,
KJcmaterial, in the steep part of the fracture toughness transition curve [79].

5.2.3 Fracture Toughness Uncertainty


The uncertainty in fracture toughness measurements arise from inherent material inhomogeneity
and the models used to characterize the toughness. The ferritic steels that are relevant to
Appendix A and CC N-830 analyses are comprised of a body-centered-cubic (BCC) iron matrix,
which contains a small amount of dissolved carbon and other alloying elements, along with
precipitates of iron carbide. Inhomogeneity arises due to grain size variation, precipitate size and
density variations through the thickness, and chemical variability arising from alloy segregation,
all due to variations in heating and cooling rates during fabrication of thick-section components.
These microstructural inhomogeneities cause uncertainties in measured ferritic steel properties
such as fracture toughness.
Material inhomogeneity is not the only source of material property uncertainty in FFS
assessments. Test methodology contributes to uncertainty as does the conversion from
properties determined from specimens tested in laboratory conditions (e.g. small specimen sizes,
uniform, quasi-static, loading rate, controlled environment) to actual components in the field.
Constraint and crack front length are among the sources of uncertainties when converting
laboratory-based fracture toughness models to predicting toughness in actual components.
Currently the assumption is that laboratory-measured toughness values apply 1:1 to field
conditions.

5.3 Treatment of Uncertainties


A discussion of the treatment of uncertainties in the context of a fracture safety acceptance
criteria is important. However, the flaw assessment procedure in ASME Appendix A does not
contain a risk-based acceptance criteria; rather, it requires that Kdriving be less than αKresistance at
all times and for all loadings, where α is a structural factor to compensate for uncertainties and to
add conservatism. Since CC N-830-1 only addresses alternative equations for describing
material toughness, the comparisons of the effects of uncertainty treatment at the end of this
chapter are limited to only those uncertainties that apply to material toughness. For
completeness, and to provide context for the subsequent quantitative discussion and comparisons
of material toughness uncertainty treatments, a qualitative discussion of the treatment of
uncertainties on flaw size and location, and stress calculations are first summarized.

5.3.1 Treatment of Uncertainty Due to Flaw Size and Location


Prior to flaw assessment, the flaw must first be located and sized using NDE techniques.
Generally, for through-wall depth and surface or subsurface location, most examinations must be
performed using volumetric techniques such as ultrasonic testing (UT). Because of uncertainties,
the procedures used in UT inspections may result in detected flaw sizes that are different than the
actual flaw dimensions. These differences depend on the equipment and the actual size of the

5-3
Implications of Proposed Changes

flaw. Generally, smaller flaws are oversized by UT equipment, and larger flaws are undersized.
These differences are, in part, caused by the physical characteristics of the equipment. For
example, flaws that are smaller than the UT wavelength cannot be sized; in this situation, an
echo from the UT equipment will typically be sized as the wavelength of the UT beam (see
Figure 5-1, top). For large indications that interact with the UT beam on multiple scans, flaw
size is more difficult to estimate because it is more difficult to precisely locate the flaw tips (see
Figure 5-1, bottom). While other factors not discussed here influence UT estimates, and while
some examples exist of flaw under-sizing by UT, the factors discussed here make the probability
of an overcall (over-estimating the physical size of the flaw) more likely than that of an
undercall. If the flaw size is overestimated the ligament either between one flaw and another
situated nearby, or between the subject flaw and the free surface of the component will be
underestimated. Under-estimation of the ligament increases the likelihood of adjacent flaws
being treated as a single larger flaw, or of near-surface flaws being classified as surface-breaking
via the application of flaw interaction rules, as will be discussed in the following paragraphs.

Figure 5-1
Schematic illustration of physical causes for systematic over-estimation of flaw size using
UT.

5-4
Implications of Proposed Changes

Article IWA-3000 contains procedures for characterizing flaws detected during NDE. IWA-
3300 provides methods to conservatively bound flaw sizes, and provides proximity rules for
determining when subsurface flaws should be treated as surface flaws, and for when multiple
flaws should be combined and treated as a single, larger flaw. These proximity rules are based
on fracture mechanics assessment of interacting flaws, and generally combine two adjacent flaws
when the stress intensity factor of one flaw affects the adjacent flaw by more than about 15%.
Thus, the simplification of combining adjacent flaws tends to be conservative. A subsurface
flaw is treated as a surface flaw if the distance between the flaw and the component surface is
less than or equal to 0.4d, where 2d is the through-thickness measured depth of the flaw. Two
flaws (either surface or internal) are combined, with the overall flaw dimension given by the
rectangle encompassing both flaws, if the smallest distance between the flaws is less than or
equal to the larger of the two flaw depths. The Section XI proximity rules have been shown to
be conservative in work that assessed surface flaw interactions [80, 81].
Much of the conservatism in the Section XI flaw evaluation procedure arises from the
simplifications made to what is, essentially, a complex problem, as described in Section B-2 of
EPRI NP-719-SR [6]:
“Recognizing the limits of ultrasonic examination techniques to define precisely
the dimensions, areas, and orientation of flaws, the code rules incorporate many
simplifications that obviate the need to determine the precise flaw size and
orientation”
Section XI flaw evaluation assumes that all observed indications, such as crack-like defects, slag
inclusions, porosity, lack of weld fusion, laminations, and any combinations thereof should be
treated as planar cracks. In addition, to simplify the analyses of detected flaws, irregularly
shaped flaws are represented by idealized simple geometric shapes. Thus, the development of
flaw standards criteria was simplified to facilitate the application of the principles of fracture
mechanics [6]. Such considerations were based, in large part, on judgements rather than precise
science, and their continued use is supported by field observations.
Experimental evidence suggests that the stress fields of two surface cracks will begin to interact
when the distance between them is between 0.0d and 0.75d, where d is the depth of the surface
flaws. This is less than the value of d adopted in IWA-3300. The interaction of the stress fields
of two subsurface flaws is more constrained than for surface flaws, so using the surface flaw
proximity criterion for subsurface flaw spacing is even more conservative [81]. Dulieu and
Lacroix provide further evidence of this conservatism in their assessment of the flaw interaction
rules for quasi-laminar hydrogen flakes [80]. They show that when flaws grouped by interaction
rules are instead analyzed as separate flaws, the resultant driving force on the un-grouped flaws
drops by a factor of 2 to 3.
To date, the best available quantitative treatment of the total uncertainties associated with NDE
may be estimated from the work performed by the EPRI NDE Center. EPRI’s Performance
Demonstration Initiative (PDI) estimates NDE uncertainties, and includes development of
models that describe the probability of detection (POD) and sizing accuracy of flaws using UT
techniques [78]. Although this work provides quantitative characterization of the uncertainty
inherent in the NDE procedures, it does not assess the uncertainties associated with the IWA-
3300 flaw proximity rules.

5-5
Implications of Proposed Changes

ASME Appendix A does not currently address NDE uncertainties. Flaw sizes estimated from
UT exams are used as input to ASME Code flaw evaluations. ASME Appendix A requires that
the end-of-evaluation-period flaw size be used, so the UT flaw size must be increased by the
amount of crack growth anticipated over the evaluated period of operation before additional
inspection or repair is performed. As a result, most UT flaw sizes are increased by some amount
for flaw assessment. For these reasons, it is recommended that the future revision of CC-N830-1
consider NDE uncertainties associated with flaw size and characterization based on EPRI’s POD
and sizing accuracy work [78]. It is anticipated that the next revision of CC N-830-1 will adopt
partial structural factors (PSFs) to enable an explicit accounting of flaw size uncertainty.

5.3.2 Treatment of Uncertainty Due to Stress


In a RPV, stresses are caused by internal pressure, thermal gradients experienced during plant
transients (actual or postulated accident events), earthquakes, attached piping loads, and welding
residual stresses (from both structural welds and cladding). These stresses can be estimated
using either closed-form equations (provided in the Code or elsewhere) or by using numerical
techniques, such as the finite element (FE) method. Advanced numerical techniques such as FE
offer accuracies for complex geometries that were, in practical terms, unobtainable as little as a
few decades ago. Because these FE advanced methods were not readily available at the time that
many of the procedures for analysis of the structural integrity of nuclear components were
developed, conservative, simplified approximations were used to provide the closed-form
solutions contained in the Code. Even today, with the ready availability of very accurate
solutions, the method chosen for a particular analysis is usually driven by factors that include the
training and experience of the engineers involved, and the economic considerations of the
problem to be addressed. More often than not, a simple closed-form solution can be chosen to
represent a more complex situation even though the closed-form solution may be a much more
conservative approach. Oftentimes the high toughness and flaw tolerance characteristics of
nuclear grade steels enable the stress analysts to adopt conservative stress estimates while still
demonstrating adequate structural integrity.
Engineers tend to gravitate towards conservative (i.e., high) estimates of stress, but beyond that,
the ASME Code methods also contribute to additional conservatism. This additional
conservatism come in two forms:
• Intentionally conservative stress estimates: The Code generally provides the analyst with
different options for stress estimation. Sometimes these options are intentionally
conservative to allow for simpler evaluation. Some examples of intentionally conservative
stress estimates that appear in the Code are as follows:
– Section XI Appendix A: Article A-3211 provides a method using a fourth-order
polynomial to represent stresses acting normal to the plane of the flaw. Generally, this
polynomial is fit to the stresses acting over the whole crack depth; however, A-3211 also
states, “Alternatively, the stress distribution that upper bounds the actual stress field over
the flaw may be used.” The Code provides an example of a “conservative linearization of
a stress field” in its Figure A-3210-3.
– Section XI Appendix G: One option for thermal stress estimation in G-2214.3 is to use
equations that estimate the maximum thermal stress during a normal heat-up or cool-
down transient such that the analyst may assume that this maximum thermal stress

5-6
Implications of Proposed Changes

persists throughout the transient. This is very conservative compared to the more
complex method of evaluating time-varying thermal stresses throughout a transient,
which tend toward zero as isothermal conditions are achieved. Thus, for portions of
transients, the recommendations of G-2214.3 can be very conservative.
– Section XI Appendix K: For Service Level A and B loadings, K-2200(a) recommends
that the applied pressure inside the vessel be estimated as 1.15 times the accumulation
pressure (PACCUM) 4 for assessment of ductile crack initiation and 1.25 times PACCUM for
assessment of ductile crack growth stability. This guidance is grossly conservative, as
demonstrated by the following facts:
o The Code defines PACCUM as 1.1 times the design pressure, PDESIGN (again see
footnote 3).
o The operating pressure, POPERATING, is 0.9 times the PDESIGN.
o POPERATING is therefore 0.9÷1.1 = 0.81×PACCUM. Combining this with the
recommendations of K-2200(a) to use a pressure of either 1.15×PACCUM or
1.25×PACCUM, the result is that K-2200(a) recommends use of a pressure between 1.4
and 1.52 times higher than POPERATING. Additionally, it should be noted that the
safety relief valves prevent the RPV pressure from ever reaching PACCUM.
• Structural Factors: Structural factors appear in various places in the Code. These structural
factors normally increase the applied value of stress to introduce a margin between the
allowed flaw size and the critical flaw size. Some examples that apply exclusively to stress
estimation are as follows:
– Section XI Appendix G: G-2215 specifies a structural factor of 2.0 on pressure for
normal heat-up and cool-down loading. For hydro-test loading, G-2400(b) specifies a
structural factor of 1.5 on pressure.
The structural factors of √10 (for normal operating conditions) and √2 (for emergency
and faulted conditions) appear in IWB-3600 and, thus, are used in flaw evaluations
performed according to Section XI Appendix A. Since the √10 or √2 structural factors
reduce the allowable toughness, they impose a conservatism on the maximum allowed
value of applied stress times the square-root of flaw size.
In summary, the conservative nature of engineering stress analysts is further compounded by
ASME Code requirements that lead to conservative stress estimates. In all cases, the use of
structural factors increases the estimated applied stress to account for unknown-unknowns.
These factors all ensure that current stress estimates achieved following Code procedures are
conservative. It is anticipated that the next revision of CC N-830 will adopt PSFs to minimize
unnecessary conservatism and enable an explicit accounting of uncertainties associated with
stress estimation. The experience by other Code bodies that have implemented PSFs [82-84] is

4
In a PWR, PACCUM represents the maximum overshoot pressure that could result from a mass imbalance. A
mass imbalance could occur during an event when the low-temperature overpressure protection (LTOP) system and
the relief values open because the injection rate is too high. In ASME Code terminology PACCUM is defined as
being equal to 1.1 times the design pressure (1.1×PDESIGN). While BWRs have safety relief values there is no
clear equivalent to an “accumulation” pressure, so for BWRs the Code definition of PACCUM = 1.1×PDESIGN is
used along with the guidance of K-2200(a).

5-7
Implications of Proposed Changes

expected to help support the development of PSFs within ASME Section XI in a revision to CC
N-830-1.

5.3.3 Treatment of Uncertainty on Fracture Toughness


Appendix A allows use of either the RTNDT-KIc curve or the RTT0-KIc curve to describe material
fracture toughness. Section XI Paragraph IWB-3612 requires that these toughness values be
reduced by structural factors of √2 or √10 for assessing flaw acceptability for emergency/faulted
or normal/upset operating conditions, respectively. As described in Chapter 1, the RTNDT index
temperature does not represent a direct measure of fracture toughness, but is obtained from a
correlative approach (using Charpy and NDT testing) to characterizing toughness. This index
temperature provides a conservative estimate of the actual toughness transition index
temperature as represented by T0, as shown in Figure 5-2. Using RTT0 to index the KIc curve
provides an index temperature directly related to the fracture toughness transition temperature
with an additional margin of 19oC to account for uncertainties. This adds additional
conservatism beyond the structural factor that leads to an overly conservative representation of
the actual toughness transition temperature for a material.
The conservatism of both RTNDT and RTT0 is further increased by using the ASME KIc curve to
represent material fracture toughness in the transition region. Figure 5-3 compares the KJc data
overlaid with the 1% and 99% bounds of the MC, the RTNDT-indexed KIc curve (taking the 50%
cumulative probability distribution (CPD)) value of RTNDT-T0 = 50oC), and the RTT0-indexed KIc
curve. The conservatism of the RTNDT-indexed KIc and RTT0-indexed KIc curves is clear.
However, neither the RTNDT-indexed KIc curve nor the RTT0-indexed KIc curve bounds the data,
and both are non-conservative in the lower shelf toughness region.
The applied stress intensity factor and material resistance to crack extension/arrest determined in
ASME Appendix A are compared for acceptability using the equations in Paragraph IWB-3612
as follows:
Kapplied < KIc/√10 for normal operating conditions, and
Eq. 5-1

Kapplied < KIc/√2 for emergency/faulted conditions.


The structural factors of √2 and √10 are intended to account for uncertainties from all sources.
As described in EPRI Report No. NP-719-SR [6], Section XI adopted conservative assumptions
to accounts for unknowns. Supporting document WRC-175 [5] referred to these unknowns as
“factors of ignorance.” These structural factors add further conservatism beyond that included
in the material fracture toughness relative to actual data as shown in Figures 5-4 and 5-5.

5-8
Implications of Proposed Changes

Figure 5-2
Cumulative probability distribution function showing the relationship between RTNDT and
T0.

Figure 5-3
Plot of KJc, 1% MC bound, 99% MC bound, the RTNDT-indexed KIc curve, and the RTT0-
indexed KIc curve.

5-9
Implications of Proposed Changes

Figure 5-4
Plot of KJc, 1% MC bound, 99% MC bound, the RTNDT-indexed KIc curve divided by √2, and
the RTT0-indexed KIc curve divided by √2 (for emergency/faulted operating conditions).

Figure 5-5
Plot of KJc, 1% MC bound, 99% MC bound, the RTNDT-indexed KIc curve divided by √10, and
the RTT0-indexed KIc curve divided by √10 (for normal/upset operating conditions).

5-10
Implications of Proposed Changes

5.4 CC N-830
The original version of CC N-830 was adopted by ASME in 2014, and permits the 5th percentile
T0-indexed KJc MC to be used in lieu of either the RTNDT- or RTT0-referenced KIc curve for
determining the fracture toughness for comparison to the applied stress intensity factor in IWB-
3612. Following the procedure in IWB-3612, the value of fracture toughness determined using
CC N-830 is divided by the structural factors described in Eqn. (5-1) to account for uncertainties
from all sources. In contrast, the T0-indexed KJc MC provides a best-estimate model of material
fracture toughness, so using a 5th percentile lower bounding value from this curve provides a
conservative estimate of material fracture toughness that is consistent with two-sigma bounds
commonly adopted in regulatory assessments. When combined with the structural factors in
IWB-3612, the conservatism in the estimate of material fracture toughness increases
substantially as shown in Figure 5-6.
Dividing the 5th percentile CC N-830 MC by √2 to assess emergency/faulted operating
conditions results in a bounding value of the data approximately equal to the 1st percentile MC
bound. Therefore, this curve bounds 99% of the toughness data and accounts for 99% of the
expected scatter. Dividing the 5th percentile CC N-830 MC by √10 to assess normal/upset
operating conditions results in excessive conservatism; the resulting curve bounds all known KJc
toughness data and is equal to the conservatism present in the ASME Appendix A RTT0-indexed
KIc curve divided by √10 (shown in red in Figure 5-6).

Figure 5-6
Plot of KJc, 1% MC bound, 99% MC bound, the CC N-830 5% MC bound, the CC N-830 5%
MC divided by √2, and the CC N-830 5% MC divided by √10 (for emergency/faulted and
normal operating conditions, respectively).

5-11
Implications of Proposed Changes

5.5 Code Case N-830-1 Uncertainty Treatment


CC N-830-1 provides for use of the T0-indexed KJc MC in lieu of either the RTNDT or RTT0-
referenced KIc curve for determining the fracture toughness for comparison to the applied stress
intensity factor. CC N-830-1 provides no guidance on selection of appropriate bounding curves
to use in an analysis to account for uncertainty but instead leaves this to the User to define and
justify. In the comparisons below, the 1% and 0.5% KJc curves are shown, without any structural
factor, to indicate what “direct use” might look like. These curves are shown relative to fracture
toughness data and other Code-provided fracture toughness representations. A goal of CC N-
830-1 in moving forward is to enable implementation of fracture toughness bounding curves,
without use of structural factors, to account for uncertainty in fracture toughness directly and
consistently.
The 1st percentile MC provides similar (but more consistent) bounds as the Appendix A approach
for emergency and faulted conditions (either the RTT0-indexed KIc curve divided by √2 or the CC
N-830 5th percentile MC divided by √2), as shown by comparison of the dashed green curves in
Figures 5-7 to the magenta curve. The 0.5th percentile MC provides a slightly less conservative
bound to the fracture toughness data compared to either the Appendix A RTT0-indexed KIc curve
divided by √10 or the CC N-830 5th percentile MC divided by √10, as shown in Figure 5-8. The
excessive conservatism demonstrated by the RTT0-indexed KIc curve divided by √10 and 5th
percentile MC divided by √10 is not necessary to adequately bound the uncertainties inherent in
fracture toughness data. These conclusions are demonstrated by the more than 7,000 data points
contained in the MC database. The blue curves in Figure 5-8 both have values well below 20
MPa√m over most of the temperature range. Not only has no fracture toughness data for a
ferritic steel ever been observed at these low values, the technical basis for the MC Weibull
distribution provides strong support for the notion that such low values cannot occur, suggesting
that this amount of conservatism is not necessary for uncertainties in fracture toughness.

5-12
Implications of Proposed Changes

Figure 5-7
Plot of KJc, with the 1% MC bound, 99% MC bound, the CC N-830-1 1% MC bound, the CC
N-830 5% MC bound divided by √2, and the Appendix A RTT0-indexed KIc curve divided by
√2 (all for emergency/faulted operating conditions).
Note that the 1% MC and the 5% MC divided by a √2 structural factor coincide over most of the
temperature region shown.

5-13
Implications of Proposed Changes

Figure 5-8
Plot of KJc, 1% MC bound, 99% MC bound, the 0.5% MC bound (CC N-830-1), the 5% MC
bound divided by √10 (CC N-830), and the RTT0-indexed KIc curve divided by √10 (Appendix
A) for normal/upset operating conditions.

5.6 Summary
The method described in ASME Section XI, Non-Mandatory Appendix A for evaluating the
fracture tolerance of flaws is a deterministic procedure containing many implicit and explicit
conservatisms to account for the known and unknown sources of uncertainty inherent in the
fracture mechanics evaluation. Implicit conservatisms are contained in NDE sizing and location
procedures, flaw proximity rules, and stress analysis methods. Implicit conservatisms are also
contained in the methods used to define a material fracture toughness, including use of a lower
bound linear elastic fracture toughness curve indexed by a Charpy-based reference temperature,
or by a fracture toughness-based reference temperature with additional margin added.
Additional conservatism is explicitly applied to the material fracture toughness term by dividing
the lower bound value by a √2 or √10 structural factor depending on whether emergency/faulted
or normal operating are being evaluated.
The best-estimate models of fracture toughness contained in CC N-830-1 provide a much more
consistent representation of an extensive database of fracture toughness values than does either
the RTNDT or RTT0-indexed KIc curve, eliminating much of the implicit conservatism inherent to
the older, linear-elastic-based model. The full distribution fracture toughness curves defined in
CC N-830-1 enables selection of a bounding value that explicitly and quantitatively accounts for
uncertainties and can thus readily support not only deterministic evaluations but also
probabilistic methods.

5-14
Implications of Proposed Changes

The1st percentile bounding KJc MC curve for the emergency/faulted accident conditions is
approximately equivalent to the combination of structural factors and the 5th percentile fracture
toughness curve in accounting for uncertainties. The 0.5th percentile bounding KJc toughness
MC, without the explicit use of structural factors, provides a conservative lower bound sufficient
to account for all uncertainties. The use of the 1% and 0.5% bounding curves without structural
factors would provide a much more consistent representation of material fracture toughness
across all temperatures and conditions compared to curves reduced by structural factors. It
would be an improvement to Appendix A procedures to eliminate the existing structural factors
through the selection of an appropriate statistical bound to the fracture toughness curve or
through a combination of partial structural factors that reflect a consistent margin requirement
over the operating temperature range.

5-15
6
POTENTIAL CODE/REGULATORY APPLICATIONS OF
CC N-830-1

6.1 Introduction
CC N-830-1 provides a complete description of the fracture toughness of ferritic steel, including
all fracture toughness metrics, from the lower shelf through the upper shelf. These best-estimate
equations define the full distribution of expected toughness values, defining the temperature
dependence and scatter inherent to fracture toughness behavior. These equations can be used to
define any bounding curve to provide the desired level of conservatism in a deterministic fracture
mechanics evaluation, or they can be used to define input distributions for ferritic steel fracture
toughness for probabilistic assessments. In principle, the CC N-830-1 fracture toughness
equations can be used to define material resistance to fracture in any fracture mechanics, FFS
evaluation. Thus, the potential Code / Regulatory applications of the Code Case include any
instance for which fracture toughness is an input to an assessment or analysis. These include the
following:
Within the ASME Code
• Section XI
– The IWB 3500-1 table of allowable planar flaws.
– The acceptance criteria of IWB-3612 for regions remote from geometric discontinuities,
or of IWB 3613 for regions near geometric discontinuities
• Assessment of found flaws following the requirements of Nonmandatory Appendix A
• Evaluation of unanticipated operating events using Nonmandatory Appendix E [85]
• Assessment of pressure temperature limits using Nonmandatory Appendix G [86], including:
– G2215 Allowable pressure for shell regions
– G-2216, Risk-Informed allowable pressure
– G2223 Allowable pressure for nozzles
– G2400 for Hydrostatic Test Temperature determination
• Assessment of RPVs having low upper-self energy levels using Nonmandatory Appendix K
[8].
Within the Requirements of the NRC as Outlined within the Code of Federal Regulations
and Related Regulatory Guides
• In support of analysis of pressurized thermal shock following 10 CFR 50.61 or 10 CFR
50.61a [87]

6-1
Potential Code/Regulatory Applications of CC N-830-1

• In support of analysis of normal operating conditions following Appendix G to 10 CFR Part


50, including [9]:
– Appendix G to 10 CFR Part 50 incorporates by reference Nonmandatory Appendix G to
Section-XI of the ASME Code, so all the above listed situations apply.
– Estimation of minimum temperature requirements for the flange
– Demonstration of equivalent margins for low upper shelf materials (i.e., materials having
an upper shelf energy projected to fall below 68J
The one part of the suite of CC N-830-1 toughness models that has seen previous Code /
Regulatory use is the Wallin KJc Master Curve; the next section summarizes these uses. This
discussion is followed by a section that discusses currently foreseen uses for the CC N-830-1
toughness models.

6.2 Past Use of the Wallin Master Curve

6.2.1 Within the ASME Code


Shortly following the first adoption of the ASTM Standard Testing procedure for T0 estimation
in 1997 [15], the industry undertook an effort to justify the use of T0 to define a reference
temperature for the KIc and KIa curves equivalent in function to RTNDT. The associated report [44]
became the technical basis for definition of RTT0 as T0 + 19.4°C; the 19.4°C margin term shifts
the median Master Curve so that it provides an ≈95% bounding curve to the dataset used in the
early 1970s to establish the ASME KIc curve. Within the Code, RTT0 was first introduced in 1999
through Code Cases N-629 and N-631, applicable to Sections XI and III, respectively [16, 17].
Use of RTT0 indexing was later incorporated directly into Appendices A and G of Section XI.
RTT0 is also used in the Code to bound KIa data to ensure appropriate bounding [7]. Further Code
applications of the Master Curve include the first publication of CC N-830 in 2014 [18], which
was discussed earlier in this report.

6.2.2 NRC Applications


Beginning in the 1990s there have been three uses of data generated within the framework of the
Wallin Master Curves that have been used by the industry with NRC approval. These are
summarized below.
In 1993 the B&W Owners’ Group (BWOG) published BAW-2202 [88], which used Master
Curve KJc data to justify use of an unirradiated value of RTNDT based only on NDT data, not on
NDT and Charpy data as defined by ASME NB-2331. The effort was motivated by the
unusually high values (between +14.4 and +50.6oC) of unirradiated RTNDT for Linde 80 welds of
weld wire heat WF-70. These values, applicable to beltline welds in Rancho Seco, Crystal
River-3, TMI-1, and Zion Units 1 and 2, led to predictions of RTPTS values for these units
exceeding the 10 CFR 50.61 screening criteria before 40 years of operation. The BWOG used a
draft version of ASTM E1921, which appears as an appendix to BAW-2202, to generate KJc data
for WF-70 welds under both quasi-static and dynamic loading rates. These data were compared,
respectively, to KIR and KIc curves indexed to a mean TNDT value of -48°C with a standard
deviation (σ) of 8.2°C to establish the margin term. The -48 and 8.2°C values were based on 24
TNDT values from BWOG operating plants and from work performed at the Oak Ridge National

6-2
Potential Code/Regulatory Applications of CC N-830-1

Laboratory [89] on the WF-70 weld harvested from the cancelled Midland Nuclear Power Plant
in Michigan. The comparison showed that the KIR and KIc curves bounded the data, and
ultimately the NRC approved use of the -48 and 8.2°C values based on this justification. BAW-
2202 does not mention why a TNDT value was used to index the KIR and KIc curves, but it may
have had something to do with the fact that the ASTM standard was still a draft, and that there
was, at the time, no context within the ASME Code to use T0 values or KJc data.
In follow-on work first submitted to the NRC in 2002, the BWOG expanded the BAW-2202
work to establish alternative unirradiated RTNDT values, and associated σ terms, for different
heats of Linde 80 welds present in the beltline regions of B&W fabricated vessels. The topical
report, BAW-2308, uses the Master Curve reference temperature approach for determining RTT0
as described in ASME Code Case N-629 and based its generic RTT0 values on large KJc data
populations for each weld wire heat [90]; in total over 300 KJc specimens were tested. The NRC
reviewed this topical report, but required plant-specific submittals for each application of the
approach. To date the following thirteen units have been granted exemptions to use the RTT0-
based RTNDT values to demonstrate compliance with the requirements of the PTS rule (10 CFR
50.61a):
• Arkansas Nuclear Unit 1
• Davis Besse
• Oconee Units 1, 2, and 3
• Three Mile Island Unit 1
• Crystal River Unit 3
• Turkey Point Units 1 and 2
• Surry Units 1 and 2
• Point Beach Units 1 and 2
In their calculation of generic RTT0 values to use with each heat the BAW-2308 approach
includes the following factors to account for uncertainties:
• An added factor (bias) of +10°C when data from precracked Charpy specimens is used to
account for the constraint differences between bend and compact tension specimens.
• A factor included in the square-root sum-of-squares (SRSS) margin calculation of between
4°C -9.5°C to account for variability associated with test procedure and material
inhomogeneity, and
• A factor included in the SRSS margin calculation of between 1°C to 6°C to account for lack
of T0 estimation precision due to finite sample sizes.
Table 6-1 summarizes the outcome of the BAW-2308 effort, and indicates a reduction of
estimated unirradiated index temperature from 15°C to 55°C depending on the weld wire heat.
Similar to the data in Figure 5-2, these data attest to the conservatism implicit to RTNDT indexing
procedures.
Beginning in 1998 the Kewaunee Nuclear Power Plant used Master Curve data and CC N-629
[91] to obtain a license amendment to use these data in its assessment of the PTS screening
criteria (10 CFR 50.61a), in setting pressure-temperature limits (10 CFR 50 Appendix G), and as

6-3
Potential Code/Regulatory Applications of CC N-830-1

part of the surveillance program (10 CFR 50 Appendix H) [92]. Kewaunee used both irradiated
and unirradiated KJc data for the limiting circumferential weld wire heat, 1P3571. Kewaunee
was the first application to the NRC using CC N-629. For this reason, Kewaunee undertook
extensive efforts to address NRC staff questions concerning the applicability of the Master Curve
to RPV material characterization in general [46]. Kewaunee’s initial submittal to the NRC
argued that the irradiated RTNDT-based index temperature for 1P3571 could be reduced by
26.7°C. Staff concerns regarding use of KJc data from pre-cracked Charpy specimens, use of
sister plant data, and material variability changed this value to 5°C in the final safety evaluation
report.
Table 6-1
Summary of unirradiated RTT0 value for various Linde 80 weld wire heats.

Basis Linde 80 Heat ID Unirradiated RTT0 [oF] Margin: σ1 [oF]

406LL44 -98.0 11.6

71249 -53.5 12.8


72105 -31.1 13.7

821T44 -84.2 9.6

Master Curve 299L44 -74.3 12.8


Toughness Data
72442 -33.2 12.2

72445 -72.5 12.0

61782 -58.5 15.4


Generic Value for -48.6** 18.0
Other Heats
RTNDT Based Values All heats -7 to +10 17

** Includes 20oF addition to address NRC’s concerns with the equivalency of bounding provided by a
RTT0 indexed KIc curve.

6.3 Currently Foreseen Uses of the CC N-830-1 Suite of Fracture Toughness


Models
As mentioned in the introduction to this Chapter, because CC N-830-1 provides a complete
description of the fracture toughness of ferritic steel from the lower shelf through the upper shelf
it can, in principle, be find use in any Code or Regulatory activity needing fracture toughness as
an input value. The following list highlights only those topics currently being pursued.
• Revision 2 of CC N-830: As mentioned previously in this report, Revision 1 of CC N-830 is
a step in the evolution and refinement of the methods provided by this CC. Work is already
underway within the WGFE on a Revision 2 of the Code Case [93]. Revision 2 will adopt
partial structural factors (PSFs) to address the impact of uncertainties on the three key
variables in a fracture safety assessment: fracture toughness, flaw size, and stress. Use of
PSFs improves on the Revision 1 approach, which assigns a structural factor only to fracture
toughness, in that it explicitly and transparently addresses the uncertainties associated with

6-4
Potential Code/Regulatory Applications of CC N-830-1

each variable, and allows adjustment of the values of the structural factors used to directly
reflect the state of knowledge concerning each variable. PSFs also provide a linkage
between deterministic and probabilistic assessments (see [91] for details).
• Nozzles: The PWROG has an effort underway entitled “Pressurized Water Reactor, Reactor
Pressure Vessel Appendix G Margins” [94]. The purpose of this project is to address an
issue identified in NRC Regulatory Issue Summary (RIS) 2014-11 [95], which states that
“All licensees should ensure that P-T limits sufficiently address all ferritic materials of the
reactor vessel, including the impact of structural discontinuities, and address the impact of
neutron fluence accumulation in accordance with the requirements of 10 CFR Part 50,
Appendix G.” This RIS identified the need for fracture toughness data for the RPV nozzle
course, which in some cases is not available in construction records. The PWROG effort has
therefore proposed to use a generic RTT0 value for nozzle-course forgings, and to also
account for the systematic through-thickness variation of RTT0 known to be responsible for
greater fracture toughness near the inner-diameter surface of the nozzle where flaws are
postulated to exist.
• PWROG Direct Fracture Toughness Initiative: The PWROG has an effort underway
entitled “Transitioning RV Integrity to Direct Fracture Toughness, Phase 1” [96]. The
topical report from this phase, which will be submitted to the NRC for review and approval,
will propose a method to use irradiated fracture toughness data to improve or demonstrate
margin in pressure-temperature limit curves. The report will address the generation of
irradiated T0 data, methods to account for material variability, and, because the irradiated T0
data will be generated in test reactors, a discussion of and, if necessary, and accounting for
flux effects.
• EPRI Effort for Multi-Data T0 Estimation: The PWROG proposed plan to measure T0 for
RPV materials will provide a large database of BOL T0 values as well as irradiated T0 values.
However, there are many critical materials for which archival test material is unavailable. In
addition, because testing of irradiated specimens is expensive, not all available archival
materials will be tested. These factors necessitate the development of alternative methods for
estimating T0 values for these materials. EPRI is supporting a program to define the
correlations necessary to enable estimation of T0 from any combination of material toughness
properties already available through previous testing, including NDT, T30, USE, KIa,
instrumented Charpy, JIc and J0.1, and strength properties. Correlations between some of
these properties have been clearly demonstrated [11, 12, 14, 64, 73, 77, 97, 98]. This
program focuses on establishing correlations between NDT, T30 and USE, and to link these
approximate measures of toughness to T0 to enable estimation of any toughness parameter
from some combination of other toughness parameters.
• EPRI T0-based Embrittlement Trend Curve: In the development of large-scale,
probabilistic models to assess fracture safety of RPVs, it is desired to move towards direct
use of fracture toughness properties instead of a correlative approach. However, this
requires development of a fracture toughness-based embrittlement trend curve (ETC), and
wide-spread acceptance of fracture toughness models. The current CC N-830-1 initiative to
implement more rigorous fracture toughness models into the ASME Code provides best
estimate models of fracture toughness in all fracture mode regimes, but it does not include a
T0-based ETC. The largest impediment to development and acceptance of a T0-based ETC

6-5
Potential Code/Regulatory Applications of CC N-830-1

is limitations on the T0 data available for materials for which Charpy data is also available
for the same or similar exposure conditions that would enable direct comparison of the
current Charpy-based ETC models. The T0 ETC program compiled a large database
containing both CVE and KJc data that was used to define both a TCVE ETC [99] and a T0
ETC [100]. Most of the fracture toughness data available to use in ETC development was
research data with varying exposures, environments and heat treatments. A T0-based ETC
was developed that was shown to fit all the available fracture toughness data with little bias;
however, the scatter about the mean was almost double that of Charpy-based ETCs. The
cause of the larger-than-expected scatter observed in the fracture toughness data was found
to be lack of descriptor fields in the fracture toughness database relative to the more
prescriptively acquired Charpy surveillance data. This is due, in large part, to the fact that
much of the fracture toughness data was developed as part of research programs whose
primary goal was in understanding a variety of effects on the measure of fracture toughness
and not specifically in monitoring radiation degradation. For this reason, the fracture
toughness database contains toughness data measured using many different specimen
configurations, test apparatus, and many exposure conditions including some post-
irradiation annealing, the details of which were not always captured in the database fields.
Comparison of the T0-based ETC with 4 Charpy-based ETCs supports the 1:1 correlation
between Charpy shift and fracture toughness shift that is often cited [100]. Due to the
inherent scatter observed in the fracture toughness data, a 1:1 correlation on the uncertainty
is harder to support.

6-6
7
SAMPLE PROBLEMS AND RESULTS

7.1 Introduction
Two sample problems were developed and solved within the WGFE to support the development
of CC N-830-1. The first sample problem was conducted in late 2015 and focused exclusively
on calculation of the allowable toughness values as outlined by the CC, with the objective of
ensuring that the CC provided sufficient clarity to enable ready implementation of the equations
contained therein. The Sample Problem 1 statement is provided in Appendix A of this report.
A second sample problem was conducted in 2016 with two objectives:
5. Calculation of allowable toughness values (same as in the first sample problem).
6. Calculation of allowable flaw sizes from these allowable toughness values.
While CC N-830-1 pertains only to determination of allowable toughness, calculation of the
allowable flaw size was performed in Sample Problem II to enable comparison of the effects of
the implementation of the CC N-830-1 allowable toughness equations with calculations made
using the current Appendix A allowable toughness metrics.
The second sample problem fully contains the first sample problem; its conduct and results were
documented in a technical paper presented at the 2017 ASME Pressure Vessel and Piping
Conference [79]. This chapter provides a synopsis of the full results described in detail in [79].
The full Sample Problem 2 statement is provided in Appendix B of this report. Since CC N-830-
1 has not yet been adopted by Section XI of the ASME Code and is still subject to change, the
version used in the conduct of Sample Problem 2 is included as Appendix C to provide context.
The following individuals solved the two sample problems:
• Yil Kim GE POWER
• Mark Kirk U.S. NRC
• Darrell Lee BWXT
• Cheng Liu & Steven Xu Kinectrics
• Do Jun Shim & Gary Stevens Structural Integrity Associates, Inc.

7.2 The Sample Problem


The second sample problem involved the calculation of allowable toughness values using the
following three methods:
Method 1: Using the toughness equations in CC N-830-1 at their 1st percentile lower bounds.
Method 2: Using the mean or median toughness values from CC N-830-1 divided by √10.

7-1
Sample Problems and Results

Method 3: Using the toughness equations in ASME Section XI Appendix A divided by √10.
Methods 1 and 2 represent allowable toughness values being considered by the WGFE for use in
CC N-830-1, while Method 3 represents the current ASME Code Appendix A calculation. For
Method 3, since the ASME Code does not specify a method to estimate J0.1 values for different
steels, the participants were allowed to specify the method they used (e.g., Regulatory Guide
1.161), or simply state that the results were “undefined” according to Section XI Appendix A.
For each method nine material Case IDs were defined, as outlined in Table 7.1. These cases
explore the full range of epistemic uncertainty inherent to the relationship between RTNDT and T0
(see Figure 5-1).
Once allowable toughness values were calculated using these three Methods and nine Case IDs,
the participants were asked to calculate allowable flaw sizes for a typical PWR at operating
pressure and under isothermal conditions at 50 °C (see Appendix B for full details). This single
allowable flaw size calculation does not represent all conditions of interest; it was performed
only to give WGFE members a sense of how changes in allowable toughness scale to changes in
allowable flaw size. Also, it should be noted that CC N-830-1 does not specify or propose to
change the Code’s allowable flaw size calculations. Thus, the allowable flaw size part of the
sample problem is not directly pertinent to the proposed revision of the Code Case; it was
performed for the information of WGFE members.
Table 7-1
Material Properties for use in Appendix A and Proposed Code Case N-830-1 Sample
Problem 2.

Low Transition Medium Transition High Transition


CDF Actual
Temperature Temperature Temperature
%-ile RTNDT-T0
Case RTNDT T0 Case RTNDT T0 Case RTNDT T0
RTNDT-T0 o
C
ID o
C o
C ID o
C o
C ID o
C o
C
th
5 10 L-5 -20 -10 M-5 100 90 H-5 180 170
th
50 48 L-50 -20 -68 M-50 100 52 H-50 180 132
95th 92 L-95 -20 -112 M-95 100 8 H-95 180 88

7.2.1 Allowable Toughness Values


Once allowable toughness values were calculated using these three Methods and nine Case IDs,
the participants were asked to calculate allowable flaw sizes for a typical PWR at operating
pressure and under isothermal conditions at 50 °C (see Appendix B for full details). This single
allowable flaw size calculation does not represent all conditions of interest; it was performed
only to give WGFE members a sense of how changes in allowable toughness scale to changes in
allowable flaw size. Also, it should be noted that CC N-830-1 does not specify or propose to
change the Code’s allowable flaw size calculations. Thus, the allowable flaw size part of the
sample problem is not directly pertinent to the proposed revision of the Code Case; it was
performed for the information of WGFE members.

7-2
Sample Problems and Results

7.2.2 Allowable Flaw Size Values


As documented in [79], achieving agreement in allowable flaw depth values among the six
participants was more difficult. While all six participants used formulae outlined in the Code
and numerical methods permitted by standard Code practice, the Code itself is not fully
prescriptive in how allowable flaw depths should be calculated. Differences between the
calculation methods used by different participants were observed to contribute to the differences
in results as follows:
• Method used to estimate through-wall stresses (thick wall vs. thin wall solutions)
• Whether or not a plastic zone correction was performed.
• Whether the possibility of failure at the point where the semi-elliptic flaw intersects the
inner-diameter of the RPV was considered in addition to the location at the deepest point of
the semi-elliptic flaw.
• The method used to fit the stress distribution and then, based on this fit, calculate the
KAPPLIED value.
• The level of discretization used for both temperature and flaw depth.
A full assessment of the effects of these differences appears in [79]. For conditions of high
allowable toughness, the resultant differences in allowable flaw depth were small, generally 5-
10% between participants. However, for low allowable toughness conditions greater differences
(30-40%) in allowable flaw depth occurred.
When comparing allowable flaw depths calculated from current Code methods (Method 3) and
potential Code Case methods (Methods 1 and 2), current Code estimates of allowable flaw depth
were more conservative (that is: smaller) than estimates based on the candidate CC methods.
This was mostly due to the generally-conservative bias of the Code’s RTNDT –indexed, KIc curve
approach (as an example, see Figure 5-5). The conservatism inherent to current Code methods
was found to vary considerably depending on the material condition (Case-ID), ranging from
modest increases (e.g., 1.1-2x for Case H-5) to quite large increases (e.g., ≈15x for Case M-95).
The cause of this significant effect of material condition on the conservatism inherent to current
Code estimates can be illustrated by comparing the various allowable toughness curves with
actual fracture toughness data for RPV steels, such as those appearing in Figures 5-2 through 5-
7. Both the current Code and the candidate CC N-830-1 allowable toughness methods produce
lower-bounding representations of the data, but to greatly varying degrees of conservatism
depending on the temperature at which the flaw evaluation is performed. The differences in
vertical extent between the bounding curves and the toughness data at any given temperature
cause the differences in allowable flaw depth. In addition, the current Code definition of an
allowable toughness curve for normal operating conditions (SF=√10) drives the allowable
toughness value below 20 MPa√m at temperatures below T-T0 = +50°C. However, neither
empirical nor theoretical evidence exists to support such low toughness values in RPV steels.
It is not possible to achieve consistent conservatism across the conditions occurring in the
operating fleet using the Code’s RTNDT-indexed, KIc curve approach due to the correlative and
intentionally biased nature of RTNDT, and because the temperature dependence of the KIc curve
does not match the temperature dependence of all currently-available fracture toughness data for
RPV steels. These issues are easily addressed by the candidate CC approaches for normal

7-3
Sample Problems and Results

operating/upset conditions and emergency/faulted conditions, both of which adopt an index


temperature (T0) defined by actual fracture toughness data and a temperature dependence defined
by those data.

7-4
8
SUMMARY AND CONCLUSIONS

The technical bases for the fracture toughness models contained in ASME CC N-830-1 are
presented in this report. The suite of best estimate fracture toughness models provides a
complete description of fracture toughness crack initiation and arrest behavior from lower shelf,
through transition, to ductile upper shelf regimes for all ferritic steels.
The best estimate models used for CC N-830-1 are based on updated techniques and available
data, sound physical bases, and extensive empirical evaluations that collectively promote
confidence in their use for flaw assessment following Nonmandatory Appendix A of ASME
Section XI and similar methods. These models are appropriate for use in both deterministic and
probabilistic assessments, as each model describes the full distribution in values about the mean
for any temperature and material condition.
Equations are presented for each fracture toughness model that allow an analyst to determine any
percentile value of interest for any of the fracture toughness parameters KJc, KIa, JIc, J0.1, and J-R.
Specific values of these parameters may be used in deterministic assessments, or the entire
distributions may be sampled for use in probabilistic assessments. Collectively, these fracture
toughness models provide a consistent, best-estimate representation of ferritic steel fracture
toughness behavior, including uncertainties, to allow for quantitative fracture toughness
assessments that ensure the safety of nuclear (and other) ferritic components.
Sample problem assessments were performed to ensure the adequacy of the technical content in
CC N-830-1, to verify the accuracy of the CC content, and to ensure that the CC could be applied
by knowledgeable engineers to produce reasonable and reliable results in typical flaw
evaluations.
This document is provided to the ASME Section XI Working Group on Flaw Evaluation for their
use to support moving forward with direct implementation of best-estimate fracture toughness
curves.

8-1
9
REFERENCES

1. ASTM E208, “Standard Test Method for Conducting Drop-Weight Test to Determine Nil-
Ductility Transition Temperature of Ferritic Steels,” Annual Book of ASTM Standards, Vol.
03.01, American Society for Testing and Materials, West Conshohocken, PA, 2002.
2. ASTM E23, “Standard Test Methods for Notched Bar Impact Testing of Metallic Materials,”
Annual Book of ASTM Standards, Vol. 03.01, American Society for Testing and Materials,
West Conshohocken, PA, 2002.
3. ASTM E 399 “Standard Test Method for Linear-Elastic Plane-Strain Fracture Toughness KIc
of Metallic Materials,” Annual Book of ASTM Standards, Vol. 03.01, American Society for
Testing and Materials, West Conshohocken, PA, 2013.
4. ASME NB-2331, 1998 ASME Boiler and Pressure Vessel Code, Rules for Construction of
Nuclear Power Plants, Division 1, Subsection NB, Class 1 Components.
5. PVRC Ad Hoc Group on Toughness Requirements, “PVRC Recommendations on
Toughness Requirements for Ferritic Materials.” Welding Research Council Bulletin No.
175, August 1972.
6. Marston, T.U., “Flaw Evaluation Procedures, Background and Application of ASME Section
XI Appendix A,” EPRI Report NP-719-SR, Electric Power Research Institute, 1978.
7. ASME Boiler and Pressure Vessel Code, Rules for In-service Inspection of Nuclear Power
Plants, Section XI, Appendix A., “Analysis of Flaws” 2013.
8. ASME Boiler and Pressure Vessel Code, Section XI, “Rules for In-service Inspection of
Nuclear Power Plant Components,” Appendix K, “Assessment of Reactor Vessels with Low
Upper Shelf Charpy Impact Energy Levels,” 2013.
9. Code of Federal Regulations, Title 10, Part 50, Appendix G, “Fracture Toughness
Requirements.”
10. Regulatory Guide 1.161, “Evaluation of Reactor Pressure Vessels with Charpy Upper-Shelf
Energy of Less Than 50 ft-lb,” United States Nuclear Regulatory Commission, June 1995.
11. Materials Reliability Program: Implementation Strategy for Master Curve Reference
Temperature, T0 (MRP-101), EPRI, Palo Alto, CA, and U.S. Department of Energy,
Washington, DC: 2004. 1009543.
12. Kirk, M., M. Erickson, G. Stevens, W. Server, and R. Cipolla, “Assessment of Fracture
Toughness Models for Ferritic Steels Used in Section XI of the ASME Code Relative to
Current Data-Based Models,” PVP2014-28540, Proceedings of PVP2014, 2014 ASME
Pressure Vessels and Piping Division Conference, July 20-24, 2014, Anaheim, CA, USA.

9-1
References

13. Code of Federal Regulation 10CFR Part 50.61, “Fracture Toughness Requirements for
Protection Against Pressurized Thermal Shock Events.”
14. EricksonKirk, M.A., and M. T. EricksonKirk, “The Relationship Between Transition and
Upper-Shelf Fracture Toughness of Ferritic Steels,” Fatigue & Fracture of Engineering
Materials & Structures, 29, March 2006, 672–684.
15. ASTM E1921-97, “Standard Test Method for Determination of Reference Temperature, T0,
for Ferritic Steels in the Transition Range,” ASTM International, West Conshohocken, PA,
2002, www.astm.org.
16. ASME Boiler and Pressure Vessel Code Case N-629, “Use of Fracture Toughness Test Data
to Establish Reference Temperature for Pressure Retaining Materials, Section XI, Division
1,” 1999.
17. ASME Boiler and Pressure Vessel Code Case N-631, “Use of Fracture Toughness Test Data
to Establish Reference Temperature for Pressure Retaining Materials Other Than Bolting for
Class 1 Vessels Section III, Division 1, 1999.
18. ASME Boiler and Pressure Vessel Code Case N-830 (BC 09-182), Direct Use of Master
Fracture Toughness Curve for Pressure Retaining Materials for Vessels of a Section XI,
Division 1, 2014.
19. ASME Boiler and Pressure Vessel Code Case N-830-1 (14-1073), Direct Use of Master
Fracture Toughness Curve for Pressure Retaining Materials for Vessels of a Section XI,
Division 1, 2017 DRAFT.
20. Wallin, K., “The Scatter in KIc Results,” Engineering Fracture Mechanics, 19(6), pp. 1085-
1093, 1984.
21. Wallin, K., Saario, T., and Törrönen, K., “Statistical Model for Carbide Induced Brittle
Fracture in Steel,” Metal Science, Vol. 18, January 1984, 13-16.
22. Wallin, K., “Irradiation Damage Effects on the Fracture Toughness Transition Curve Shape
for Reactor Vessel Steels,” Int. J. Pres. Ves. & Piping, 55, pp. 61-79, 1993.
23. Wallin, K., “Master Curve Analysis of Ductile to Brittle Region Fracture Toughness Round
Robin Data: The ‘EURO’ Fracture Toughness Curve,” VTT Manufacturing Technology,
VTT Publication 367, 1998.
24. Kirk, M.T., “The Technical Basis for Application of the Master Curve to the Assessment of
Nuclear Reactor Pressure Vessel Integrity,” NRC Technical Note, ADAMS ML093540004.
25. ASME Section XI Division I, Code Case N-851, “Alternative Method for Establishing the
Reference Temperature for Pressure Retaining Materials, Adopted Nov. 2014.
26. ASME Section III, Division 1, Record # 11-1383R1.
27. Zerilli, F.J., and Armstrong, R.W., “Dislocation-Mechanics-Based Constitutive Relations for
Material Dynamics Calculations,” Journal of Applied Physics, Vol. 61, No. 5, March 1987,
1816-1825.

9-2
References

28. Natishan, M. and Kirk, M., “A Micromechanical Evaluation of the Master Curve,” Fatigue
and Fracture Mechanics, 30th Volume, ASTM STP-1360, K. Jerina and P. Paris, Eds.,
American Society for Testing and Materials, 1998.
29. Natishan, M.E., Wagenhoefer, M., and Kirk, M.T., “Dislocation Mechanics Basis and Stress
State Dependency of the Master Curve,” Fracture Mechanics, 31st Symposium, ASTM STP
1389, K. Jerina and J. Gallagher, Eds., American Society for Testing and Materials, 2000.
30. Natishan, M. and Kirk, M., “A Physical Basis for the Master Curve,” Proc. of the 1999
ASME Pressure Vessel and Piping Conference, ASME, July 1999.
31. Curry D.A., and Knott J.F., “The Relationship Between Fracture Toughness and
Microstructure in the Cleavage Fracture of Mild Steel,” Metal Science, 10, pp. 1-6, 1976.
32. Curry D.A., and Knott J. F., “Effects of Microstructure on Cleavage Fracture Stress in Steel,”
Metal Science, 12, pp. 511-514, 1978.
33. Curry, D.A. and Knott, J.F., “Effect of Microstructure on Cleavage Fracture Toughness of
Quenched and Tempered Steels,” Metal Science, 13, pp. 341-345, 1979.
34. Knott, J.F., J. Iron Steel Inst., 204, p. 104, 1966.
35. McMahon, C.J. and Cohen, M., “Initiation of Cleavage in Polycrystalline Iron,” Acta
Metalla., Vol. 13, pp. 591-604, 1965.
36. Smith, E., Physical Basis of Yield and Fracture, Conference Proceedings, Institute of Physics
and the Physical Society, London, p. 36, 1966.
37. Smith, E., “Cleavage Fracture in Mild Steels,” Int. J. Fract. Mech., Vol. 4, pp. 131-145, 1968.
38. Ritchie, R.O., Knott, J., and Rice, J., “On the Relationship Between Critical Tensile Stress
and Fracture Stress in Mild Steels,” Journal of the Mechanics and Physics of Solids, Volume
21, Issue 6, 1973, pp. 395-410.
39. Curry, D.A., “Comparison Between Two Models of Cleavage Fracture,” Met. Sci., pp.78-80,
Feb. 1980.
40. Wallin, K., “Fracture Toughness of Engineering Materials, Estimation and Application,”
EMAS Publishing, ISBN: 978-0-9552994-6-9, 2011.
41. Wallin, K., “The Size Effect in KIc Results,” Engineering Fracture Mechanics, 22, pp. 149-
163, 1985.
42. Wallin, K., “Statistical Modelling of Fracture in the Ductile to Brittle Transition Region,”
Defect Assessment in Components – Fundamentals and Applications, ESIS/EGF9, J.G.
Blauel and K.-H. Schwalbe, Eds., pp. 415-445, 1991.
43. Wallin, K., “Statistical Aspects of Constraint with Emphasis on Testing and Analysis of
Laboratory Specimens in the Transition Region,” Constraint Effects in Fracture, ASTM STP-
1171, E.M. Hackett, K.-H. Schwalbe, and R.H. Dodds Eds., American Society for Testing
and Materials, 1993.
44. “Application of Master Curve Fracture Toughness Methodology for Ferritic Steels,” EPRI-
TR-108390 Revision 1, 1999.

9-3
References

45. Kirk, M., Lott, R., Kim, C., and Server, W., “Empirical Validation of the Master Curve For
Irradiated And Unirradiated Reactor Pressure Vessel Steels,” Proceedings of the 1998
ASME/JSME Pressure Vessel and Piping Symposium, July 26-30, 1998, San Diego,
California, USA.
46. Lott, R.G., Kirk, M.T., and Kim, C.C., “Master Curve Strategies for RPV Assessment,”
Westinghouse Electric Company, WCAP-15075, September 1998, Available on the USNRC
website at Legacy ADAMS 9811240260 and at ADAMS ML111861647.
47. W. A. VanDerSluys, C. L. Hoffmann, K. K. Yoon, D. E. Killian and J. B. Hall, “Fracture
Toughness Master Curve Development: Fracture Toughness of Ferritic Steels and ASTM
Reference Temperature (To),” Welding Research Council, WRC-457, December 2000.
48. W.A. VanDerSluys, C.L. Hoffmann, K.K. Yoon, W.L. Server, R.G. Loft, S. Rosinski, M.T.
Kirk, S. Byrne and C.C. Kim, “Fracture Toughness Master Curve Development: Application
of Master Curve Fracture Toughness Methodology for Ferritic Steels,” Welding Research
Council, WRC-458, January 2001.
49. W.A. VanDerSluys, C.L. Hoffman, W.L. Server, R.G. Lott, M.T. Kirk and C.C. Kim,
“Fracture Toughness Master Curve Development: Strategies for RPV Assessment,” Welding
Research Council, WRC-459, February 2001.
50. J. Heerens, D. Hellmann, “Development of the Euro fracture toughness dataset,” Engineering
Fracture Mechanics 69 (2002) 421–449.
51. J. Heerens, R.A. Ainsworthb, R. Moskovicc, K. Wallin, “Fracture toughness characterisation
in the ductile-to-brittle transition and upper shelf regimes using pre-cracked Charpy single-
edge bend specimens,” International Journal of Pressure Vessels and Piping 82 (2005) 649–
667.
52. Rathbun, H. J., Odette, G R, He, M. Y., “On the Size Scaling of Cleavage Toughness in the
Transition: A Single Variable Experiment and Model Based Analysis,” Proceedings of the
10th International Conference on Fracture, Honolulu, Hawaii – 2001.
53. Rathbun, H. J., G. R. Odette, M. Y. He, G. E. Lucas, and T. Yamamoto, “Size Scaling of
Cleavage Toughness in the Transition: A Single Variable Experiment and Model-Based
Analysis,” United States Nuclear Regulatory Commission, NUREG.CR-6790, August 2002.
54. Leax, T.R., “Temperature Dependence and Variability of Fracture Toughness in the
Transition Regime for A508 Grade 4N Pressure Vessel Steel,” Journal of ASTM
International, Volume 3, Issue 1 (January 2006).
55. Wallin, K., “The Elusive Temperature Dependence of the Master Curve,” Proceedings of the
13th International Conference on Fracture, June 2013, Beijing, China.
56. IAEA TRS-429, “Guidelines for Application of the Master Curve Approach to Reactor
Pressure Vessel Integrity in Nuclear Power Plants,” International Atomic Energy Agency,
Vienna Austria, 2005.
57. Pisarski, H.G. and Wallin, K., “Application and Verification of the SINTAP Fracture
Toughness Estimation Procedure for Welds and Parent Metals,” Proceedings of the 18th

9-4
References

International Conference on Offshore Mechanics and Artic Engineering, OMAE99/MAT-


2043, ASME, July 1999.
58. EricksonKirk, M.T., M. A. EricksonKirk, Charles Roe, and Xian J. Zhang, “Use of Large
Databases to Identify Trends in the Behavior of Ferritic Steels,” PVP2009-77788,
Proceedings of PVP2009, 2009 ASME Pressure Vessel and Piping Division Conference, July
27-31, 2009, Prague, Czech Republic.
59. EricksonKirk,M.A., M. EricksonKirk, and T. Williams, “The Interrelationships of KIa, KIc,
And JIc, and the Implications of These Relationships on Use of Fracture Models Over the
Ranges of Hardening Observed in Ferritic Steels,” PVP2006-ICPVT-11-93651, Proceedings
of PVP2006-ICPVT-11: ASME 2006 Pressure Vessels and Piping/ICPVT-11 Conference,
July 23-27, 2006, Vancouver BC, Canada.
60. Wallin, K., and Rintamaa, R., “Master Curve Based Correlation Between Static Initiation
Toughness KIc and Crack Arrest Toughness KIa,” Proceedings of the 24th MPA-Seminar,
Stuttgart, October 8 and 9, 1998.
61. Kirk, M. T., Natishan, M. E., and Wagenhofer, M., “A Physics-Based Model for the Crack
Arrest Toughness of Ferritic Steels,” Fatigue and Fracture Mechanics, 33rd Volume, ASTM
STP-1417, W. G. Reuter, and R. S. Piascik, Eds., American Society for Testing and
Materials, West Conshohocken, PA, 2002.
62. H. Hein, E. Keim, H. Schnabel, J. Barthelmes, C. Eiselt, F. Obermeier, J. Ganswind, and M.
Widera, “Final Results from the CARINA Project on Crack Initiation and Arrest of Irradiated
German RPV Steels for Neutron Fluences in the Upper Bound,” 26th ASTM Symposium on
Effects of Radiation on Nuclear Materials, June 12-13, 2013, Indianapolis, IN (USA).
63. ASTM E1221-10, “Standard Test Method for Determining Plane-Strain Crack-Arrest
Fracture Toughness, KIa, of Ferritic Steels.” Annual Book of ASTM Standards, Vol. 03.01,
American Society for Testing and Materials, West Conshohocken, PA, 2010.
64. EricksonKirk, M.A., and M.T. EricksonKirk, “An Upper-Shelf Fracture Toughness Master
Curve for Ferritic Steels,” International Journal of Pressure Vessels and Piping, 83 (2006)
571-583.
65. Kirk, M.T., M. Erickson, R. Link, C. Roe, “Large Database Assessment of the Upper Shelf
Master Curve,” Proceedings of PVP2017, ASME Pressure Vessels and Piping Division
Conference, July 16-20, 2017, Honolulu Hawaii, USA.
66. Wagenhofer, M., “Modeling the transition region fracture behavior of ferritic steels.” Ph.D.
Dissertation, University of Maryland, College Park, MD, 2002.
67. Heerens, J. and D. Hellman, “Development of the Euro Fracture Toughness Dataset,”
Engineering Fracture Mechanics, Vol. 69, 2002, pp. 419-449.
68. Wallin, K., “Master Curve Analysis of the “Euro” Fracture Toughness Dataset,” Engineering
Fracture Mechanics, Vol. 69, 2002, pp. 451-481.
69. Heerens, J, M. Pfuff, D. Hellmann and U. Zerbst, “The Lower Bound Toughness Procedure
Applied to the Euro Fracture Toughness Dataset,” Engineering Fracture Mechanics, Vol. 69,
2002, pp. 483-495.

9-5
References

70. Neale, B.K., “An Assessment of the Fracture Toughness in the Ductile-to-Brittle-Transition
Regime Using the Euro Fracture Toughness Dataset,” Engineering Fracture Mechanics, Vol.
69, 2002, pp. 495-509.
71. Moskovic, R., “Modelling of Fracture Toughness Data in the Ductile-to-Brittle Transition
Temperature Region by Statistical Analysis,” Engineering Fracture Mechanics, Vol. 69,
2002, pp. 511-530.
72. ASTM E1820-01 (2002) Standard test method for measurement of fracture toughness.
ASTM. Annual Book of ASTM Standards, Vol. 03.01. American Society for Testing and
Materials, West Conshohocken, PA.
73. Kirk, M.T., H. Hein, M. Erickson, W. Server, G. Stevens, “A Fracture-Toughness Based
Transition Reference Temperature for Use in the ASME Code With the Crack Arrest (KIA)
Curve,” PVP 2014-28311, 2014 ASME Pressure Vessels and Piping Conference, July 20-24,
2014, Anaheim, CA, USA.
74. Eason, E., Wright, J., and Nelson, E., “Multivariable Modeling of Pressure Vessel and Piping
J-R Data,” United States Nuclear Regulatory Commission, NUREG/CR-5729, 1991.
75. Wallin, K., “Low-cost J-R Curve Estimation Based on Charpy Upper Shelf Energy,” Fatigue
and Fracture of Engineering Materials and Structures, 24, 537-549, 2001.
76. Matsuzawa, H., and Osaki, T., “Fracture Toughness of Highly Irradiated Pressure Vessel
Steels in the Upper Shelf Temperature Region,” Proceedings of PVP2006-ICPVT-11,
Vancouver, British Columbia, Canada, July 23-27, 2006 (PVP2006-ICPVT11-93032).
77. Kirk, M.T., M.A. Erickson, G. Stevens, “Models of the Temperature Dependence and Scatter
in J-R and J0.1 for Ferritic Reactor Pressure Vessel Steels,” PVP2015-45253, Proceedings of
PVP2015 2015 ASME Pressure Vessels and Piping Division Conference, July 19-23, 2015,
Boston, Massachusetts, USA.
78. Nondestructive Evaluation: Probabilistic Analysis of Performance Demonstration Ultrasonic
Flaw Detection and Through-Wall Sizing Results for Reactor Pressure Vessel Inspections.
EPRI, Palo Alto, CA: 2012. 1025787.
79. M. Kirk, S. Xu., C. Liu, M. Erickson, Y. Kim, D. Lee, D.J. Shim, and G. Stevens., “Solution
of a Sample Problem Related to Revision 1 of Code Case N-830,” 2017 ASME Pressure
Vessel and Piping Conference, July 16-20, 2017, Waikoloa Hawaii, PVP2017-66150.
80. Pierre Dulieu and Valéry Lacroix, “Structural Integrity Assessment of Doel 2 and Tihange 2
RPVs Affected by Hydrogen Flakes: Refined X-FEM Analyses,” 2016 ASME Pressure
Vessel and Piping Conference, Vancouver, British Columbia, PVP2016-63766.
81. Valéry Lacroix, Pierre Dulieu, and Damien Couplet, “Alternative Characterization Rules for
Quasi-Laminar Flaws,” 2014 ASME Pressure Vessel and Piping Conference, Anaheim,
California, PVP2014-28200.
82. BS 7910:2013, “Guide to Methods for Assessing the Acceptability of Flaws in Metallic
Structures,” British Standards Institute, 2013.
83. API 579-1, ASME FFS-1, “Fitness for Service,” June 2016.

9-6
References

84. Risk-Informed Method to Determine ASME Section XI Appendix G Limits for Ferritic Reactor
Pressure Vessels - An Optional Approach Proposed for ASME Section XI Appendix G. MRP-
250 and BWRVIP-215NP. EPRI, Palo Alto, CA: 2009. 1016600.
85. ASME Boiler and Pressure Vessel Code, Section XI, “Rules for In-service Inspection of
Nuclear Power Plant Components, Appendix E., “Evaluation of Unanticipated Operating
Events,” 2013.
86. ASME Boiler and Pressure Vessel Code, Section XI, “Rules for In-service Inspection of
Nuclear Power Plant Components,” Appendix G, “Fracture Toughness Criteria for Protection
Against Failure,” 2013.
87. Code of Federal Regulations, 10 CFR 50.61, “Fracture Toughness Requirements for
Protection Against Pressurized Thermal Shock Events.”
88. K.K. Yoon, “Fracture Toughness Characterization of WF-70 Weld Metal,” Report to the
B&W Owners Group Materials Committee, BAW-2202, September 1993.
89. D. E. McCabe, R. K. Nanstad, S. K. Iskander, R. L. Swain, “Unirradiated Material Properties
of Midland Weld WF-70,” United States Nuclear Regulatory Commission Report
NUREG/CR-6249, October 1994.
90. “Initial RTNDT of Linde-80 Weld Materials,” Report to the PWR Owners Group, BAW-2308
Rev. 2A, March 2008.
91. NRC Safety Evaluation Report on Kewaunee Master Curve Submittal, Letter of 1st May 2001
from Lamb to Reddemann, ADAMS ML011210180.
92. Code of Federal Regulations 10 CFR 50 Appendix H, “Reactor Vessel Material Surveillance
Program Requirements,” Published by the Office of the Federal Register, National Archives
and Records Administration, Washington, D.C., 2000.
93. Erickson, M.A., and M. T. Kirk, “Development of a Partial Structural Factor Approach for
Direct Fracture Toughness Implementation into the ASME Boiler and Pressure Vessel
Code,” 2017 ASME Pressure Vessel and Piping Conference, July 16-20, 2107, Waikoloa
Hawaii, PVP2017-66148.
94. Chris Koehler, Heather Malikowski, Brian Hall, and Justin Webb, “PWR RPV Nozzle
Appendix G Margins,” presentation at 19 January 2016 NRC meeting, ADAMS ML
16021A002.
95. NRC Regulatory Issue Summary 2014-11, “Information on Licensing Applications for
Fracture Toughness Requirements for Ferritic Reactor Coolant Pressure Boundary
Components,” October 14, 2014, ADAMS ML14149A165.
96. J. Brian Hall, Elliot Long, Ben Mays, Heather Malikowski, and Chris Koehler, “Plan for
Transitioning RV Integrity to Direct Fracture Toughness,” presented at the International
Light Water Reactors Material Reliability Conference, Chicago, August 2016.
97. Kirk, M., and M. Erickson, “Assessment of the Temperature Dependence of Ferritic Steel
Fracture Toughness on or Near the Lower Shelf,” Proceedings of PVP2015 2015 ASME
Pressure Vessels and Piping Division Conference, July 19-23, 2015, Boston, Massachusetts,
USA, PVP2015-45850.

9-7
References

98. Kirk, M.T., M. Erickson, G. Stevens, H. Gustin and W. Server, “Options for Defining the
Upper Shelf Transition Temperature (Tc) for Ferritic Pressure Vessel Steels,” Proceedings of
PVP2015, 2015 ASME Pressure Vessels and Piping Division Conference, July 19-23, 2015,
Boston, Massachusetts, USA, PVP2015-45307.
99. Materials Reliability Program: Developing an Embrittlement Trend Curve Using the Charpy
“Master Curve Transition Reference Temperature (MRP-289). EPRI, Palo Alto, CA: 2011,
1020703.
100. Materials Reliability Program: Development of a T0 –Based Embrittlement Trend Curve
and Comparison With the Charpy Master Curve Embrittlement Trend Curve (MRP-389).
EPRI, Palo Alto, CA: 2014. 3002003040.
101. Kirk, M.T., “Evaluation and/or Validation of the J-R Curve Prediction Model Proposed
for use in Revision 1 to ASME Code Case N-830,” presentation at the International Group on
Radiation Damage Mechanisms (IGRDM) 20th Meeting, Santiago de Compostela, Spain,
October 2017. (ADAMS ML17271A110).

9-8
A
SAMPLE PROBLEM 1 STATEMENT

Sample Problem Statement


Code Case N-830-1, “Direct Use of Fracture Toughness for Flaw Evaluations of
Pressure Boundary Materials in Section XI, Division 1, Class 1 Vessels”
Code Action No. 14-1073

Purpose: The purpose of this document is to define a sample problem to test the procedures of
proposed Code Case N-830-1. This problem will be solved in two phases. Phase I
involves developing solutions for allowable toughness (e.g., JIc, KIc) using two
different methods: (A) those proposed by the proposed revision to Code Case N-830,
and (B) those used by existing ASME Code Appendix A rules. Phase II (LATER)
will include developing solutions for allowable flaw size for these two methods.

Inputs: Figure B-1 defines the inputs needed for Phase I of the sample problem.

Requested Outputs:

A. Calculate the variation with temperature of allowable toughness (i.e., of KJc, KIa, JIC, and J0.1)
for the nine (9) cases identified in Figure B-1 (Case ID-s are given in RED TYPE) using the
proposed Code Case methodology (Attachment 1) for Emergency and Faulted operating
conditions.
B. Calculate the variation with temperature of allowable toughness (i.e., of KIc, KIa, and J0.1) for
the nine (9) cases identified in Figure B-1 (Case ID-s are given in RED TYPE) using the
existing Code Appendix A methodology for Emergency and Faulted operating conditions.
Note: The ASME Code does not currently specify a method to estimate J0.1 values
for different steels. In the past, the equations in USNRC Regulatory Guide
1.161 have been used. A Member solving this Sample Problem may elect to
use Regulatory Guide 1.161, or another method. If another method is used it
should either be documented or adequately referenced.
C. Make graphs to compare the results of the (A) and (B) calculations for each of the nine (9)
cases.
D. Fill out the attached spreadsheet with one worksheet, labelled with the Case ID, for each test
case.

A-1
Sample Problem 1 Statement

Figure A-1
Sample Problem Properties

A-2
B
SAMPLE PROBLEM 2 STATEMENT

Sample Problem 2 Statement 5

Proposed Code Case N-830-R1, “Direct Use of Fracture Toughness for Flaw
Evaluations of Pressure Boundary Materials in Section XI, Division 1, Class 1
Vessels”
Code Action No. 14-1073

Purpose: The purpose of this document is to define the Phase II sample problem to test the
procedures described in proposed Code Case N-830-R1. The Phase I sample problem
involved developing solutions for allowable toughness (e.g., JIc, KIc) using two
different methods: (A) those of the proposed revision to Code Case N-830, and (B)
those of Nonmandatory Appendix A of ASME Code Section XI. Phase II includes
developing solutions for allowable flaw size using these three methods (two based on
CC N-830-R1 and one using the procedures of Appendix A). In the interest of
simplicity, a low temperature isothermal pressurization is analyzed. Because at this
time Nonmandatory Appendix A of ASME Code Section XI applies only to
conditions in fracture mode transition (that is, NOT on the upper shelf) the
temperature to be analyzed is arbitrarily set at 50o C.

Inputs: This problem pertains to a typical PWR: a 0.2032 m thick pressure vessel (A508
Grade 2 Class 3) with an inner radius of 2.032 m. A semi-elliptical ID surface flaw
of depth 0.01 m and length 0.0333 m has been found, oriented axially in the vessel.
The vessel operating pressure is 15.3 MPa and the analysis temperature is 50o C. The
vessel is made from an A508 Grade 2 Class 3 forging that has a specified minimum
yield strength of 450 MPa. Assume isothermal conditions at the pressure and
temperature to be used in this analysis. Ignore the plastic zone correction (for
simplicity).

The material property inputs for nine cases are defined in Table B-1.

5
Sections of the problem statement in red are additions that were made to the original problem statement to provide
clarification and more detailed instructions to the participants.

B-1
Sample Problem 2 Statement

Table B-1
Material Properties for use in the Sample Problem 2

Low Transition Medium Transition High Transition


CDF Actual
Temperature Temperature Temperature
%-ile RTNDT-T0
Case RTNDT T0 Case RTNDT T0 Case RTNDT T0
RTNDT-T0 o
C
ID o
C o
C ID C o o
C ID o
C o
C
th
5 10 L-5 -20 -30 M-5 100 90 H-5 180 170
th
50 48 L-50 -20 -68 M-50 100 52 H-50 180 132
95th 92 L-95 -20 -112 M-95 100 8 H-95 180 88

Requested Outputs:
E. Calculations: Calculate the allowable toughness and critical flaw size for each of the nine
sets of material properties using the three different methods described below. For each of the
three methods (a) use the standard ASME Code Appendix A methodology to define the
driving force, and (b) assume that the aspect ratio of the critical flaw is fixed at a/ℓ =0.3. For
Method 1 and 2 use J0.1 for upper shelf calculations. Use the 1% lower bound for Method 1
and the Median J0.1/√10 for Method 2. For critical flaw size evaluations for all three
methods, use 75% as the cut-off for a/t.

The three methods are distinguished by how the allowable toughness is calculated, as
described below:
Method 1: Determine the allowable toughness using the equations in CC-N-830-R1
using 1st percentile toughness values (i.e., for normal operating conditions).
Method 2: Determine the allowable toughness using the mean or median toughness
values from CC-N-830-R1 divided by √10.
Method 3: Determine the allowable toughness using the toughness equations in
Appendix A divided by √10.

Note on Method 3: The ASME Code does not currently specify a method to estimate J0.1 values
for different steels. In the past, most analysts made use of the equations in USNRC Regulatory
Guide 1.161. Problem solvers may elect to use Regulatory Guide 1.161, or another method, or
simply state that the answer is “undefined” according to Appendix A. If another method is used,
it should be referenced.

F. Tabular output format: The output is requested in the tabular format as shown in Table
C-2. For any case for which the input flaw size exceeds the allowable flaw size, please
highlight the allowable flaw size in red.
G. Please provide a step-by-step description of how you determine the critical flaw size
(method of driving force determination, etc.) so that we might better understand any
discrepancies.

B-2
Sample Problem 2 Statement

Table B-2
Table for Presentation of Results of the Phase II Sample Problem
Ratio Ratio
Per CC N-830-R1 Per CC N-830-R1 Per ASME SC-XI
Case # Desired Results Method 1/ Method 1/
Method 1 Method 2 Method 3
Method 3 Method 3
Allowable Toughness (MPa√m)
L5
Allowable Flaw Size (mm)
Allowable Toughness (MPa√m)
L50
Allowable Flaw Size (mm)
Allowable Toughness (MPa√m)
L95
Allowable Flaw Size (mm)
Allowable Toughness (MPa√m)
M5
Allowable Flaw Size (mm)
Allowable Toughness (MPa√m)
M50
Allowable Flaw Size (mm)
Allowable Toughness (MPa√m)
M95
Allowable Flaw Size (mm)
Allowable Toughness (MPa√m)
H5
Allowable Flaw Size (mm)
Allowable Toughness (MPa√m)
H50
Allowable Flaw Size (mm)
Allowable Toughness (MPa√m)
H95
Allowable Flaw Size (mm)

B-3
C
DRAFT CC N-830-1 (VERSION USED FOR SAMPLE
PROBLEM 2)

Case
14-1073 N-830-1
Editor’s Note: This proposal replaces
Rev. 0 N-830 in its entirety
11/8/2016

Case N-830-1

Direct Use of Fracture Toughness for Flaw Evaluations of Pressure


Boundary Materials in Class 1 Ferritic Steel Components

Section XI, Division 1


Inquiry: What alternative fracture toughness models and reference equations may be used for
analytical evaluations performed in accordance with Nonmandatory Appendices A or K in lieu
of the current requirements of these Appendices when determining the values for KIc, KIa, JIc,
J0.1, and J-R?
Reply: It is the opinion of the Committee that the fracture toughness models based on the Master
Curve Method in accordance with ASTM E-1921 may be used in lieu of the current requirements
of Nonmandatory Appendices A or K when determining values for KIc, KIa, JIc, J0.1, and J-R
using the procedures and equations given below.

-1000 Scope
(a) This Case applies to Division 1, Class 1 ferritic steel components subject to the scope of
applicability of the ductile crack extension toughness equations based on product form.
(b) This Case defines the variation of fracture toughness as a function of temperature over the
entire material toughness range of interest to operating Class 1 vessels (lower shelf, transition
region, and upper-shelf).
(c) This Case may be used as an alternative to the determination of the following:
(1) Crack initiation fracture toughness reference curve, KIc, of Nonmandatory Appendix A,
Paragraph A-4200 for pressure retaining materials other than bolting,
(2) Crack arrest fracture toughness reference curve, KIa, of Nonmandatory Appendix A,
Paragraph A-4200 for pressure retaining materials other than bolting,

C-1
Draft CC N-830-1 (Version Used for Sample Problem 2)

(3) J-integral fracture resistance for the material at a ductile flaw extension of 0.1 in. (2.5
mm), J0.1, of Nonmandatory Appendix K Class 1 for pressure retaining materials other
than bolting, and
(4) J-integral fracture resistance curve for the material and its variation with ductile flaw
extension, Δa, J-R, of Nonmandatory Appendix K Class 1 for pressure retaining materials
other than bolting.
(d) When using this Case as part of an analytical evaluation, it is the responsibility of the user to
account for the variability inherent to the fracture toughness properties. For traditional
deterministic analysis, variability in the fracture toughness curve can be accounted for
through the selection of a statistical limit associated with a particular percentile of the
toughness distribution. For a probabilistic assessment, the variability about the mean and
median trends is evaluated through a sampling simulation using a suitable numerical method
such as the Monte Carlo method or other techniques. Guidance on this subject is given in -
3000 and -4000.
(e) This Case provides the fracture toughness information necessary to support both
deterministic and probabilistic assessments and is organized in seven sections, as follows:
-1000 Scope: provides the applicability and organization of the Case
-2000 Reference Toughness Temperature: describes data needed to use this Code
Case.
-3000 Toughness Variability: describes values that can be used along with the
equations of -4000 to account for toughness variability.
-4000 Toughness Curves: provides equations describing the variation with
temperature of, and variability associated with, cleavage and ductile crack
initiation toughness, and of cleavage crack arrest toughness.
-5000 Applicability Limits: provides the limits over which the curves of -4000 can be
applied.
-6000 Unit Conversions: input values and equations within this Code Case are
expressed in System International (SI) units. Conversions to US Customary
(USC) units are provided.
-7000 Nomenclature: symbols and abbreviations used in this Code Case are defined.

-2000 Reference Temperature


The principle parameter for calculating fracture toughness parameters, KJc, KIa, JIc, J0.1, and J-R,
is the reference temperature, T0. The temperature, T0, is based on the Master Toughness Curve
method and is calculated from measured toughness data in accordance with ASTM E 1921, or
from database information on representative RPV steels. The product form of the steel is
accounted for in -4400.

C-2
Draft CC N-830-1 (Version Used for Sample Problem 2)

The value of T0 may be adjusted to account for the effects of measurement or material
uncertainty6. Also, the value of T0 shall be adjusted to account for the effects of irradiation if the
component of interest is exposed to neutron irradiation in excess of 1×1017 n/cm2 (E > 1 MeV) 7.
The equations in -4000 assume that the value of T0 has been adjusted appropriately for these
effects. The cognizant regulatory authority may specify margins or adjustments to T0 intended to
account for these effects.

-3000 Toughness Variability


The equations in -4000 describe both the temperature dependence of and the variability in the
fracture toughness of ferritic steels from the lower shelf through to the upper shelf. Bounding
toughness curves can be generated from the equations in -4000 by using the values of p and Mp
from Table C-1, where p is the probability limit to be used and Mp is the standard normal
distribution with a mean of zero and a standard deviation of unity. Selection of values of p and
Mp appropriate for a particular application is the responsibility of the user. The values of p and Mp
selected should be applied consistently to all toughness curves. When the fracture toughness
equations are used as part of a probabilistic analysis, then the entire range of the distribution (that
is, p ranging from 0 to 1) would be used.
Table C-1
Values of p and Mp Corresponding to Different Bounding Toughness Curves

p Mp

0.005 2.58

0.010 2.33
0.015 2.17

0.020 2.05

0.025 1.96
0.030 1.88

0.035 1.81

0.040 1.75
0.045 1.70

0.050 1.64

0.055 1.60
0.060 1.55

6
Methods to perform adjustments for measurement uncertainty can be found in ASTM Standard Test Method
E1921-14, “Standard Test Method for Determination of Reference Temperature, To, for Ferritic Steels in the
Transition Range.”
7
Methods to perform adjustments for irradiation can be found in ASTM Standard Guide E900-15, “Standard Guide
for Predicting Radiation-Induced Transition Temperature Shift in Reactor Vessel Materials.”

C-3
Draft CC N-830-1 (Version Used for Sample Problem 2)

Table C-1 (continued)


Values of p and Mp Corresponding to Different Bounding Toughness Curves

p Mp

0.065 1.51

0.070 1.48

0.075 1.44
0.080 1.41

0.085 1.37

0.090 1.34
0.095 1.31

0.100 1.28

Note: Values of Mp can be calculated for any value of p using


MS-Excel with the formula Mp = -NORMSINV(p).

-4000 Toughness Curves


This article provides the equations to calculate the toughness parameters, KJc, KIa, JIc, Jx, and J-R.
Regarding J0.1 and J-R provided in this Case, J-R is calculated first based on the equations in
-4400. The J resistance toughness for a fixed amount of ductile crack extension, ∆a, called JX,
where X is a fixed amount of ductile crack extension, is calculated from the J-R curve. As an
example, the value, J0.1 is the case of JX where X = 0.01 in.
Values of K are expressed in MPa√m, values of J are expressed in kJ/m2, and values of
temperature are expressed in °C. J can be converted to K as follows:

𝑱𝑱⋅𝑬𝑬
𝑲𝑲 = � Eq. C-1 (a)
𝟏𝟏−𝝂𝝂𝟐𝟐

where,
{𝟐𝟐𝟐𝟐𝟐𝟐𝟐𝟐𝟐𝟐𝟐𝟐−𝟕𝟕𝟕𝟕.𝟒𝟒𝟒𝟒}
𝑬𝑬 = Eq. C-1 (b)
𝟏𝟏𝟏𝟏𝟏𝟏𝟏𝟏

where E is elastic modulus in units of GPa.

-4100 Cleavage Crack Initiation toughness, KJc


The median cleavage crack initiation toughness curve is defined as follows:

𝑲𝑲𝑱𝑱𝑱𝑱(𝒎𝒎𝒎𝒎𝒎𝒎𝒎𝒎𝒎𝒎𝒎𝒎) = 𝟑𝟑𝟑𝟑 + 𝟕𝟕𝟕𝟕 ⋅ 𝒆𝒆𝒆𝒆𝒆𝒆[𝟎𝟎. 𝟎𝟎𝟎𝟎𝟎𝟎(𝑻𝑻 − 𝑻𝑻𝒐𝒐 )] Eq. C-2 (a)

The cleavage crack initiation toughness curve at percentile p is defined as follows:


𝒑𝒑
𝑲𝑲𝑱𝑱𝑱𝑱 = 𝟐𝟐𝟐𝟐 + (𝑲𝑲𝒐𝒐 − 𝟐𝟐𝟐𝟐){−𝒍𝒍𝒍𝒍(𝟏𝟏 − 𝒑𝒑)}𝟏𝟏/𝟒𝟒 Eq. C-2 (b)

C-4
Draft CC N-830-1 (Version Used for Sample Problem 2)

Eq. C-2(b) can be used to produce both lower and upper bound curves. For example, using a
value of p=0.05 would produce a 5% lower bound curve while using a value of p=0.95 would
produce a 95% upper bound curve.
In eq. C-2(b),
𝑲𝑲𝒐𝒐 = 𝟑𝟑𝟑𝟑 + 𝟕𝟕𝟕𝟕 ⋅ 𝒆𝒆𝒆𝒆𝒆𝒆[𝟎𝟎. 𝟎𝟎𝟎𝟎𝟎𝟎(𝑻𝑻 − 𝑻𝑻𝒐𝒐 )] Eq. C-2 (c)

There is no effect of product form on eq. C-2.


Eq. C-2 applies to a crack front length of 25.4 mm (1 in.) in straight-fronted laboratory test
specimens. While an adjustment to eq. C-2 to account for different crack front lengths in
laboratory test specimens have been developed, currently there is insufficient basis to
recommend a generic equation that applies to the non-straight fronted cracks (e.g., surface
breaking cracks, fully embedded cracks, etc.) of interest in a structural analysis. For this
application, eq. C-2 is recommended unless the user can demonstrate that a crack front length
other than 25.4 mm (1 in.) is appropriate to the structural situation of interest.

-4200 Cleavage Crack Arrest Toughness, KIa


The crack arrest toughness curve at percentile p is defined as follows:
𝒑𝒑
𝑲𝑲𝑰𝑰𝑰𝑰 = 𝑲𝑲𝒎𝒎𝒎𝒎𝒎𝒎𝒎𝒎 �𝟏𝟏 − 𝟎𝟎. 𝟏𝟏𝟏𝟏𝑴𝑴𝒑𝒑 � Eq. C-3 (a)
for lower bound curves 𝑰𝑰𝑰𝑰

(𝟏𝟏−𝒑𝒑)
𝑲𝑲𝑰𝑰𝑰𝑰 = 𝑲𝑲𝒎𝒎𝒎𝒎𝒎𝒎𝒎𝒎 �𝟏𝟏 + 𝟎𝟎. 𝟏𝟏𝟏𝟏𝑴𝑴𝒑𝒑 � Eq. C-3 (b)
for upper bound curves 𝑰𝑰𝑰𝑰

where,

𝑲𝑲𝒎𝒎𝒎𝒎𝒎𝒎𝒎𝒎
𝑰𝑰𝑰𝑰 = 𝟑𝟑𝟑𝟑 + 𝟕𝟕𝟕𝟕 ⋅ 𝒆𝒆𝒆𝒆𝒆𝒆[𝟎𝟎. 𝟎𝟎𝟎𝟎𝟎𝟎(𝑻𝑻 − 𝑻𝑻𝑲𝑲𝑲𝑲𝑲𝑲 )] Eq. C-3 (c)

𝑻𝑻𝑲𝑲𝑲𝑲𝑲𝑲 = 𝑻𝑻𝒐𝒐 + 𝟒𝟒𝟒𝟒. 𝟗𝟗𝟗𝟗𝒆𝒆𝒆𝒆𝒆𝒆[−𝟎𝟎. 𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝑻𝑻𝒐𝒐 ] Eq. C-3 (d)

As an example, a value for p of 0.05 will produce a 5% lower bound using eq. C-3a and a
95% upper bound using eq. C-3b. When using eq. C-3, the value of p shall not exceed 0.5
(0 < p < 0.5). There is no effect of component thickness, crack front length or product form
on eq. C-3.

-4300 Ductile Crack Initiation Toughness, JIc


The ductile crack initiation toughness curve at percentile p or (1-p) can be defined as follows:
𝒑𝒑
𝑱𝑱𝑰𝑰𝑰𝑰 = 𝑱𝑱𝒎𝒎𝒎𝒎𝒎𝒎𝒎𝒎 − 𝝈𝝈𝚫𝚫𝑱𝑱𝑰𝑰𝑰𝑰 𝑴𝑴𝒑𝒑 Eq. C-4 (a)
for lower bound curves 𝑰𝑰𝑰𝑰

(𝟏𝟏−𝒑𝒑)
𝑱𝑱𝑰𝑰𝑰𝑰 = 𝑱𝑱𝒎𝒎𝒎𝒎𝒎𝒎𝒎𝒎 + 𝝈𝝈𝚫𝚫𝑱𝑱𝑰𝑰𝑰𝑰 𝑴𝑴𝒑𝒑 Eq. C-4 (b)
for upper bound curves 𝑰𝑰𝑰𝑰

As an example, a value for p of 0.05 would produce a 5% lower bound using eq. C-4a and
a 95% upper bound using eq. C-4b. When using eq. C-4, the value of p shall not exceed
0.5 (0 < p < 0.5).

C-5
Draft CC N-830-1 (Version Used for Sample Problem 2)

In eqs. C-4a and C-4b,

𝑱𝑱𝒎𝒎𝒎𝒎𝒎𝒎𝒎𝒎
𝑰𝑰𝑰𝑰 = 𝟏𝟏. 𝟕𝟕𝟕𝟕{𝟏𝟏𝟏𝟏𝟏𝟏𝟏𝟏 ∙ 𝐞𝐞𝐞𝐞𝐞𝐞[−𝟎𝟎. 𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎(𝑻𝑻 + 𝟐𝟐𝟐𝟐𝟐𝟐. 𝟏𝟏𝟏𝟏)] − 𝟑𝟑. 𝟑𝟑𝟑𝟑𝟑𝟑} + 𝑱𝑱𝒄𝒄(𝑼𝑼𝑼𝑼) − 𝚫𝚫𝑱𝑱𝑰𝑰𝑰𝑰(𝑼𝑼𝑼𝑼) Eq. C-4 (c)

𝟏𝟏−𝝊𝝊𝟐𝟐
𝑱𝑱𝒄𝒄(𝑼𝑼𝑼𝑼) = {𝟑𝟑𝟑𝟑 + 𝟕𝟕𝟕𝟕 × 𝐞𝐞𝐞𝐞𝐞𝐞[𝟎𝟎. 𝟎𝟎𝟎𝟎𝟎𝟎(𝟒𝟒𝟒𝟒. 𝟖𝟖𝟖𝟖𝟖𝟖 − 𝟎𝟎. 𝟐𝟐𝟐𝟐𝟐𝟐𝟐𝟐𝑻𝑻𝒐𝒐 )]}𝟐𝟐 Eq. C-4 (d)
𝑬𝑬𝑼𝑼𝑼𝑼

𝚫𝚫𝑱𝑱𝑰𝑰𝑰𝑰(𝑼𝑼𝑼𝑼) = 𝟏𝟏. 𝟕𝟕𝟕𝟕{𝟏𝟏𝟏𝟏𝟏𝟏𝟏𝟏 ∙ 𝐞𝐞𝐞𝐞𝐞𝐞[−𝟎𝟎. 𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎(𝑻𝑻𝑼𝑼𝑼𝑼 + 𝟐𝟐𝟐𝟐𝟐𝟐. 𝟏𝟏𝟏𝟏)] − 𝟑𝟑. 𝟑𝟑𝟑𝟑𝟑𝟑} Eq. C-4 (e)

{𝟐𝟐𝟐𝟐𝟐𝟐𝟐𝟐𝟐𝟐𝟐𝟐−𝟕𝟕𝟕𝟕.𝟒𝟒𝑻𝑻𝑼𝑼𝑼𝑼 }
𝑬𝑬𝑼𝑼𝑼𝑼 = Eq. C-4 (f)
𝟏𝟏𝟏𝟏𝟏𝟏𝟏𝟏

𝑻𝑻𝑼𝑼𝑼𝑼 = 𝟒𝟒𝟒𝟒. 𝟖𝟖𝟖𝟖𝟖𝟖 + 𝟎𝟎. 𝟕𝟕𝟕𝟕𝟕𝟕𝟕𝟕𝑻𝑻𝒐𝒐 Eq. C-4 (g)

and where σΔJIc is defined as

𝝈𝝈𝜟𝜟𝜟𝜟𝜟𝜟𝜟𝜟 = 𝑨𝑨 ⋅ 𝒆𝒆𝒆𝒆𝒆𝒆[𝑩𝑩(𝑻𝑻 − 𝟐𝟐𝟐𝟐𝟐𝟐)] Eq. C-4 (h)

𝑨𝑨 = 𝟗𝟗. 𝟎𝟎𝟎𝟎 ⋅ 𝒆𝒆𝒆𝒆𝒆𝒆(𝟏𝟏. 𝟏𝟏𝟏𝟏 ⋅ 𝑷𝑷) Eq. C-4 (i)

𝑩𝑩 = 𝐌𝐌𝐌𝐌𝐍𝐍{𝟎𝟎, (𝟎𝟎. 𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎 ⋅ 𝑷𝑷 − 𝟎𝟎. 𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎)} Eq. C-4 (j)

𝑷𝑷 = 𝐌𝐌𝐌𝐌𝐌𝐌{𝟏𝟏, 𝐌𝐌𝐌𝐌𝐌𝐌[𝟎𝟎, 𝐌𝐌𝐌𝐌𝐌𝐌(𝑷𝑷𝟏𝟏 , 𝑷𝑷𝟐𝟐 )]} Eq. C-4 (k)

𝑱𝑱𝑰𝑰𝑰𝑰(𝟐𝟐𝟐𝟐𝟐𝟐)
𝑷𝑷𝟏𝟏 = − 𝟎𝟎. 𝟒𝟒𝟒𝟒 Eq. C-4 (l)
𝟏𝟏𝟏𝟏𝟏𝟏

𝑱𝑱𝑰𝑰𝑰𝑰(𝟐𝟐𝟐𝟐𝟐𝟐)
𝑷𝑷𝟐𝟐 = + 𝟎𝟎. 𝟓𝟓𝟓𝟓 Eq. C-4 (m)
𝟖𝟖𝟖𝟖𝟖𝟖

The value JIc(288) in eqs. C-4l and C-4m is calculated using eq. C-4c and a value of 288°C for T.
There is no effect of component thickness or crack front length on eq. C-4.

-4400 Ductile Crack Extension Toughness, J-R and JX


The ductile crack extension toughness, which defines the value of J at a specified amount of
ductile crack extension at percentile p or (1-p), is as follows:
𝒑𝒑
𝑱𝑱𝑿𝑿 = 𝒆𝒆𝒆𝒆𝒆𝒆�𝒍𝒍𝒍𝒍[𝑱𝑱𝒎𝒎𝒎𝒎𝒎𝒎𝒎𝒎 ] − 𝑴𝑴𝒑𝒑 × 𝑹𝑹𝑹𝑹𝑹𝑹𝑹𝑹� Eq. C-5 (a)
for lower bound curves 𝑿𝑿

(𝟏𝟏−𝒑𝒑)
= 𝒆𝒆𝒆𝒆𝒆𝒆�𝒍𝒍𝒍𝒍[𝑱𝑱𝒎𝒎𝒎𝒎𝒎𝒎𝒎𝒎 ] + 𝑴𝑴𝒑𝒑 × 𝑹𝑹𝑹𝑹𝑹𝑹𝑹𝑹� Eq. C-5 (b)
for upper bound curves 𝑱𝑱𝑿𝑿 𝑿𝑿

As an example, a value for p of 0.05 would produce a 5% lower bound using eq. C-5a and a 95%
upper bound using eq. C-5b. The value of p used in eqs. C-5a and C-5b shall not exceed 0.5.
In eqs. C-5a and C-5b,
𝑱𝑱𝒎𝒎𝒎𝒎𝒎𝒎𝒎𝒎
𝑿𝑿 = 𝑪𝑪 × 𝚫𝚫𝒂𝒂𝒏𝒏 Eq. C-5 (c)

𝑪𝑪 = 𝟏𝟏. 𝟔𝟔𝟔𝟔 × 𝑱𝑱𝒎𝒎𝒎𝒎𝒎𝒎𝒎𝒎


𝑰𝑰𝑰𝑰 Eq. C-5 (d)

𝒏𝒏 = 𝟎𝟎. 𝟑𝟑𝟑𝟑𝟑𝟑𝟑𝟑 × 𝑪𝑪𝟎𝟎.𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎 Eq. C-5 (e)

𝑿𝑿 = ∆𝒂𝒂 Eq. C-5 (f)

C-6
Draft CC N-830-1 (Version Used for Sample Problem 2)

Here X is used as a subscript for compactness; X signifies a particular amount of ductile crack
extension, or Δa. Values of J at a particular ductile crack extension, for example the value of J
at 2.54 mm (or, equivalently 0.1 in.), can be determined by using these equations. The entire J-R
curve can be produced by solving these equations for a range of ∆a values
The mean toughness, JIcmean is defined in -4300. Values of root mean square deviation (RMSD)
depends on product form as shown in Table C-2. There is no effect of component thickness or
crack front length on eq. C-5.
Table C-2
RMSD values for different product forms.

Product Form RMSD

Forging (A508) 0.106

Plate (A533B) 0.142

RPV Welds 0.134


Linde 80 RPV Welds 0.180

Any other product form* 0.180

* A default of 0.180 is assumed unless the user can justify a lower value.

-5000 Applicability Limits

-5100 Ductile Crack Extension Range


As expressed in eq. C-6a, the value of ductile crack extension used in eq. C-5 must be between
0.5 and 10 mm.
𝟎𝟎. 𝟓𝟓𝐦𝐦𝐦𝐦 ≤ 𝚫𝚫𝒂𝒂 ≤ 𝟏𝟏𝟏𝟏𝐦𝐦𝐦𝐦 Eq. C-6 (a)

-5200 Lower Temperature Limits on KJc and KIa


As expressed in eq. C-6b, eq. C-2 for KJc may be used to temperatures as low as (T0 - 160°C).
𝐓𝐓 ≥ 𝑻𝑻𝟎𝟎 − 𝟏𝟏𝟏𝟏𝟏𝟏 °𝐂𝐂 Eq. C-6 (b)

As expressed in eq. C-6c and eq. C-3 for KIa may be used to temperatures of (TKIa - 100°C).
𝐓𝐓 ≥ 𝑻𝑻𝑲𝑲𝑲𝑲𝑲𝑲 − 𝟏𝟏𝟏𝟏𝟏𝟏 °𝐂𝐂 Eq. C-6 (c)

-5300 Upper Temperature Limits on JIc, J-R, and Jx


As expressed in eq. C-6d, eqs. C-4 and C-5 may be used to temperatures as high as +300°C
𝐓𝐓 < + 𝟑𝟑𝟑𝟑𝟑𝟑 °𝐂𝐂 Eq. C-6 (d)

-5400 Intermediate Temperature Limits


The equations of this Code Case account for the interaction between brittle and ductile fracture
in RPV steels, i.e., when to transition from KJc and KIa to JIc, J-R, and Jx. As such these
equations can be used together, as illustrated in Figure C-1, to develop intermediate limits

C-7
Draft CC N-830-1 (Version Used for Sample Problem 2)

concerning the temperature at which analysis should transition from linear elastic fracture
mechanics (i.e., from using KJc and KIa) to elastic plastic fracture mechanics (i.e., using JIc, Jx,
and J-R).

Figure C-1
Illustration of Intermediate Temperature Limits when 5th Percentile Bounding Curves are
used

-6000 Units Conversions


Table C-3 provides multiplying factors that can be used to convert toughness values from the SI
units to USC units.
Table C-3
Unit Conversion Coefficients

Multiply SI by this
Toughness Curve Symbols SI Units USC Units
factor to determine USC

Cleavage Crack Initiation KJc


MPa√m ksi√in 0.909
Cleavage Crack Arrest KIa
Ductile Crack Initiation JIc
kJ/m2 (in-lbs)/in2 5.713
Ductile Crack Extension Jx & J-R

-7000 Nomenclature
Tables C-4 and C-5 contains a list of symbols and definitions.

C-8
Draft CC N-830-1 (Version Used for Sample Problem 2)

Table C-4
Symbols

Category Symbol Unit Description

KIc MPa√m plane strain fracture initiation toughness

K-equivalent of the value of J measured at cleavage


KJc MPa√m
crack initiation

KIa MPa√m fracture toughness measured at cleavage crack arrest


Different Fracture
ductile crack initiation toughness measured according to
Toughness Metrics JIc kJ/m2
ASTM E1820

ductile crack initiation toughness after “x” mm of ductile


Jx kJ/m2
crack extension

variation of ductile fracture toughness with stable crack


J-R kJ/m2
extension

temperature at which the KJc master curve has a median


T0 °C
value of 100 MPa√m

Index temperature at which the KJa master curve has a median


TKIa °C
Temperatures value of 100 MPa√m

temperature at which the median KJc master curve


TUS °C
crosses the mean JIc upper shelf master curve

Temperature T °C temperature at the crack-tip

p dimensionless percentile for the lower bounding curves


Parameters to
Define Statistical T-statistic multiplier for the lower bounding curves.
Mp dimensionless Standard normal distribution with a mean of zero and
Bounding Curve
standard deviation of 1.0

KJcp MPa√m value of KJc at percentile p


Used in the KIc or KJc
Kmin MPa√m 20 MPa√m
equations
Ko MPa√m value of KJc at the 63.2nd percentile

Used in the KIa


KIap MPa√m value of KIa at percentile p
equation

C-9
Draft CC N-830-1 (Version Used for Sample Problem 2)

Table C-4 (continued)


Symbols

Category Symbol Unit Description

JIcp kJ/m2 value of JIc at percentile p


JXp kJ/m2 value of JX at percentile p

leading coefficient of an exponential fit to the J-R curve.


C kJ/m2
value of C in the equation J = C(Δa)n
slope of an exponential fit to the J-R curve. The value of
n dimensionless
n in the equation J = C(Δa)n

σΔJIc kJ/m2 standard deviation of JIc


JIcmean kJ/m2 mean value of JIc

Used in the J-R and Jc(US) kJ/m2 value of Jc at TUS


JX equations
ΔJIc(US) kJ/m2 value of ΔJIc at TUS
E GPa Young’s modulus

A kJ/m2 fitting parameter in σΔJIc equation

B 1/°C fitting parameter in σΔJIc equation

P dimensionless percentile in σΔJIc equation

P1 dimensionless percentile in σΔJIc equation

P2 dimensionless percentile in σΔJIc equation

ν dimensionless Poisson’s ratio, a value of 0.3 can be used

Table C-5
Definitions

Word Definition

Mean The arithmetic average of a distribution

Median The 50th percentile of a distribution

C-10
The Electric Power Research Institute, Inc. (EPRI, www.epri.com)
conducts research and development relating to the generation, delivery
and use of electricity for the benefit of the public. An independent,
nonprofit organization, EPRI brings together its scientists and engineers
as well as experts from academia and industry to help address
challenges in electricity, including reliability, efficiency, affordability,
health, safety and the environment. EPRI members represent 90% of the
electric utility revenue in the United States with international participation
in 35 countries. EPRI’s principal offices and laboratories are located in
Palo Alto, Calif.; Charlotte, N.C.; Knoxville, Tenn.; and Lenox, Mass.

Together...Shaping the Future of Electricity

Program:
Pressurized Water Reactor Materials Reliability Program (MRP)

© 2017 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.

3002010332

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

You might also like