Download as pdf or txt
Download as pdf or txt
You are on page 1of 411

Wei | Nikbin | McKeighan | Harlow

ASTM INTERNATIONAL
Selected Technical Papers

Fatigue and Fracture


Test Planning, Test

Acquisitions and Analysis


Fatigue and Fracture Test Planning, Test Data
Data Acquisitions
and Analysis

STP 1598
Editors:
Zhigang Wei
Kamran Nikbin
Peter C. McKeighan
Gary D. Harlow

ASTM INTERNATIONAL
Helping our world work better
ASTM International

ISBN: 978-0-8031-7639-3
Stock #: STP1598
www.astm.org
Selected Technical Papers
STP1598

Editors: Zhigang Wei, Kamran Nikbin, Peter C. McKeighan, and D. Gary Harlow

Fatigue and Fracture Test


Planning, Test Data Acquisitions
and Analysis

ASTM STOCK #STP1598


DOI: 10.1520/STP1598-EB

ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959
Printed in the U.S.A.
Library of Congress Cataloging-in-Publication Data
Names: Wei, Zhigang, 1970- editor. | Nikbin, Kamran M., editor. | McKeighan,
P. C. (Peter C.), editor. | Harlow, D. Gary, editor.
Title: Fatigue and fracture test planning, test data acquisitions and
analysis / editors, Zhigang Wei, Kamran Nikbin, Peter C. McKeighan, D.
Gary Harlow.
Description: West Conshohocken, PA : ASTM International, [2017] | Series:
Selected technical papers ; STP1598 | “ASTM Stock #STP1598.” | Includes
bibliographical references.
Identifiers: LCCN 2017000407 | ISBN 9780803176393 (pbk.)
Subjects: LCSH: Materials--Fatigue.
Classification: LCC TA418.38 .F364 2017 | DDC 620.1/1260724--dc23
LC record available at https://lccn.loc.gov/2017000407

ISBN: 978-0-8031-7639-3

Copyright © 2017 ASTM INTERNATIONAL, West Conshohocken, PA. All rights reserved. This material may
not be reproduced or copied, in whole or in part, in any printed, mechanical, electronic, film, or other
distribution and storage media, without the written consent of the publisher.

Photocopy Rights
Authorization to photocopy items for internal, personal, or educational classroom use, or the internal,
personal, or educational classroom use of specific clients, is granted by ASTM International provided that
the appropriate fee is paid to the Copyright Clearance Center, 222 Rosewood Drive, Danvers, MA 01923,
Tel: (978) 646-2600; http://www.copyright.com/

The Society is not responsible, as a body, for the statements and opinions expressed in this publication.
ASTM International does not endorse any products represented in this publication.

Peer Review Policy


Each paper published in this volume was evaluated by two peer reviewers and at least one editor. The
authors addressed all of the reviewers’ comments to the satisfaction of both the technical editor(s) and
the ASTM International Committee on Publications.

The quality of the papers in this publication reflects not only the obvious efforts of the authors and the
technical editor(s), but also the work of the peer reviewers. In keeping with long-standing publication
practices, ASTM International maintains the anonymity of the peer reviewers. The ASTM International
Committee on Publications acknowledges with appreciation their dedication and contribution of time
and effort on behalf of ASTM International.

Citation of Papers
When citing papers from this publication, the appropriate citation includes the paper authors, “paper
title,” STP title, STP number, book editor(s), ASTM International, West Conshohocken, PA, year, page
range, paper doi, listed in the footnote of the paper. A citation is provided on page one of each paper.

Printed in Mayfield, PA
April, 2017
Foreword
THIS COMPILATION OF Selected Technical Papers, STP1598, Fatigue and Frac-
ture Test Planning, Test Data Acquisitions and Analysis, contains peer-reviewed
papers that were presented at a symposium held May 4–5, 2016, in San Antonio,
Texas, USA. The symposium was sponsored by ASTM International Committee
E08 on Fatigue and Fracture and Subcommittee E08.03 on Advanced Apparatus
and Techniques.

Symposium Chairpersons and STP Editors:

Zhigang Wei
Tenneco Inc.
Grass Lake, MI, USA

Kamran Nikbin
Imperial College London
London, UK

Peter C. McKeighan
Symmetry Engineering and Forensic Consulting LLC
Shingle Springs, CA, USA

D. Gary Harlow
Lehigh University
Bethlehem, PA, USA
Contents

Overview vii

Establishing a Multi-Laboratory Test Plan for Environmentally Assisted Fatigue 1


Matthias Bruchhausen, Kevin Mottershead, Caitlin Hurley, Thomas Métais,
Román Cicero, Marc Vankeerberghen, and Jean-Christophe Le Roux

Experimental Study on Surrogate Nuclear Fuel Rods Under Reversed


Cyclic Bending 19
Hong Wang and Jy-An John Wang

Low Cycle Fatigue of Cast Austenitic Steel 37


Xijia Wu, Guangchun Quan, and Clayton Sloss

Fracture Mechanical Testing of In-Service Thermally Aged Cast Stainless Steel 58


Martin Bjurman, Björn Forssgren, and Pål Efsing

Calorimetric Studies and Self-Heating Measurements for a Dual-Phase


Steel Under Ultrasonic Fatigue Loading 81
Noushin Torabian, Véronique Favier, Saeed Ziaei-Rad, Justin Dirrenberger,
Frédéric Adamski, and Nicolas Ranc

Fatigue Studies on Impacted and Unimpacted CFRP Laminates 94


Raghu V. Prakash, Mathew John, Deepika Sudevan, Andrea Gianneo,
and Michele Carboni

Application of Kresidual Measurements to Fracture Toughness Evaluations 119


Gongyao Wang, Kimberly Maciejewski, and Mark James

Sensitivity Study on Parameters that Influence Automated Slope Determination 133


Stephen M. Graham

Contribution to the Evaluation of Stress-Strain and Strain-Life Curves 151


Michael Wächter and Alfons Esderts

v
Methods Development for Nonlinear Analysis of Fatigue Data 186
Bruce A. Young, Richard C. Rice, Steven R. Thompson, and Doug Hall

Data Processing Procedure for Fatigue Life Prediction of


Spot-Welded Joints Using a Structural Stress Method 198
Hong-Tae Kang, Xiao Wu, Abolhassan K. Khosrovaneh, and Zhen Li

More Accurate Elastic Compliance Equation and Its Inverse Solution


for Compact Specimens 212
Xian-Kui Zhu

A Novel Nonlinear Kinematic Hardening Model for Uniaxial/


Multiaxial Ratcheting and Mean Stress Relaxation 227
Hao Wu and Zheng Zhong

Fatigue Damage Indicators Based on Plastic Deformation 246


Grzegorz Socha

A Fatigue Failure Mode Transition Criterion for Sizing Load-Carrying


Fillet-Welded Connections 258
Shizhu Xing and Pingsha Dong

Analysis of Nonproportional Multiaxial Fatigue Test Data of Various


Aluminum Alloys Using a New Damage Parameter 278
Jifa Mei and Pingsha Dong

A Theory for Mathematical Framework and Fatigue Damage Function


for the S-N Plane 299
Hoda Eskandari and Ho Sung Kim

Verification and Validation of Accelerated Testing Methods for


Fatigue Life Assessment 337
Limin Luo, Jason Hamilton, Zhigang Wei, and Robert Rebandt

Load Spectrum Test and Fatigue Failure Study of High-Speed


Train Carbody Anti-Yawing Seat 359
Wenjing Wang, Jinyi Bai, Sichun Li, Hongwei Zhao, and Weiguang Sun

Practical and Technical Challenges of the Exhaust System Fatigue


Life Assessment Process at Elevated Temperature 371
Mark T. Seitz, Jason D. Hamilton, Richard K. Voltenburg,
Limin Luo, Zhigang Wei, and Robert G. Rebandt

vi
Overview
ASTM STP1598 contains a collection of 20 peer-reviewed papers from the Sympo-
sium on Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis
held May 4–5, 2016, in San Antonio, Texas, USA. The symposium was sponsored
by ASTM Committee E08 on Fatigue and Fracture in conjunction with the 2016
May standards development meetings of the Committee. The symposium was
attended by a number of professionals representing several countries, including
the United States, Canada, United Kingdom, Germany, Australia, Netherlands,
Sweden, China, and India. The driving force behind this symposium is the revision
of several relevant ASTM standards, especially ASTM E739-10 (2015), Standard
Practice for Statistical Analysis of Linear or Linearized Stress-Life (S-N) and Strain-
Life (ε-N) Fatigue Data.
Understanding and preventing fatigue failure and fracture of engineering mate-
rials and structures are critical in several industries. Material testing is fundamen-
tal to gaining a better understanding of fatigue and fracture phenomena, as well as
to guide materials selection, product design, and quality control. In fact, engineer-
ing design, development, and validation heavily relies on accurate test data and the
proper interpretation of test data. Although a significant amount of knowledge and
understanding has been gained over the last several decades via material testing,
there still remains a substantial amount of improvement needed due to procedural
deficiencies and limitations. The need for testing improvement is becoming more
critical as materials are increasingly stretched to their limits by extreme conditions
of temperature, stress, corrosive environments, and longer service life cycles. With
new applications, some of the previously tested materials and procedures prove
inadequate and inconsistent, demanding a collective and interdisciplinary effort
to generate reliable and high-quality data. To embrace the new developments, the
following areas of particular interest were selected as the main themes for the sym-
posium: 1) test planning, 2) data acquisition and processing, and 3) data analysis
and interpretation. Although many of the papers and presentations include content
focused on these three topics, the symposium co-chairs were pleased to welcome
other closely related topics and emerging issues in fatigue and fracture.
The major objective of the symposium was to provide a forum for engineers, man-
agers, researchers and scholars worldwide to exchange ideas, share best practices,

vii
discuss challenges, and identify opportunities and directions for future develop-
ments and applications. Specific objectives include:

1) Showcase the most current research and advances in these areas;


2) Promote a systematic, unified materials test plan for improved data acquisition
and analysis; and
3) Collect information and supporting documents for updating existing fatigue,
creep, and fracture test standards and identify the needs for new standards.

Two keynote lectures at the symposium were presented by Krishnaswamy Ravi-


Chandar (The University of Texas at Austin) and Youshi Hong (Institute of Mechanics,
Chinese Academy of Sciences), respectively, at the beginning of each day of the two-day
symposium. A panel discussion on “The Challenges and Opportunities in Fatigue and
Fracture Test Planning, Test Data Acquisitions, and Analysis” was held at the end of the
first day of the symposium. The panel consisted of the following experts in their respec-
tive areas: Michael Shepard (MTS Systems Corporation), Steven Thompson (AFRL/
RXSA), Dan Lingenfelser (HBM nCode Federal LLC), Peter McKeighan (Symmetry
Engineering and Forensic Consulting LLC ), and Charlotte Belsick (Lockheed Martin).
Bruce Young (Battelle) served as a substitute chair in the last day of the symposium.
The papers presented in the symposium were arranged into four sessions:

Session 1: Testing Planning and Performance Characterization


Session 2: Data Acquisition, Quality Assurance, and Analysis
Session 3: Modeling/Simulation, Interpretation, and Correlation
Session 4: Verification, Validation, and Applications

The papers collected in this STP are arranged in the same order. These papers
provide a diverse source of new information regarding test planning, data acquisition
and analysis that can help accelerate the revision of the existing standards and the
development of new standards. These papers also represent a significant contribution
to ASTM E08’s commitment to expanding the knowledge base that supports design
and testing as related to fatigue and fracture.
The symposium co-chairs express our sincere gratitude to ASTM staff for all their
contributions to planning throughout the many months preceding the symposium
and the STP1598 publication. Additionally, Dr. Markus Heinimann (Arconic) and
Charlotte Belsick (Lockheed Martin) are also highly appreciated for their help and
support. Furthermore, this STP would not have been possible without the atten-
tiveness and countless hours volunteered by our peer reviewers to ensure that all of
the manuscripts were suitable for publication. Finally, special thanks are given to
the authors and reviewers of the papers for their outstanding efforts in writing and
reviewing efforts that make the symposium and the STP possible. It is our sincere
hope that these selected technical papers contribute significantly to the further
advancement of the relevant topics.

viii
FATIGUE AND FRACTURE TEST PLANNING, TEST DATA ACQUISITIONS AND ANALYSIS 1

STP 1598, 2017 / available online at www.astm.org / doi: 10.1520/STP159820160047

Matthias Bruchhausen,1 Kevin Mottershead,2


Caitlin Hurley,3 Thomas Métais,4 Román Cicero,5
Marc Vankeerberghen,6 and Jean-Christophe Le Roux7

Establishing a Multi-Laboratory
Test Plan for Environmentally
Assisted Fatigue
Citation
Bruchhausen, M., Mottershead, K., Hurley, C., Métais, T., Cicero, R., Vankeerberghen, M., and Le
Roux, J.-C., “Establishing a Multi-Laboratory Test Plan for Environmentally Assisted Fatigue,”
Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis, ASTM STP1598, Z. Wei,
K. Nikbin, P. C. McKeighan, and D. G. Harlow, Eds., ASTM International, West Conshohocken,
PA, 2017, pp. 1–18, http://dx.doi.org/10.1520/STP1598201600478

ABSTRACT
The European project INCEFA-PLUS will characterize environmental fatigue
in pressurized water reactor (PWR) conditions. The aim is to develop new
guidelines for assessing environmental fatigue damage susceptibility of nuclear
power plant (NPP) components. The consortium consists of 16 public and private
organizations from across Europe. The project is structured in two phases: The
first phase is an extensive fatigue testing program; in the second phase, a
procedure for estimating the environmental fatigue degradation of the materials
will be formulated. During the test phase, a selection of austenitic stainless steels

Manuscript received February 29, 2016; accepted for publication September 15, 2016.
1
European Commission, Joint Research Centre, Westerduinweg 3, 1755 LE Petten, The Netherlands
2
Amec Foster Wheeler, Clean Energy, Europe, Walton House, Birchwood Park, Birchwood, Warrington,
Cheshire WA3 6GA, United Kingdom
3
VTT Technical Research Centre of Finland Ltd., Espoo, 02044 Finland http://orcid.org/
0000-0003-4810-1997
4
EDF-DIPNN SEPTEN, 12-14 Avenue Antoine Dutrièvoz, 69628 Villeurbanne, France
5
Inesco Ingenieros, 39005 Santander, Spain
6
SCK CEN, Nuclear Materials Science Institute, Boeretang 200, B-2400 Mol, Belgium
7
EDF-R & D, Materials and Mechanics of Components Dept., Avenue des Renardières-Ecuelles, 77818 Moret Sur
Loing Cedex, France
8
ASTM Symposium on Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis on May 4–5,
2016 in Grand Hyatt, San Antonio, TX.

Copyright V
C 2017 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.
2 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

used in NPPs will be characterized with regard to fatigue. The test matrix will
focus on the effects of mean strain, strain amplitude, hold time periods, and
surface roughness on fatigue life. Sensitivities to these parameters will be tested
in PWR environments with additional tests in air for reference purposes. The
study of hold time effects will lead to very long testing times, limiting the
total number of tests. It is therefore crucial to establish a test matrix that allows
the study of the principal effects of interest while taking into account the
nuisance effects, such as different specimen geometries, particular material
microstructures, and other laboratory-dependent factors that may not be well-
controlled. Methods and considerations for establishing a single test matrix are
presented in this work.

Keywords
design of experiments, environmentally assisted fatigue, round-robin, nuclear
power plant, pressurized water reactor

Introduction
The rising share of energy from strongly fluctuating renewable sources has forced
today’s nuclear power plants (NPPs) to move from base base-load operation at con-
stant load toward a load-following operating scheme. As a consequence of the oper-
ation at variable outputs, the components in NPPs are exposed to much higher load
variations than originally foreseen and environmentally assisted fatigue risks have
become a major aging mechanism, especially in the context of long-term operation.
Consequently, the nuclear industry and its regulators need to develop new
approaches for assessing fatigue, and especially environmentally assisted fatigue, in
light water reactor (LWR) environments. Currently, the most advanced approach is
that of the U.S. Nuclear Regulatory Commission’s NUREG/CR-6909, Rev. 1 [1] where
the environmental effect on fatigue life Nf is expressed by means of an Fen factor
defined as the ratio of the fatigue life in air at room temperature NfRT
;air to the fatigue
 300
life in LWR environment at operating temperature (300 C) Nf ;LWR , as shown in Eq 1:

NfRT
;air
Fen ¼ : (1)
Nf300
;LWR

According to NUREG/CR-6909, Rev. 1 [1] the Fen factor for austenitic stainless
steel can be expressed as (Eq 2):

Fen ¼  expðT 0 e0 O0 Þ (2)

where T 0 , e0 , and O0 are transformed parameters describing the temperature, strain


rate, and dissolved oxygen content, respectively. Other relevant parameters such as
surface finish are not included in the Fen factor but were included in a subfactor
directly applied to the mean data air curve to develop the fatigue design curve in air.
BRUCHHAUSEN ET AL., DOI 10.1520/STP159820160047 3

Thus, different damaging effects are taken into account by determining factors
for individual effects and multiplying these factors to calculate the cumulative impact
of all effects. There is, however, experimental evidence that this method can lead to
overly penalizing estimates for the fatigue life under plant conditions [2–5].
To address some of these issues, the project, Increasing Safety in NPPs by Cov-
ering Gaps in Environmental Fatigue Assessment (INCEFA-PLUS), has recently
been launched within the European Horizon 2020 framework [6–8]. The project
aims at developing new guidelines for the assessment of environmental fatigue
damage susceptibility of NPP components. It is carried out by a consortium consist-
ing of 16 private and public organizations from across Europe.
The focus is on studying the influence of the mean strain, strain amplitude,
hold time, and surface roughness parameters on fatigue life. To acquire data for
estimating the impact of these parameters and their possible interactions on fatigue
life, a consistent test matrix has been established. The test matrix is based on a
design of experiments (DoE) approach and takes into account a number of nui-
sance factors in addition to the parameters directly targeted within the project.
DoE is a method for establishing optimized test plans that was pioneered in
the first half of the twentieth century [9]. The widespread use of computers has led
to its application in many areas of science and engineering, including materials sci-
ence and fatigue in particular [10–13]. However, examples of using DoE for coor-
dinating round-robin studies including a larger number of laboratories seem to be
less common.
The INCFEA-PLUS project started in July 2015 and is expected to last for five
years. The testing will be carried out in three phases, each of which is scheduled to
last one year. The current work pertains to the establishment of a test matrix for the
first testing phase.

Description of the Methodology


The DoE approach establishes a test plan by systematically varying all relevant
independent variables (“factors”) to study their impact on the dependent variables
(“responses”) [9,14,15]. This is a fundamental difference for another frequently
used approach sometimes referred to as one-factor-at-a-time (OFAT). The OFAT
method implicitly assumes that the impacts of the different factors on the response
are independent from each other. In reality, however, the impact of one factor often
depends on the value of another factor. This dependence of the impact of one factor
on the value of another factor is called “interaction.”
The difference between the OFAT and DoE approaches is shown schemati-
cally in Fig. 1 for an example with two factors, X1 and X2 . In both cases, five
experiments (“runs”) are carried out with both factors in an upper, a lower, and
a central level. In the OFAT approach, the factors are changed separately from
the center point, whereas they are modified simultaneously in the DoE ap-
proach. From the distribution of the test conditions in the x1–x2 plane, it is
4 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 1 Schematic comparison of the (a) OFAT and the (b) DoE approaches.

clear that the case shown in Fig. 1a does not contain any information about the
interactions.
The advantage of the DoE approach is that it allows taking interactions into ac-
count. An example for interaction are the impacts a rough, plant-relevant surface
finish (as compared to a polished laboratory surface) and the PWR environment
(as compared to air) have on the fatigue life of stainless steels. There is experimental
evidence that the combined effect of both, the rough surface and the PWR environ-
ment, is less detrimental on the fatigue life than an estimate based on the individual
impacts of both factors would lead us to expect [4,12].
During a test campaign, a number of runs (tests) i are carried out where the
settings of the factors X1 and X2 are varied between runs (with possibly some
duplications).
When a problem with a single response Y and two factors X1 and X2 is consid-
ered, the corresponding model equation for run i is:

Yi ¼ b0 þ b1 X1i þ b2 X2i þ b12 X1i X2i þ ei : (3)

In this equation, X1i and X2i are the values of the factors for this specific run, b0 ,
b1 , and so on are the unknown model parameters, and ei is the error term. b12 is the
model parameter characterizing the interaction between the factors X1 and X2 .
For more factors and interactions, the model in Eq 3 can be extended:

X
k
Yi ¼ b0 þ bj Xji þ RR bij Xli Xji þ ei : (4)
i<j
j¼1

Eq 4 can be written in matrix form as Eq 5:

Y ¼ Xb þ e: (5)
BRUCHHAUSEN ET AL., DOI 10.1520/STP159820160047 5

For a model with k factors including all main effects and two-factor interactions
and n experimental runs, the model matrix X has the entries (Eq 6):
21 X11 X21  X11 X21  Xk1;1 Xk1 3
61 X12 X22  Xk1;1 Xk2 7
6 .. .. 7
6 7
X ¼ 61 . . 7: (6)
6. 7
4 .. 5
1 X1n X2n  X1n X2n    Xk1;n Xkn

If not all possible interactions or effects are taken into account in the model,
the corresponding terms are omitted in the model matrix and the matrix becomes
smaller.
The unknown model parameters (which are to be determined from the test
data) are the components of the vector b, as shown in Eq 7:
2 3
b0
6 b1 7
6 .. 7
6 7
6 . 7
6 7
6 bk 7
6 7
6 b12 7
6 7
b¼6 .. 7: (7)
6 . 7
6 7
6 b1k 7
6 7
6 b23 7
6 7
6 .. 7
4 . 5
bk1;k

In most cases, interactions of more than two factors are not relevant; therefore,
they are currently not considered for INCEFA-PLUS. Initially, only two levels for
each factor are considered, so no quadratic terms are included in the model.
It is helpful to use the normalized values 61 for the upper and lower levels of
the factors Xi instead of their physical values. This allows direct assessment of the
relative impacts of the different effects by comparing the values of the correspond-
ing model parameters b. Also, the impact of numerical and categorical factors in
the investigated factor range can be compared.
Depending on the aim of a test campaign (e.g., screening, prediction), different
optimization criteria and algorithms can be used for optimizing the test matrix.
Two frequently used optimization approaches are the D-optimal and the I-optimal
design. In D-optimal designs, the variance of the model regression coefficients
(parameters b in Eq 4) is minimized. Mathematically, this corresponds to maximiz-
ing the determinant jX 0 X j where X 0 is the transposed X. D-optimal designs are used
for studies focusing on the identification of those effects that have the most signifi-
cant impact on the response. They are best suited for screening studies. In I-optimal
designs, the smallest average variance of the prediction is sought. These designs are
6 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

TABLE 1 Specifications of the air environment.

Field Name Unit Symbol Comment



Temperature C T 300 C 6 3 C
Pressure bar p Atmospheric

best suited for studies aiming at an accurate prediction of the response Y. Because
the focus of the first phase of the experimental campaign is identifying the most rel-
evant effects, a D-optimal design will be used.

Application to the Project


FACTORS OF INTEREST
In relation to the project, mean strain, strain amplitude, surface roughness, hold
time, and the environment (air or PWR conditions) are the factors under consider-
ation. Initially, two levels will be used for every factor. These are:
• Mean strain:  e ¼ 0 %, 0.5 %.
• Strain amplitude: e0 ¼ 0.3 %, 0.6 %.
• Hold time: th ¼ 0 s, 72 h. (Three hold times will be introduced in each test, reg-
ularly spaced in time with respect to anticipated fatigue life.)
• Surface roughness: Ra ¼ 0.2 lm, 7 lm. (This corresponds approximately to
Rt ¼ 3 lm, 50 lm.)
• Environment: Air, PWR water (the environmental specifications for air and
PWR water are listed in Table 1 and Table 2, respectively).

The other test parameters that are not studied (such as strain rate: 0.01 %/s for
rising strain and 0.1 %/s for decreasing strain) are kept fixed for all tests.
All these factors can be well-controlled, and establishing an optimized test plan
for this case would be straightforward using a commercially available software

TABLE 2 Specifications of the PWR water chemistry.

Field Name Unit Symbol Comment



Temperature C T 300 C 6 3 C
Pressure bar p 150 bar
Lithium content ppm Li 2 ppm 6 0.2 ppm as LiOH
Boron content ppm B 1000 ppm 6 100 ppm as boric acid
Dissolved hydrogen cc(STP)H2/kg H2 25 6 5 cc(STP)H2/kg H2O
pH @ T - pH300 6.95
Anionic contamination <10 ppb
Oxygen <5 ppb
Cationic contamination <100 ppb
TOC <200 ppb
BRUCHHAUSEN ET AL., DOI 10.1520/STP159820160047 7

package. There are, however, additional “nuisance” factors and constraints to be


considered.

NUISANCE FACTORS AND CONSTRAINTS


The number of tests that can be carried out is of course limited. However, thanks to
the relatively large number of partners committed to the project, 74 tests are already
foreseen in the first test phase. For the entire project, approximately 300 tests are
expected to be carried out.
Another aspect is the duration of the individual tests; tests under real plant con-
ditions would simply take too long, especially when hold times are included in the
tests. The test conditions therefore need to be chosen with regard to achieving real-
istic test durations.
These considerations are valid for any fatigue-related test program. For the particu-
lar project discussed in this work, there are additional factors that need to be considered:
• Laboratory: The tests will be distributed over ten different project partners. It
is expected that “laboratory” will not be a determining factor in the tests; this
however has to be verified, so an additional categorical factor “laboratory” is
introduced. In cases where the same organization carries out tests in air and in
PWR environment, two completely different test installations will be used.
The factor “laboratory” refers to the actual test rig in which the tests are car-
ried out—not to the organization. For the purpose of this study, these test rigs
are considered as different laboratories. Finally, seven laboratories will be test-
ing in air and nine will be testing in a PWR environment.
• Specimen Type: For the tests in PWR conditions, two types of specimens will be
used. Although most of the tests in PWR conditions will be carried out using
standard fatigue specimens in an autoclave, some laboratories will use hollow
specimens with the simulated PWR water flowing inside the specimen. This dif-
ference of specimen geometry has consequences for the stress state. While the
full specimens have a purely axial stress in the gage length, the hollow specimens
are exposed to an additional circumferential stress introduced by the pressurized
water. A previous study on thermomechanical fatigue has found differences be-
tween full and hollow specimens, but these were smaller than the differences
among laboratories [16]. Because these tests were performed without internal
pressurization of the hollow specimen, the impact of the specimen type (“full” or
“hollow”) still needs to be investigated for PWR testing. A recent study showed
systematic differences between hollow and smooth specimens [17].
• Material: Although most of the tests will be carried out on a common heat of 304L
stainless steel, some partners will also bring data from national programs into the
project. These tests will be performed on materials of interest for the respective na-
tional programs. In order to align the testing on these different materials as far as
possible with the project’s test plan, for the common material, suggestions for test
conditions for those materials will be included in the common test program. This
requires adding “material” as an additional categorical variable.

“Outlier specimens” will be microstructurally analyzed to establish whether


there are any flaws that might explain their premature failure. Such microstructural
8 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

anomalies can be considered as an additional nuisance factor. However, the micro-


structural analysis can only be performed post-test and is therefore not incorporat-
ed in the planning of the test campaign.
The constraints that need to be incorporated in the test program are summa-
rized in Table 3. These constraints lead to a large number of factor settings being
predefined in the test matrix that is reproduced in Table A1 in the appendix. The
entries in the blue fields of the table are fixed by the five constraints in Table 3 and
cannot be modified for optimizing the test plan.

IMPLEMENTATION OF CONSTRAINTS IN THE TEST MATRIX


The commercial software package JMP 12 [18] has been used for determining an
optimized test program. As mentioned in the earlier section, “Description of the
Methodology,” establishing the D-optimal design means finding the settings of
the test parameters in each run leading to the maximum of the determinant
jX 0 X j. The program uses the coordinate exchange algorithm [19] for that pur-
pose. Initially, a random value (61) is attributed to each factor in every run.
Then the program replaces the value for each factor successively by the opposite
value (e.g., 1 is replaced by þ 1). The effect of the exchange of the factor setting
on the determinant jX 0 X j is calculated, and the value resulting in the larger
determinant is kept. The algorithm cycles through all factors until no further
modification leading to a larger determinant can be found. At that point, a local
maximum has been found.
In order to find a global maximum, the next step is to start the same process
with a new random seed of þ1 and 1 for all entries in the matrix and to restart
the same procedure with different initial values. The best of all locally optimized
designs is considered to be the globally optimized design. Typically, around 1,000
random seeds are used.
The program allows enforcing combinations of factor values by means of a
covariates table. This table has a row for every individual test run in the test cam-
paign. Each row has several fields that define the laboratory, the specimen type, the
environment, and the material for each test. Therefore, these test parameters are
fixed in advance and cannot be changed anymore during the optimization of the
test matrix. This allows implementing Constraints 1–4 listed in Table 3.

TABLE 3 Constraints predefining some factor settings in the test matrix.

1 Each laboratory carries out a predefined number of tests. The number of tests varies among
laboratories.
2 Each laboratory will test either in air or in PWR conditions, not in both.
3 Each laboratory will use only a single type of specimen: either full or hollow.
4 Some laboratories will carry out (a part of) their test program on their own material.
5 Although full specimens can have either surface finish (smooth or rough), hollow specimens
can only be manufactured with a smooth surface finish.
BRUCHHAUSEN ET AL., DOI 10.1520/STP159820160047 9

The program also allows explicit interdicting of certain combinations of factor


values. By disallowing the combination of “hollow specimens” with a “rough surface
finish,” Constraint 5 can be implemented.
As can be seen from Eqs 3 through 6, the test matrix X and therefore the deter-
minant jX 0 X j do not depend only on the number of factors and runs in the test
program but also on the effects included in the model. Before the D-optimal test
matrix can be determined, the terms to be included in the model have to be
selected.

EFFECTS TO BE CONSIDERED
Only main effects and second-order interactions will be taken into account during
the first phase of the project. The number of parameters NP in a model with k fac-
tors at two levels including all main effects, all second order interactions, and the
intercept (b0 in Eq 3) is:

k  ðk  1Þ
NP ¼ k þ 1 þ : (8)
2

It is clear from Eq 8 that the number of parameters NP increases quadratically


with k, which can lead to very complex models.
In the present case, however, not all of the possible effects are considered equal-
ly important by the project partners. Some of the possible interactions will be dis-
carded, which reduces the complexity of the model.
If an interaction between the factor “laboratory” and one of the other factors
was active, that would mean that, for example, the effect of “hold time” or “strain
amplitude” would vary among laboratories. This is considered unlikely, and interac-
tions between the factor “laboratory” and any of the other factors will be discarded.
The main effect “laboratory” will be maintained throughout. This reduces the num-
ber of model parameters by seven (the factor “laboratory” has seven possible
interactions).
Most of the tests will be carried out on the same heat of 304L. The specimens
from this material are all being manufactured in the same workshop to minimize
variations during the production process itself. Also, some of the tests in the frame-
work of national programs will be performed on 304L (albeit on a different heat).
For details, refer to the test matrix in the table in the appendix. Due to the limited
number of tests being performed on “national program” materials, the “material”
factor interactions cannot be studied and have therefore been discarded from the
test matrix. In contrast, the main effect “material” is maintained. This further
reduces the number of model parameters by six (the factor “material” has six
remaining potential interactions since the factor “laboratory” was already removed
in the previous step).
Some of the remaining interactions cannot be studied because the factors in-
volved in these interactions cannot be varied. For example, the “hollow specimens”
10 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

will all have a smooth surface finish; therefore, the data collected during the project
will not allow studying the interaction between the type of specimen and the surface
finish. These interactions are marked “assessment not possible” in Table 4. Because
two out of the four interactions in that category were already discarded during one
of the previous steps, the total number of model parameters is reduced by two,
which leads to 22 remaining parameters.
Some of the remaining main effects and interactions cannot be separated be-
cause they are changed simultaneously between runs. These problems are due to
the constraints that the factors “laboratory,” “specimen type,” and “environment”
are predefined for every test and cannot be varied. The factors “laboratory” and to a
lesser degree of certainty “specimen type” are not expected to have a major impact
on the fatigue life. To confirm (or invalidate) these expectations, the impact of these
factors and the corresponding interactions (labeled with “screening” in Table 4) on
the test outcome will nevertheless be checked during the evaluation. To that end, a
screening evaluation will be carried out first.
For this screening evaluation, the runs will be divided into three groups: the
first group contains all tests in air, the second group contains all tests in PWR water
on full specimens, and the third group contains the tests in PWR water on hollow
specimens. In each of these groups, the effect of the factor “laboratory” can be in-
vestigated. If the corresponding model parameters b are small, the laboratories have
no major effect and can be removed from the analysis. The same procedure will
then be applied to all the tests in PWR water (Groups 2 and 3) where the impact of
the “specimen type” will be assessed. If the factor “specimen type” does not have a
major impact either, it can be removed from the analysis, and the complete data set
can be analyzed, taking into account only the effects in the green fields in Table 4.
This is also the model on which the optimization of the test matrix is based.

Analysis of the Test Matrix


The full test matrix is reproduced in the table in the appendix. In this section, some
of its characteristics will be discussed.
A useful property for the design is to allow the model effects to be determined
independently from one another. The degree to which a given design allows this
can be analyzed using the covariance among the different effects. The covariance of
two random variables x and y is defined in Eq 9:

covðX; Y Þ ¼ E½ðX  Eð X ÞÞðY  EðY ÞÞ (9)

where Eð X Þ refers to the expected value (i.e., mean) of X.


The correlation of two variables is the covariance normalized by the product of
the standard deviations (Eq 10):

covðX; Y Þ
corrðX; Y Þ ¼ : (10)
rX rY
BRUCHHAUSEN ET AL., DOI 10.1520/STP159820160047 11

TABLE 4 List of all possible parameters (intercept, main effects, and second-order interactions).
The effects in the gray cells (“assessment not possible,” “discarded”) will not be assessed;
those in the orange cells (“screening”) will be analyzed in a first screening analysis (see
text for details). The remaining effects (green) are targeted by the project.

Parameter Comment

Intercept (b0 in Eq 3)
Hold time
Mean strain
Strain amplitude
Surface roughness
Environment (air/water)
Specimen Type screening evaluation
Lab screening evaluation
Material
Hold time * Mean strain
Hold time * Strain amplitude
Hold time * Surface roughness
Hold time * Environment
Hold time * Specimen type screening evaluation
Hold time * Lab discarded
Mean strain * Strain amplitude
Mean strain * Surface roughness
Mean strain * Environment
Mean strain * Specimen type screening evaluation
Mean strain * Lab discarded
Strain amplitude * Surface roughness
Strain amplitude * Environment
Strain amplitude * Specimen type screening evaluation
Strain amplitude * Lab discarded
Surface roughness * Environment
Surface roughness * Specimen type assessment not possible
Surface roughness * Lab discarded
Environment * Specimen type assessment not possible
Environment * Lab assessment not possible
Specimen type * Lab assessment not possible
Material * Hold time discarded
Material * Mean strain discarded
Material * Strain amplitude discarded
Material * Surface roughness discarded
Material * Environment discarded
Material * Specimen type discarded
Material * Lab discarded
12 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

The correlation is a measure of the linear dependence between two varia-


bles. If the correlation is 0, the variables are independent. For the test matrix
that means that if the correlation of two effects is 0, their respective impacts on
the test outcome (i.e., the fatigue life) can be determined independently from
each other.
The color map on correlations in Fig. 2 shows the correlations among all the
effects included in the model.
The factor “material” is different from the other factors in that it is a categori-
cal (i.e., non-numerical) variable with four levels. To be able to carry out calcula-
tions, this type of variable is transferred into a set of four variables (e.g., m1, m2,
m3, and m4). Each of these can take the two values 0 and 1. For example, m1 ¼ 1
means “Material 1” one was selected for that test, whereas m1 ¼ 0 means “Material 1”
was not selected. Because exactly one material is used in every test, the factors
m1 to m4 are coupled. This coupling can be expressed as Eq 11:

FIG. 2 Color map of correlations for the final design.


BRUCHHAUSEN ET AL., DOI 10.1520/STP159820160047 13

X
4
mi ¼ 1: (11)
j¼1

It follows that there are only three independent variables mi. Consequently,
only three are listed in Fig. 2.
The red diagonal in the plot (corr ¼ 1) indicates the trivial fact that each effect
is correlated with itself. The important point to notice is that most of the other cor-
relations are close to 0. The current design will allow separating the physical effects
based on the data and estimating the targeted effects independently from one
another.
The standard deviation for the different model parameters depends on the
characteristics of the measurements (hardware, procedures) as well as on the test
matrix and can therefore not be determined before the actual tests have been car-
ried out. However, different test matrices can be compared with regard to their
contributions to the overall standard deviation. One of the measures that can be
used for that purpose is the “fractional increase of confidence interval length” (FI)
[20]. The FI measures by which fraction a design increases the confidence interval
of the parameter estimates compared to an ideal, orthogonal design in which the

TABLE 5 Comparison of the “fractional increase of confidence interval length” for the INCEFA-PLUS
and the design with reduced constraints.

INCEFA-PLUS Design with Ratio (INCEFA-PLUS/


Term Design Reduced Constraints Red. Constraints)

Intercept 1.8009 1.6318 1.1036


Environment 0.0491 0.0267 1.8390
Material 1 2.5930 2.4443 1.0608
Material 2 1.1504 1.0173 1.1308
Material 3 1.2587 1.2239 1.0284
Mean strain 0.0267 0.0256 1.0430
Strain amplitude 0.0154 0.0163 0.9448
Hold time 0.0157 0.0153 1.0261
Surface roughness 0.0157 0.0156 1.0064
Environment*mean strain 0.0267 0.0149 1.7919
Environment*strain amplitude 0.0148 0.0163 0.9080
Environment*hold time 0.0164 0.0270 0.6074
Environment*surface roughness 0.0157 0.0156 1.0064
Mean strain*strain amplitude 0.0151 0.0156 0.9679
Mean strain*hold time 0.0161 0.0162 0.9938
Mean strain*surface roughness 0.0153 0.0163 0.9387
Strain amplitude*hold time 0.0161 0.0265 0.6075
Strain amplitude*surface roughness 0.0161 0.0150 1.0733
Hold time*surface roughness 0.0151 0.0156 0.9679
14 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

correlations between any two different effects are 0. The FI should be as close as
possible to zero. The FI for the ith parameter and n experimental runs is given by
Eq 12 [20]:
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
FI ¼ nðX 0 X Þ1 ii  1: (12)

To assess the impact of the constraints (Table 3) on the performance of


the design, a second (hypothetical) design is calculated. In this hypothetical
design, the number of tests per laboratory per material is maintained (Con-
straints 1 and 3), but all other constraints have been lifted. Although this less-
constrained design would allow including more effects in the model, the model
itself has not been modified for the comparison exercise. The resulting “design
with reduced constraints” is compared to the INCEFA-PLUS design in terms of
the FI.
The results are listed in Table 5. The performance for most parameters is very
similar for both designs. The largest differences show for the effects “Environment”
and “Environment*mean strain.” The worst performance of both designs occurs for
the material effects. This is related to the strong imbalance between the number of
tests for the different materials with very limited number of tests on the materials
from national programs. This is also reflected in the higher correlation of the mate-
rial effects visible in the correlation map (Fig. 2).
Hence, the biggest disadvantage the constraints have on the INCEFA-PLUS
project is not a loss of certainty of the parameter estimates but the fact that some of
the main effects and interactions cannot be estimated simultaneously.

Conclusions
The DoE methodology is used to establish a common test matrix for a European
project on environmentally assisted fatigue. A number of constraints has to be
taken into account in the formulation of the matrix. These constraints prevent a
simultaneous evaluation of all effects targeted in the study and the nuisance
effects. A stepwise evaluation is proposed to eliminate nuisance effects from the
evaluation and to allow the remaining effects to be addressed with a single
model.
The resulting design is almost orthogonal. The constraints do not lead to a
strong increase of the parameter estimates compared to a less constricted, hypothet-
ical design. The main consequence of the constraints is that not all relevant effects
can be analyzed simultaneously.

ACKNOWLEDGMENTS
This project has received funding from the Euratom Research and Training Pro-
gramme 2014–2018 under Grant Agreement No. 662320.
BRUCHHAUSEN ET AL., DOI 10.1520/STP159820160047 15

Appendix

TABLE A1 Final test matrix.

Specimen Mean Strain Hold Surface


Lab Env. Type Material Strain Amplitude Time Roughness

A air full 304L (common) 1 1 1 1


A air full 304L (common) 1 1 1 1
A air full 304L (common) 1 1 1 1
A air full 304L (common) 1 1 1 1
A air full 304L (national) 1 1 1 1
A air full 304L (national) 1 1 1 1
B air full 304L (common) 1 1 1 1
B air full 304L (common) 1 1 1 1
B air full 304L (common) 1 1 1 1
C air full 304L (common) 1 1 1 1
C air full 304L (common) 1 1 1 1
C air full 304L (common) 1 1 1 1
C air full 304L (common) 1 1 1 1
C air full 304L (common) 1 1 1 1
D air full 304L (common) 1 1 1 1
D air full 304L (common) 1 1 1 1
D air full 304L (common) 1 1 1 1
E air full 304L (common) 1 1 1 1
E air full 304L (common) 1 1 1 1
E air full 304L (common) 1 1 1 1
E air full 304L (common) 1 1 1 1
E air full 304L (common) 1 1 1 1
E air full 304L (common) 1 1 1 1
E air full 304L (common) 1 1 1 1
F air full 304L (common) 1 1 1 1
F air full 304L (common) 1 1 1 1
F air full 304L (common) 1 1 1 1
G air full 304L (common) 1 1 1 1
G air full 304L (common) 1 1 1 1
G air full 304L (common) 1 1 1 1
G air full 304L (common) 1 1 1 1
G air full 304L (common) 1 1 1 1
G air full 304L (common) 1 1 1 1
G air full 304L (common) 1 1 1 1
H PWR full 304L (common) 1 1 1 1
H PWR full 304L (common) 1 1 1 1
H PWR full 304L (national) 1 1 1 1
16 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

TABLE A1 (Continued)

Specimen Mean Strain Hold Surface


Lab Env. Type Material Strain Amplitude Time Roughness

H PWR full 304L (national) 1 1 1 1


I PWR hollow 304L (common) 1 1 1 1
I PWR hollow 304L (national) 1 1 1 1
J PWR full SS1 (national) 1 1 1 1
K PWR full 304L (common) 1 1 1 1
K PWR full 304L (common) 1 1 1 1
K PWR full 304L (common) 1 1 1 1
K PWR full 304L (common) 1 1 1 1
K PWR full 304L (common) 1 1 1 1
K PWR full 304L (common) 1 1 1 1
K PWR full 304L (common) 1 1 1 1
L PWR full 304L (common) 1 1 1 1
L PWR full 304L (common) 1 1 1 1
L PWR full 304L (common) 1 1 1 1
L PWR full 304L (common) 1 1 1 1
L PWR full 304L (common) 1 1 1 1
L PWR full 304L (common) 1 1 1 1
L PWR full 304L (common) 1 1 1 1
M PWR hollow 304L (common) 1 1 1 1
M PWR hollow 304L (common) 1 1 1 1
M PWR hollow 304L (common) 1 1 1 1
N PWR full 304L (common) 1 1 1 1
N PWR full SS2 (national) 1 1 1 1
N PWR full SS2 (national) 1 1 1 1
N PWR full SS2 (national) 1 1 1 1
N PWR full SS2 (national) 1 1 1 1
N PWR full SS2 (national) 1 1 1 1
O PWR hollow 304L (common) 1 1 1 1
O PWR hollow 304L (common) 1 1 1 1
O PWR hollow 304L (common) 1 1 1 1
O PWR hollow 304L (common) 1 1 1 1
O PWR hollow 304L (common) 1 1 1 1
O PWR hollow 304L (common) 1 1 1 1
O PWR hollow 304L (common) 1 1 1 1
P PWR full 304L (common) 1 1 1 1
P PWR full 304L (common) 1 1 1 1
P PWR full 304L (common) 1 1 1 1

Note: The settings in blue fields are predefined from the five constraints in Table 3 and cannot be
modified for optimizing the design. The Materials SS1 and SS2 refer to two stainless steels from na-
tional programs. The values 6 1 for the numerical factors refer to the lower and higher level of the
respective factors.
BRUCHHAUSEN ET AL., DOI 10.1520/STP159820160047 17

References

[1] Chopra, O. and Stevens, G. L., “Effect of LWR Coolant Environments on the Fatigue Life
of Reactor Materials,” NUREG/CR-6909, Revision 1, U.S. Nuclear Regulatory Commis-
sion, Washington, DC, 2014.

[2] Tice, D. R., Green, D., and Toft, A., “Environmentally Assisted Fatigue Gap Analysis and
Roadmap for Future Research—Gap Analysis Report,” Technical Report 1023012, Electric
Power Research Institute, Palo Alto, CA, 2011.

[3] Tice, D. R., Green, D., and Toft, A., “Environmentally Assisted Fatigue Gap Analysis and
Roadmap for Future Research—Roadmap,” Technical Report 1026724, Electric Power
Research Institute, Palo Alto, CA, 2012.

[4] Poulain, T., Mendez, J., Hena, G., and De Baglion, L., “Influence of Surface Finish in
Fatigue Design of Nuclear Power Plant Components,” Procedia Eng., Vol. 66, 2013,
pp. 233–239.

[5] Métais, T., Karabaki, E., De Baglion, L., Solin, J., Le Roux, J.-C., Reese, S., and Courtin, S.,
“European Contributions to Environmental Fatigue Issues Experimental Research in
France, Germany, and Finland,” presented at the ASME 2014 Pressure Vessels and Piping
Conference, Anaheim, CA, July 20–24, 2014, http://dx.doi.org/10.1115/PVP2014-28207

[6] Community Research and Development Information Service (CORDIS), “Increasing


Safety in NPPs by Covering Gaps in Environmental Fatigue Assessment,” http://cordis.
europa.eu/project/rcn/197289_en.html (accessed January 20, 2016).

[7] INCEFA-PLUS, “Increasing Safety in NPPs by Covering Gaps in Environmental Fatigue


Assessment,” http://incefaplus.unican.es (accessed January 20, 2016).

[8] Mottershead, K., Bruchhausen, M., Métais, T., Cicero, S., Tice, D., and Norman, P.,
“INCEFA-PLUS (Increasing Safety in Nuclear Power Plants by Covering Gaps in Envi-
ronmental Fatigue Assessment),” presented at the ASME 2016 Pressure Vessels and
Piping Conference, Vancouver, BC, Canada, July 17–21, 2016, http://dx.doi.org/10.1115/
PVP2016-63149

[9] Montgomery, D. C., Design and Analysis of Experiments, 8th ed., Wiley, Hoboken, NJ,
2013.

[10] Andersson, C. and Liu, J., “Effect of Corrosion on the Low Cycle Fatigue Behavior of
Sn-4.0Ag-0.5Cu Lead-Free Solder Joints,” Int. J. Fatigue, Vol. 30, No. 5, 2008,
pp. 917–930.

[11] Glane, S. and Dilger, K., “Consideration of Manufacturing Effects on Fatigue Design for
Welded Chassis Components,” Weld. World, Vol. 60, No. 1, 2016, pp. 71–81.

[12] Le Duff, J. A., Lefrançois, A., Vernot, J. P., and Bossu, D., “Effect of Loading Signal Shape and
of Surface Finish on the Low Cycle Fatigue Behavior of 304L Stainless Steel in PWR Envi-
ronment,” proceedings of the ASME 2010 Pressure Vessels and Piping Division/K-PVP
Conference, Bellevue, WA, July 18–22, 2010, http://dx.doi.org/10.1115/PVP2010-26027

[13] Brandl, E., Heckenberger, U., Holzinger, V., and Buchbinder, D., “Additive Manufactured
AlSi10Mg Samples Using Selective Laser Melting (SLM): Microstructure, High Cycle
Fatigue, and Fracture Behavior,” Mater. Design, Vol. 34, 2012, pp. 159–169.
18 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

[14] Goos, P. and Jones, B., Optimal Design of Experiments: A Case Study Approach, Wiley,
Hoboken, NJ, 2012.

[15] NIST/SEMATECH e-Handbook of Statistical Methods, National Institute of Standards


and Technology, Washington, DC, 2012, http://www.itl.nist.gov/div898/handbook/pri/
section1/pri11.htm (accessed November 16, 2015).

[16] Loveday, M. S., Bicego, V., Hähner, P., Klingelhöffer, H., Kühn, H.-J., and Roebuck, B.,
“Analysis of a European TMF Inter-Comparison Exercise,” Int. J. Fatigue, Vol. 30, No. 2,
2008, pp. 382–390.

[17] Twite, M., Platts, N., Mclennan, A., Meldrum, J., and McMinn, A., “Variations in Measured
Fatigue Life in LWR Coolant Environments Due to Different Small Specimen Geo-
metries,” presented at the ASME 2016 Pressure Vessels and Piping Conference, Vancou-
ver, BC, Canada, July 17–21, 2016, http://dx.doi.org/10.1115/PVP2016-63584

[18] SAS, “JMP Statistical Discovery Software,” SAS Institute, Inc., Buckinghamshire, UK,
http://www.jmp.com/en_nl/home.html (accessed February 16, 2016).

[19] Meyer, R. K. and Nachtsheim, C. J., “The Coordinate-Exchange Algorithm for Construct-
ing Exact Optimal Experimental Designs,” Technometrics, Vol. 37, No. 1, 1995, pp. 60–69.
R
[20] SAS Institute, Inc., JMPV 10 Design of Experiments Guide, SAS Institute, Inc., Cary, NC,
2012.
FATIGUE AND FRACTURE TEST PLANNING, TEST DATA ACQUISITIONS AND ANALYSIS 19

STP 1598, 2017 / available online at www.astm.org / doi: 10.1520/STP159820160051

Hong Wang1 and Jy-An John Wang2

Experimental Study on Surrogate


Nuclear Fuel Rods Under
3
Reversed Cyclic Bending
Citation
Wang, H. and Wang, J.-A. J., “Experimental Study on Surrogate Nuclear Fuel Rods Under
Reversed Cyclic Bending,” Fatigue and Fracture Test Planning, Test Data Acquisitions and
Analysis, ASTM STP1598, Z. Wei, K. Nikbin, P. C. McKeighan, and D. G. Harlow, Eds., ASTM
International, West Conshohocken, PA, 2017, pp. 19–36, http://dx.doi.org/10.1520/STP1598201600514

ABSTRACT
The mechanical behavior of spent nuclear fuel (SNF) rods under reversed cyclic
bending or bending fatigue must be understood to evaluate their vibration
integrity in a transportation environment. This is especially important for high-
burnup fuels (>45 GWd/MTU), which have the potential for increased structural
damage. It has been demonstrated that the bending fatigue of SNF rods can be
effectively studied using surrogate rods. In this investigation, surrogate rods
made of stainless steel 304 cladding and aluminum oxide pellets were tested
under load or moment control at a variety of amplitude levels at 5 Hz using the
Cyclic Integrated Reversible-Bending Fatigue Tester developed at Oak Ridge
National Laboratory. The behavior of the rods was further characterized using
flexural rigidity and hysteresis data, and fractography was performed on the

Manuscript received February 29, 2016; accepted for publication October 5, 2016.
1
Oak Ridge National Laboratory, Materials Science and Technology Division, PO Box 2008, MS-6069, Oak Ridge,
TN, 37831 http://orcid.org/0000-0002-0173-0545
2
Oak Ridge National Laboratory, Materials Science and Technology Division, P.O. Box 2008, MS-6069, Oak Ridge,
TN, 37831 http://orcid.org/0000-0003-2402-3832
3
This manuscript has been authored by UT-Battelle, LLC under Contract No. DE-AC05-00OR22725 with the
U.S. Department of Energy. The Department of Energy will provide public access to these results of federally
sponsored research in accordance with the DOE Public Access Plan (http://energy.gov/downloads/doe-
public-access-plan).
4
ASTM Symposium on Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis on May 4–5,
2016 in Grand Hyatt, San Antonio, TX.

Copyright V
C 2017 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.
20 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

failed rods. The proposed surrogate rods captured many of the characteristics of
deformation and failure mode observed in SNF, including the linear-to-nonlinear
deformation transition and large residual curvature in static tests, pellet-pellet
interface and pellet cladding mechanical interaction, failure mechanisms, and large
variations in the initial structural condition. Rod degradation was measured and
characterized by measuring the flexural rigidity; the degradation of the rigidity
depended on both the moment amplitude applied and the initial structural
condition of the rods. It was also shown that a cracking initiation site can be
located on the internal surface or the external surface of cladding. Finally, fatigue
damage to the bending rods can be described in terms of flexural rigidity, and the
fatigue life of rods can be predicted once damage model parameters are properly
evaluated. The developed experimental approach, test protocol, and analysis
method can be used to study the vibration integrity of SNF rods in the future.

Keywords
spent nuclear fuels, vibration integrity, surrogate rods, reversed bending, cyclic
fatigue, cladding failure

Introduction
Spent fuel rod vibration during normal transport is important because of the
sustained loads imposed by transport, and it is particularly significant for high-
burnup fuel (>45 GWd/MTU). As the burnup of a fuel increases, a number of
changes occur that may affect the performance of the fuel and cladding during
storage and transportation. These changes include increased thickness of the
cladding corrosion layer, increased cladding hydrogen content, increased fission
gas release, and formation of high-burnup structures at the surfaces of the fuel
pellets. With respect to cladding hydrogen content, the solubility of hydrogen in
zirconium increases at higher temperatures. When the concentration of hydro-
gen exceeds the solubility limit, zirconium hydrides form. Depending on size,
distribution, and orientation, these hydrides can embrittle the cladding and re-
duce ductility. Furthermore, the presence of hydrides can facilitate cracking if
the hydrides are aligned radially, perpendicular to the tensile stress field. Even if
a through-wall crack is not formed, the extent of cracking needs to be evaluated
to determine if the cladding will fail as a result of stresses caused by normal han-
dling or transportation [1].
Testing a spent nuclear fuel (SNF) rod is not trivial. First, SNF rods are highly
radioactive. SNF testing must be conducted in a hot-cell environment and the rods
can be accessed only by manipulators. Thus, the test setup, specimen loading, and
test operations must be as simple as possible. Second, a fuel rod has a composite
structure that originally consists of fuel pellets and cladding, which has been modi-
fied significantly by the high-burnup process. Various failure modes could be trig-
gered during transportation, including fracture and splitting. Bending-induced
WANG AND WANG, DOI 10.1520/STP159820160051 21

failure can be captured only by an effective testing approach. Finally, rod vibration
during transportation is not well-characterized.
An innovative hot-cell testing system, the Cyclic Integrated Reversible-
Bending Fatigue Tester (CIRFT), was developed recently by Oak Ridge National
Laboratory (ORNL) [2–6]. Although the use of the CIRFT to test SNF rods gen-
erated many interesting data [7], a number of important issues that arise from
the testing system and the rod itself—including fatigue and failure mechanisms
of SNF rods—remain to be addressed. It has been shown that a direct examina-
tion of these issues in a hot cell is prohibitive because of high cost and limited
access.
The mechanical behavior of SNF rods can be effectively studied using surrogate
rods by exploiting similar methods of controlling fatigue and failure mechanisms
between the surrogate rods and the actual rods. In this investigation, surrogate nu-
clear fuel rods made of stainless steel (SS) 304 cladding and aluminum oxide or alu-
mina pellets are used to study the related fatigue and failure behaviors and
mechanisms. After a brief introduction to the experimental technique, test results
and discussion will be provided.

Experimental Technique
TESTING SETUP
The bending test system is composed of a U-frame testing setup for applying
bending loads on the spent fuel rod test specimen and a unit for measuring the
curvature of the rod during bending. Dual linear motors (Bose ElectroForce Sys-
tem Dual LM2 TB, MN) [8] are used to apply the forces symmetrically, and line-
ar variable displacement transformers (LVDTs) and load cells are installed to
measure/control displacements (disp1 and disp2) and loads (load1 and load2) at
the respective loading points. The U-frame setup is mounted to a breadboard or
reaction base along with the dual linear motors. The use of the U-frame setup
converts the forces at two loading points into a moment applied to the rod speci-
men. The rod is coupled to the U-frame using two rigid sleeves or grips. A com-
pliant layer made of cast epoxy is used between the grips and the rod to protect
the rod specimen from any contact damage. The deformation of the rod is mea-
sured by three LVDTs; using three LVDTs eliminates the effect of the epoxy layer
on the deformation measurement, which would be significant if a single LVDT
were used. The main components of the testing system are shown in Fig. 1 along
with the installed rod specimen that is located ahead of three LVDT probes, as
indicated by the arrow. The sign designation of the curvature used in this study
is shown in Fig. 2.

SPECIMEN PREPARATION
The model material for claddings is SS 304 tubes with an inner diameter (ID) of
9.708 mm, outer diameter (OD) of 11.07 mm, and length of 152.40 mm. SS claddings
22 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 1 (a) U-frame setup integrated to Bose dual LM2 TB and (b) the enlarged view of
the specimen section with three LVDTs mounted to measure the deflections of
the rod at three points.

load1, load2 motor1, motor2


with two LVDTs
for disp1, disp2
(inside covers)

Loading arms
of U-frame

(a)
Rod specimen

Specimen
grips

(b) Three LVDTs to measure


deflections of rod

FIG. 2 Schematic illustration of bending of rod specimen under loads or displacements


from the dual linear motors and sign designation of curvature.

(a) (b)
WANG AND WANG, DOI 10.1520/STP159820160051 23

FIG. 3 Schematic showing model rod when (a) uncapped and (b) capped. (Drawing is
not to scale.) PCI ¼ pellet-to-cladding interface; PPI ¼ pellet-to-pellet interface;
SS ¼ stainless steel.

(a)

(b)

are prepared from commercial SS 304 tubes purchased from McMaster-Carr.5 The
original OD and ID are modified using turning and reaming. Model fuel pellets are
short alumina cylinders 9.53 mm in diameter and 15.24 mm in length, prepared
from high-temperature nonporous alumina rods also purchased from McMaster-
Carr. The model materials for SNF rod components and the geometrical sizes of
claddings and pellets are selected to simulate the SNF rod specimens in hot-cell
tests [4].
The pellet–cladding interface (PCI) and pellet-pellet interface (PPI) are filled
with cast epoxy (3M DP420, MN). The same epoxy is used in mounting the grips
onto the rod. A 24-h curing period generally is needed to allow the epoxy used in
this study to reach full strength. A schematic showing the structure of the surrogate
rod is given in Fig. 3.

TESTING PROCEDURE
Both monotonic testing and reversed cyclic testing are performed. The monotonic
testing is conducted under displacement control. The displacement of each motor is
set at a rate of 0.2 mm/s to 10.00 mm and back to 0 mm at the same rate.
The fully reversed cyclic testing consists of measurements and cycling. The
measurements are conducted at various numbers of cycles using three cycles of
0.05-Hz sine waves at predetermined amplitudes. The cycling is conducted under
load control using a 5-Hz sine wave. The selection of load amplitudes for cycling is

5
Certain commercial equipment, instruments, or materials are identified in this paper to specify the experi-
mental procedure adequately. Such identification is not intended to imply recommendation or endorsement
by Oak Ridge National Laboratory, nor is it intended to imply that the materials or equipment identified are
necessarily the best available for the purpose.
24 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 4 Schematic showing the curvature determination of a rod by using deflections


measured at three points.

based on the results of the monotonic test. Finally, the cyclic test stops whenever
the following events occur: (1) the disp1 or disp2 is out of predetermined limits
(64 to 66 mm) or (2) the cycle number exceeds one or two million.
After the fatigue tests, all specimens are examined using an optical microscope
(Nikon Nomarski Measure Scope MM-11, Tokyo, Japan), and then fractography is
conducted on the selected specimens.

DATA ANALYSIS
The moment is calculated using Eq 1:
M ¼ FL (1)

where F is the averaged value of applied loads (load1 and load2), and L is the loading
arm length, 101.60 mm. Given the deflections from three LVDTs, d1, d2, and d3, as
shown in Fig. 4, the curvature j of the bending rod can be evaluated through a three-
point circle equation, Eq 2, as follows:
 1=2
j ¼ x02 þ ðy0 þ d2 Þ2 ; (2)

where x0 and y0 are the coordinates of center of the circle and given by
 
1 d1 þ d2 h
x0 ¼ y0 þ  ;
ma 2 2
2ma mb h þ ma ðd2 þ d3 Þ  mb ðd1 þ d2 Þ
y0 ¼ ;
2ðmb  ma Þ

and
h
ma ¼ ;
d2  d1
h
mb ¼ ;
d3  d2
WANG AND WANG, DOI 10.1520/STP159820160051 25

where h is the sensor spacing, 12 mm. The curvature generated by compressive loads
is designated as negative, and the curvature generated by tensile loads is designated
as positive (Fig. 2). As a result, the tension is on the y side (below the neutral axis)
for a negative curvature and on the þy side (above the neutral axis) for a positive
curvature.
The moment-curvature loop or M-j loop (Fig. 5) can be characterized by moment
range DM and curvature range Dj, flexural rigidity R, and flexural hysteresis UM . The
latter two are defined by Eqs 3 and 4:

R ¼ DM=Dj; (3)
þ
UM ¼ Mdj: (4)

In fully reversed bending, the ratio of the minimum moment to the maximum
moment is equal to 1. So,

DM ¼ 2Ma ; (5)
Dj ¼ 2ja ; (6)

where Ma and ja are the moment and curvature amplitudes, respectively. The similar
quantities to those in Eq 5 and Eq 6 were used in a recent work and shown to be quite
effective [9]. The hysteresis results from the curvature phase delay to the moment,

FIG. 5 Quantities used in the characterization of the moment-curvature (M-j) loop.

UM = ∫ Mdk

DM

Dk
26 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

being mainly associated with the energy dissipation attributed to the damages to the
rod induced by cyclic loading.

Experimental Results
STATIC TESTS
Two surrogate rods (#19 and #20) are monotonically loaded according to the
procedure suggested in the previous section, and both survive loading to the
maximum displacement without failure. The load and deflection data are proc-
essed using Eqs 1 and 2; the M-j plots obtained for the two rods are similar to
each other, as shown in Fig. 6. The plots usually begin with a linear response and
then exhibit a nonlinear response starting from approximately 25 Nm (C). A de-
tailed examination indicates a slight slope change around 10 Nm (B). Apparently,
neither rod attains the ultimate strength of the materials at the maximum curva-
ture (D). The subsequent unloading reveals a slope similar to that of the second
linear stage of the plots and a significant amount of residual curvature (E). The
transient point from the linear to the nonlinear stage is significant because it sig-
nifies that an important damage mechanism has been activated. These results
provide original input to the selection of load amplitude in the cyclic bending
tests. The M-j responses reveal that the surrogate rods, including the various
characteristic points and large residual curvature tested, in fact resemble those

FIG. 6 Moment-curvature (M-j) curves based on monotonic loading where the


characteristic points are labeled.

A E
WANG AND WANG, DOI 10.1520/STP159820160051 27

observed in the testing of SNF [7]; thus they partially justify the use of the pro-
posed surrogate rods to investigate SNF fatigue in the equivalent loading condition.

CYCLIC TESTS
Three moment amplitudes are selected for the reversed cyclic bending tests: 20.32,
25.40, and 30.48 Nm. Three or five rods are tested at each of these amplitudes.
Three rods (#13, #21, and #26) tested at 20.32 Nm survive more than 2.6  106
cycles without failure. All other rods are tested to failure and, as expected, the num-
ber of cycles to failure depends on the moment amplitudes, as shown in Fig. 7.
Online monitoring is enabled to capture the variation in curvature during test-
ing. Measurement data sets are acquired when a test reaches a designated number
of cycles and is interrupted. However, the following discussion will focus more on
the data obtained from online monitoring. The fatigue responses for flexural rigidi-
ty and hysteresis based on Eqs 3 and 4 are given in Fig. 8.
The fatigue life of the rods depends on their initial condition and the level of
moment amplitude applied. As can be seen, there is a large variation in the condi-
tion of the rods, as measured by the pre-fatigue rigidity. For example, the initial ri-
gidity level for the rods tested at 20.32 Nm vary widely between 80 and 95 Nm2. A
similar degree of variation is found for other amplitudes. The curvature at cycle
zero generally cannot be captured during a test, so the initial rigidity is evaluated

FIG. 7 Moment amplitude as a function of cycles or cycles to failure. Tests stopped


with no failure (#13, #21, and #26) are indicated by arrows. The prediction line is
based on the damage model, as will be discussed in the later section on fatigue
life prediction.
28 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 8 Variations in (a) flexural rigidity and (b) flexural hysteresis as a function of cycle
number. Moment amplitudes are shown.

20.32 Nm

30.48 Nm 25.40 Nm

10

(a)

30.48 Nm

25.40 Nm

20.32 Nm

(b)

at the low cycle numbers, such as cycle number one. The variations in the initial
condition of the rods are associated with the materials and the preparation of the
rod specimens. Nevertheless, the range of initial flexural rigidity obtained is similar
to the effective rigidity estimated based on the contributions of the components of
WANG AND WANG, DOI 10.1520/STP159820160051 29

FIG. 9 Relationship between flexural rigidity and flexural hysteresis obtained by


reversed cyclic bending tests.

30.48 Nm
20.32 Nm

25.40 Nm

the rods (cladding and pellets) [4]. It is observed that the degradation rate of the
rigidity depends on the level of moment amplitude. A higher moment means a
more significant degradation, as expected. At a lower level, such as 20.32 Nm, the
decreasing trend is quite appreciable, even though large local fluctuations are
involved.
It is interesting to see that the flexural hysteresis usually is quite noticeable in
the pre-fatigue stage, even at a moment as low as 20.32 Nm. This should be under-
standable because such a level of moment is nonetheless higher than at the critical
point (B) where the slope of the M-j curve changes. The subsequent cycling is
shown to have enhanced the level of flexural hysteresis. A larger moment corre-
sponds to a more significant enhancement of the hysteresis. A correlation is clearly
seen between the decrease in flexural rigidity and the increase in flexural hysteresis,
as illustrated in Fig. 9.
For the rods tested to failure, the failure usually occurs near a PPI. It is apparent
that there is a substantial stress concentration within the SS cladding arising from
the pellet cladding mechanical interaction (PCMI) near the PPI. It is the stress con-
centration that induces cracking of the SS cladding. The crack initiation sites are
seen near the maximum stress points over the cross-sections of the bending rods
and can be located at either the internal surface (#11) or the external surface of the
cladding (#23), as shown in Fig. 10. Obviously, a detailed microstructure of initiation
site cannot be obtained using the optical microscope at such magnification but
30 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 10 (a) Crack initiation site, as indicated by the arrow near the internal surface of
the cladding for rod #11 tested under 620.32 Nm Nf ¼ 6.78  105 cycles;
magnification 40. (b) Crack initiation site as indicated by the arrow on the
external surface of the cladding for rod #23 tested under 625.40 Nm,
Nf ¼ 1.96  104 cycles; magnification 40.

SS cladding

End face of
alumina
pellet

Debris from
locally crushed
alumina

(a) (b)

needs an advanced analysis such as scanning electron microscopy. On the external


surface of rod #23, the cracking initiation site is found to align with the machining
marks quite well, indicating a machining flaw is the origin of failure. However, note
that the internal surface initiation of cracking is not necessarily related to low-
moment cycling, nor is the external surface initiation related to intermediate or
high-moment cycling. In fact, if the stress intensity factor amplitude of a flaw
exceeds a threshold value of the cladding material [10], that flaw will be activated to
propagate; the location of the potential flaws are mostly a result of material fabrica-
tion processes. Cracking can also be seen near the two maximum-stress areas on
the end face of the alumina pellet. The debris resulting from the locally crushed alu-
mina can be observed near the cracking initiation site on the SS cladding.
Note that the PPI-dominated failure and the role of PCMI in the failure of clad-
ding revealed in this study are very similar to those observed in SNF rods in hot-
cell tests [7]. The similarity of the failure modes between surrogate rods and real
SNF rods is another important proof of the feasibility of using surrogate rods to
study the fatigue failure mechanism and fatigue life prediction in SNF rods.

Discussion
FATIGUE DAMAGE CHARACTERIZATION
Fatigue damage in composites can be modeled using the reduction in stiffness. The
stiffness is characterized in different ways: a modulus term based on a stress/strain
WANG AND WANG, DOI 10.1520/STP159820160051 31

graph [11] or a fatigue modulus based on the resultant strain at the applied stress
level [12]. For the surrogate rods—in which the structural heterogeneities such as
PPIs, PCIs, and the pellets themselves are at the same scale as the rods—the charac-
terization of the fatigue-damaged rod, D, can be described in terms of flexural rigid-
ity R as follows:

R
D¼1 ; (7)
R0

where R0 stands for the initial rigidity. This approach is straightforward because it
links the moment and the curvature that are used in the application. Because the rigid-
ity is being monitored online and the data are acquired for a typical cyclic test
(Fig. 8a), the variation in the damage can be evaluated accordingly.
Several damage models have been proposed for the composite structures
[13–16]. Our experimental observation shows that, for the surrogate rods, D can be
expressed effectively as a power law function of cycle number N,

D ¼ AN C ; (8)

where A and C are model parameters, and A ¼ DjN¼1 . Eq 8 can be converted into a
linear equation by using logarithms on both sides:

log D ¼ a1 log N þ a2 ; (9)

where

a1 ¼ C;
a2 ¼ log A:

The parameter a1 describes the increase rate and a2 the initial state (N ¼ 1) of
fatigue damage, both of which can be obtained using curve-fitting. The variations of
D as a function of N were similar for the rods under tested moment amplitudes. A
representative curve is shown in Fig. 11 along with the curve fitting.
The curve-fitting parameters are apparently dependent on applied moment am-
plitude Ma, as shown in Fig. 12. A linear relation can be defined between a1 or a2
and the moment amplitude:

a1 ¼ a11 Ma þ a12 ; (10a)


a2 ¼ a21 Ma þ a22 ; (10b)

where aij (i,j ¼ 1,2) are the secondary curve-fitting parameters, and the subscripts i, j
represent the fitting for the primary parameters in Eq 9. The relation of the initial
damage value, as measured by a2, to the moment in Eq 10b exists because the rigidity
depends upon the moment once loading enters into the nonlinear stage (Fig. 6). In the
latter case, a higher moment means a lower rigidity R0 and a lower level of initial fa-
tigue damage, according to Eq 7.
32 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 11 Damage evolution of rod during the cyclic test for #15 tested under 25.40 Nm;
fitting parameters are shown.

FIG. 12 Damage model parameters as a function of applied moment amplitudes where


curve-fitting parameters are given.
WANG AND WANG, DOI 10.1520/STP159820160051 33

FAILURE CRITERION
The knowledge of fatigue damage as obtained here can be used to predict fatigue
life. A fatigue failure criterion needs to be established; in previous studies, static test
results were used for such a purpose [12,13].
However, the current static setup cannot test the specimen to failure because of
limited displacement capacity, and an alternative approach is explored. Let us con-
sider the fatigue damage at failure Df, or,

Rf
Df ¼ 1  ; (11)
R0

where Rf is the rigidity at failure. The study shows that Df, as defined in Eq 11, is actu-
ally not equal to 1 at failure; rather, its valued depends on the applied moment ampli-
tude. A linear relation can be obtained between these two quantities as follows,

Df ¼ b1 Ma þ b2 ; (12)

where b1 and b2 are the curve-fitting parameters. It has been observed that, given the
condition of the rod, the damage at failure, Df, increases with the applied moment am-
plitude (Fig. 13)—namely, a more significant drop in R at failure as seen in the rigidity
fatigue curves in Fig. 8a. The determination of parameters for Df in Eq 12 uses an ex-
panded data set by including the cyclic tests under 20.32 Nm that are stopped but
with no failure to enhance the estimated confidence level. The application of the
parameters obtained here to the fatigue life analysis may provide a conservative
estimate.

FIG. 13 Damage at failure, Df, as a function of moment amplitude.


34 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 14 Flow chart for fatigue life prediction of rod under reversed cyclic bending;
R ¼ flexural rigidity and D ¼ fatigue damage.

FATIGUE LIFE PREDICTION


A fatigue life prediction approach can be developed as shown in Fig. 14. It begins
with a given rod, as characterized by R0, and applied moment amplitude, Ma. At a
specified number of cycles, the damage can be evaluated using Eqs 9 and 10, and the
current damage value can be compared with the critical level to see if it reaches a fail-
ure condition. The predicted results are given in Fig. 7, where the prediction line
agrees well with the test results. With all the failure mechanisms of SNF captured, we
believe that the analysis approach developed by testing surrogate rods can be equiva-
lently applied to the life analysis of SNF rods under similar loading conditions.

Conclusions
The mechanical behavior of SNF rods under reversed cyclic bending or bending fa-
tigue needs to be understood in order to evaluate the vibration integrity of SNF in
a transportation environment. It has been demonstrated that the bending fatigue
of SNF rods can be effectively studied using surrogate rods because of the similari-
ty of controlling failure mechanisms in surrogate and in actual rods. Surrogate
rods made of SS 304 cladding and alumina pellets are tested under moment con-
trol at a variety of amplitude levels at 5 Hz. The behavior of the rods is further
characterized by measuring flexural rigidity and hysteresis, and fractography is
performed on the failed rods. The following conclusions can be drawn based on
the tests and analysis:
1. The surrogate rods capture the main characteristics of deformation and failure
modes as observed in SNF, as validated by the static tests and fractography.
The important features include the linear-to-nonlinear deformation transition
WANG AND WANG, DOI 10.1520/STP159820160051 35

and large residual curvature in the static tests and the PPI and PCMI control
of failure mechanisms.
2. The range of initial flexural rigidity observed even for a given moment signi-
fies the variable initial conditions associated with materials and the prepara-
tion of rod specimens. Such a range of variation simulates the uncertainty
regarding the initial conditions of SNF rods related to the interfaces and
cladding.
3. The degradation of rods can be measured and characterized using flexural ri-
gidity data and the rigidity degradation dependent on both the moment ampli-
tude and the initial structural condition of the rods.
4. A correlation exists between flexural rigidity and flexural hysteresis. Therefore,
the use of either quantity should characterize equivalently the degradation of
rods induced by cyclic fatigue.
5. The cracking initiation site can be located on either the internal or the external
surface of the cladding. A structural flaw can serve as a cracking initiation site
and may be subjected to further propagation, depending on whether the stress
intensity factor amplitude exceeds the threshold value of the stress intensity
factor amplitude of the cladding material.
6. The fatigue damage to the bending rods can be described in terms of flexural
rigidity, and a power law function is effective to characterize the damage evo-
lution. The fatigue life of rods can be predicted once damage model parame-
ters are properly evaluated.

ACKNOWLEDGMENTS
The research was jointly sponsored by the Office of Nuclear Regulatory Research, U.S.
Nuclear Regulatory Commission (NRC), and the U.S. Department of Energy (DOE)
Used Fuel Disposition Campaign programs under DOE contract DE-AC05-
00OR22725 with UT-Battelle, LLC.
The authors would like to thank NRC program managers Michelle Bales and Pat-
rick Raynaud, ORNL program managers Bruce Bevard and Rob Howard, and Pacific
Northwest National Laboratory program manager Harold Adkins for providing guid-
ance and support. The authors would also like to thank Dr. Lianshan Lin of ORNL for
reviewing this manuscript.

References

[1] Hanson, B., Alsaed, H., Stockman, C., Enos, D., Meyer, R., and Sorenson, K., “Gap Analy-
sis to Support Extended Storage of Used Nuclear Fuel,” FCRD-2011-000136 Rev. 0, U.S.
Department of Energy Used Fuel Disposition Campaign, U.S. DOE, Washington,
DC, 2012.

[2] Wang, J.-A. J., Wang, H., Tan, T., Jiang, H., Cox, T., and Yan, Y., “Progress Letter Report
on U Frame Test Setup and Bending Fatigue Test for Vibration Integrity Study (Out-of-
Cell Fatigue Testing Development–Task 2.2),” ORNL/TM-2011/531, Oak Ridge National
Laboratory, Oak Ridge, TN, 2012.
36 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

[3] Wang, J.-A. J., Wang, H., Cox, T., and Yan, Y., “Progress Letter Report on U-Frame Test
Setup and Bending Fatigue Test for Vibration Integrity Study (Out-of-Cell Fatigue Test-
ing Development–Task 2.3),” ORNL/TM-2012/417, Oak Ridge National Laboratory, Oak
Ridge, TN, 2012.

[4] Wang, H., Wang, J.-A. J., Tan, T., Jiang, T. H., Cox, T. S., Howard, R. L., Bevard, B. B., and
Flanagan, M. E., “Development of U-frame Bending System for Studying the Vibration
Integrity of Spent Nuclear Fuel,” J. of Nucl. Mater., Vol. 440, Nos. 1–3, 2013, pp. 201–213.

[5] Wang, J.-A. J., Wang, H., Cox, T., and Baldwin, C., “Progress Letter Report on Bending
Fatigue Test System Development for Spent Nuclear Fuel Vibration Integrity Study
(Out-of-Cell Fatigue Testing Development–Task 2.4),” ORNL/TM-2013/225, Oak Ridge
National Laboratory, Oak Ridge, TN, 2013.

[6] Wang, J.-A. J., Wang, H., and Tan, T., 2014. Reversal bending fatigue testing. U.S. Patent
8,863,585 B2, filed February 14, 2012, and issued October 21, 2014.

[7] Wang, J.-A. and Wang, H., “Mechanical Fatigue Testing of High-Burnup Fuel for Trans-
portation Applications,” NUREG/CR-7198/ ORNL/TM-2014/214, Office of Nuclear Regu-
latory Research, U.S. Nuclear Regulatory Commission, Washington, DC, 2015.

[8] ElectroForce System Group, ElectroForce LM2 TestBench, Test Instrument Reference
Manual, Bose Corp., Eden Prairie, MN, 2012.

[9] Wang, H. and Wang, J.-A. J., “Bending Testing and Characterization of Surrogate Nucle-
ar Fuel Rods Made of Zircaloy-4 Cladding and Aluminum Oxide Pellets,” J. of Nucl.
Mater., Vol. 479, 2016, pp. 470–482.

[10] Anderson, T. L., Fracture Mechanics Fundamentals and Applications, CRC Press, Boca
Raton, FL, 1991, pp. 606–619.

[11] Yang, J. N., Jones, D. L., Yang, S. H., and Meskini, A., “A Stiffness Degradation Model for
Graphite/Epoxy Laminates,” J. Compos. Mater., Vol. 24, No. 7, 1990, pp. 753–769.

[12] Hwang, W. and Han, K. S., “Fatigue of Composites—Fatigue Modulus Concept and Life
Prediction,” J. Compos. Mater., Vol. 20, No. 2, 1986, pp. 156–165.

[13] Clark, S. D., Shenoi, R. A., and Allen, H. G., “Modelling the Fatigue Behavior of Sandwich
Beams under Monotonic, 2-Step and Block-Loading Regimes,” Compos Sci. Technol.,
Vol. 59, No. 4, 1999, pp. 471–486.

[14] Mao, H. and Mahadevan, S., “Fatigue Damage Modelling of Composite Materials,”
Compos. Struct., Vol. 58, No. 4, 2002, pp. 405–410.

[15] Abbadi, A., Azari, Z., Belouettar, S., Gilgert, J., and Freres, P., “Modelling the Fatigue
Behaviour of Composites Honeycomb Materials (Aluminium/Aramide Fibre Core) using
Four-point Bending Tests,” Int. J. Fatigue, Vol. 32, No. 11, 2010, pp. 1739–1747.

[16] Giancane, S., Panella, F. W., and Dattoma, V., “Characterization of Fatigue Damage in
Long Fiber Epoxy Composite Laminates,” Int. J. Fatigue, Vol. 32, No. 1, 2010, pp. 46–53.
FATIGUE AND FRACTURE TEST PLANNING, TEST DATA ACQUISITIONS AND ANALYSIS 37

STP 1598, 2017 / available online at www.astm.org / doi: 10.1520/STP159820160030

Xijia Wu,1 Guangchun Quan,2 and Clayton Sloss3

Low Cycle Fatigue of Cast


Austenitic Steel
Citation
Wu, X., Quan, G., and Sloss, C., “Low Cycle Fatigue of Cast Austenitic Steel,” Fatigue and
Fracture Test Planning, Test Data Acquisitions and Analysis, ASTM STP1598, Z. Wei, K. Nikbin,
P. C. McKeighan, and D. G. Harlow, Eds., ASTM International, West Conshohocken, PA,
2017, pp. 37–57, http://dx.doi.org/10.1520/STP1598201600304

ABSTRACT
Cast austenitic stainless steel 1.4848 is used to manufacture automotive exhaust
system components. Low cycle fatigue (LCF) of 1.4848 austenitic steel was
investigated through strain-controlled fatigue testing at strain rates of 0.02/s,
0.002/s, and 0.0002/s in the temperature range from room temperature (RT) to
900 C. Its cyclic behavior was characterized in relation to deformation
mechanisms. At RT, the material behavior was rate-independent and cyclically
stable, which occurred by plasticity. The material exhibited significant cyclic
hardening at intermediate temperatures, 400 C to 600 C, with negative strain-
rate sensitivity. In this temperature range, dynamic strain aging (DSA)
presumably occurred due to slip dragging solute atoms. At high temperatures,
800 C and 900 C, the material exhibited positive rate-dependence in the
hysteresis behavior, and the cyclic stress response tended to stabilize with
increasing cycles. The high-temperature behavior was presumably controlled by
a combination of plasticity and dislocation-glide creep. The integrated creep-
fatigue theory (ICFT) was used to describe the deformation and life behaviors
based on the identified mechanisms, which were corroborated by fractographic
observations.

Manuscript received February 10, 2016; accepted for publication June 17, 2016.
1
National Research Council Canada, 1200 Montreal Rd., Ottawa, Ontario, K1A 0R6, Canada http://
orcid.org/0000-0002-0250-112X
2
Tenneco Automotive Operating Co., Inc., 3901 Willis Rd., Grass Lake, MI 49240 http://orcid.org/
0000-0002-8132-7202
3
Wescast Industries Inc., 150 Savannah Oaks Dr., Brantford, Ontario, N3T 1L8, Canada http://orcid.org/
0000-0002-4468-936X
4
ASTM Symposium on Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis on
May 4–5, 2016 in Grand Hyatt, San Antonio, TX.

Copyright V
C 2017 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.
38 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

Keywords
low cycle fatigue, austenitic stainless steel, damage mechanism, damage
modeling, life prediction

Introduction
In the automotive industry, austenitic and ferritic cast stainless steels are widely used
for manufacturing exhaust components to withstand higher gas temperatures exceed-
ing 1,000 C. To ensure satisfactory durability under such severe operating conditions,
low cycle fatigue (LCF) and thermomechanical fatigue (TMF) behaviors of these mate-
rials need to be fully characterized. In this regard, even though the strain-controlled
fatigue test standard—ASTM E606, Standard Test Method for Strain-Controlled Fa-
tigue Testing—has been well established, there are still challenges about interpretation
and analysis of the test data, especially the life prediction methods for stainless steels.
Wrought austenitic stainless steels (e.g., 316L) have been found to exhibit non-
stabilizing cyclic stress responses [1,2]. Cyclic hardening was found to be associated
with dynamic strain aging (DSA) [1,2]. However, the conventional methods of fa-
tigue analysis via the Coffin–Manson relation [3] and energy-based correlations [4]
are based on the assumption that the material’s cyclic behavior would be stabilized
(or saturated), at least after a short period of initial cycling, usually within several
dozens of cycles. Such an assumption is not valid for austenitic stainless steels.
When the material is cyclically unstable, it imposes two serious questions on fatigue
assessment for a component made of such material: (1) What stress-strain relation-
ship should be chosen for stress analysis of the component experiencing stress
changes cycle by cycle? (2) At which stage should the stress/strain amplitude or hys-
teresis energy be determined for predicting the fatigue life?
So far, limited studies have been conducted on LCF and TMF properties of austenitic
cast stainless steels. Kim and coworkers studied LCF of two cast austenitic stainless steels
[5,6]. They correlated the experimental fatigue life with the total controlled strain range.
Seifert et al. studied the LCF and TMF behavior of a German designated alloy—1.4849
cast austenitic steel; they analyzed the fatigue life with the presumed stable stress-strain
relationship [7]. The effects of DSA and environment on fatigue life of austenitic stainless
steels have not been well understood. The current fatigue analysis method, as described
in the appendix of ASTM E606, does not seem to provide further delineation.
In this study, LCF tests were carried out on 1.4848 cast austenitic steel at three
strain rates—0.02/s, 0.002/s, and 0.0002/s—in a temperature range from 20 C to
900 C. The cyclic stress-strain/hysteresis behaviors are discussed with respect to the
underlying deformation mechanisms. The LCF life is formulated using the integrat-
ed creep-fatigue theory (ICFT), which recognizes the specific role of each mecha-
nism such as plasticity, DSA, creep, and oxidation in the holistic damage
accumulation process consisting of nucleation and propagation of surface cracks in
coalescence with internally distributed damage.
WU ET AL., DOI 10.1520/STP159820160030 39

Experimental
Austenitic cast stainless steel 1.4848 was used in this study to make material coupons.
Its chemical composition is given in Table 1. Strain-controlled LCF tests were conducted
on a material testing systems (MTS) uniaxial servohydraulic test frame, model MTS
810, equipped with an applied test systems (ATS) series 3210 three-zone radiation fur-
nace. The load frame was equipped with an MTS 661 series load cell and MTS Model
609 alignment fixture. Specimen gripping was achieved with MTS 680 hydraulic grips
using threaded grip inserts with a unified thread standard (UNF) designation 5/8
18 UNF (UNF designates fine thread density). During the test, the temperature was
controlled using three K-type thermocouples with fixed mounting positions within the
furnace hot zone. Closed-loop control of the specimen loading was achieved by mea-
suring the elongation of the specimen with an MTS Model 632.54 extensometer and
using an MTS Model GT493 digital controller regulating the force applied to the speci-
men. The load applied to the specimen was monitored with an MTS Model 661 load
cell. The furnace alarm circuit was coupled to the test controller via an external relay,
allowing the test controller to disengage power to the furnace upon specimen rupture.
Fatigue testing followed ASTM E606-04, and it was carried out at strain rates of
0.02/s, 0.002/s, and 0.0002/s in the temperature range of 20 C to 900 C. Postmor-
tem factographic examination of the fracture surfaces was conducted using a Philips
XL30S scanning electron microscope (SEM).

Results and Discussion


CYCLIC BEHAVIOR
Without further specification, the following discussion of cyclic stress-strain behav-
ior is in regard to engineering stress-strain, which is not much different from the
true-stress-strain behavior under cyclic conditions with strain range less than 2 %.
The room temperature (RT) cyclic behaviors of austenitic cast stainless steel 1.4848
at strain rate of e_ ¼ 0:02=s are shown in Fig. 1. The half-life hysteresis loops are shown
in Fig. 2 with the zero points setting at the maximum compression state. The half-life
hysteresis loop of 1 % strain range at strain rates of e_ ¼ 0:02=s is also shown in Fig. 2
for comparison. The experimental data have shown that, after an initial period of ap-
proximately 200 cycles at RT, the cyclic behavior of this material becomes stable, rate-
independent, and Masing, which means that there is a unique stress-strain relationship
that describes the material’s cyclic behavior at all strain ranges.

TABLE 1 Composition of 1.4848 austenitic steel (wt%).

C Si Mn Ni Cr Fe

0.35 1.5 0.6 20 25 Bal.

Note: C ¼ carbon; Si ¼ silicon; Mn ¼ manganese; Ni ¼ nickel; Cr ¼ chromium; Fe ¼ iron.


40 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 1 Cyclic peak-valley stresses of austenitic cast steel at RT.

FIG. 2 Stabilized hysteresis loops of austenitic cast steel at room temperature.


WU ET AL., DOI 10.1520/STP159820160030 41

FIG. 3 Cyclic peak-valley stresses of austenitic cast steel at 400 C.

At 400 C, however, the peak-valley stress responses of the material exhibited con-
tinuous cyclic hardening until fracture, as shown in Fig. 3. This could be attributed to
dynamic strain hardening, as was observed in type 316L austenitic stainless steel in the
temperature range of 250 C to 550 C [1,2]. DSA is a phenomenon of interaction be-
tween diffusing solute atoms and mobile dislocations during plastic deformation. It is
manifested by pinning of slow-moving dislocations, restricting the cross-slip of screw
dislocations, and hence enhancing slip inhomogeneity, resulting in an increased stress
(hardening) to reach the prescribed strain level. In this temperature range, deforma-
tion-induced martensitic transformation could not be responsible for the observed phe-
nomena because it would mostly occur with large deformation at low temperatures
(<20 C, i.e., under “cold working” conditions) [8]. Interestingly, in this cast austenitic
steel, serrated plastic flow was not observed as in other wrought austenitic steels previ-
ously studied. This could be due to the difference in the initial dislocation structures of
the cast and the wrought materials. In the wrought material, the local high-dislocation
density may boost the plastic strain rate, exceeding the imposed strain rate and thus
causing the stress drop [9]. Furthermore, comparing the mid-life behavior, 1.4848 aus-
tenitic cast stainless steel was observed to experience negative strain-rate sensitivity.
Hence, DSA is believed to be responsible for the cyclic hardening behavior in the inter-
mediate temperature range. The mid-life hysteresis loops of the cast austenitic steel
loaded at strain rates of e_ ¼ 0:02=s and 0.002/s at 400 C are shown in Fig. 4a and b, re-
spectively. The scatter bar indicates the stress range from initial to final states of cycling.
At the high strain rate of e_ ¼ 0:02=s, the cyclic stress-strain behavior of this material
42 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 4 Mid-life hysteresis behaviors of austenitic cast steel at 400 C, (a) at strain rate
of 0.02/s and (b) at strain rate of 0.002/s.

(a)

(b)
WU ET AL., DOI 10.1520/STP159820160030 43

FIG. 5 Cyclic peak-valley stresses of austenitic cast steel at 600 C.

was nearly Masing, whereas at the lower strain rate, e_ ¼ 0:02=s, the DSA effect was
stronger and the cyclic stress-strain behavior became non-Masing. The non-Masing be-
havior means that the individual hysteresis behavior does not generally follow the cyclic
stress-strain relationship. Cyclic behavior at 600 C was similar to that at 400 C but
with stronger negative strain-rate sensitivity and more significant cyclic hardening at all
strain rates, as shown in Figs. 5 and 6.
The cyclic stress behaviors of the material at different strain rates at 800 C are
shown in Fig. 7. Even though the initial response appeared to be rate-dependent, after a
short number of cycles, the material became cyclically stabilized. It is suspected that, at
this temperature, the effect of DSA was balanced by the positive strain-rate sensitivity
of creep deformation. The mid-life hysteresis loops of the austenitic cast steel at the
strain rate of e_ ¼ 0:02=s and 0.0002/s at this temperature are shown in Fig. 8. The cy-
clic stress-strain behavior of this austenitic steel at 800 C is highly non-Masing.
At 900 C, the material appeared to be cyclically stable right from the beginning and,
in some cases, it exhibited slight cyclic softening, as shown in Fig. 9. The mid-life hystere-
sis loops of the material at this temperature are shown in Fig. 10a and b for the two slow
strain rates, 0.002/s and 0.0002/s, respectively. At this temperature, the deformation
behavior showed positive strain-rate sensitivity due to time-dependent creep defor-
mation. The cyclic stress-strain behavior of this austenitic steel at 900 C appears to
be closely Masing. The slight variations could be due to material variability.
For LCF life analysis, conventionally, one would use the Ramberg-Osgood equa-
tion to fit the cyclic stress-strain data to evaluate the inelastic strain for correlation with
44 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 6 Mid-life hysteresis behaviors of austenitic cast steel at 600 C, (a) at strain rate
of 0.02/s and (b) at strain rate of 0.002/s.

(a)

(b)
WU ET AL., DOI 10.1520/STP159820160030 45

FIG. 7 Cyclic peak-valley stresses of austenitic cast steel at 800 C.

FIG. 8 Mid-life hysteresis behaviors of austenitic cast steel at 800 C.


46 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 9 Cyclic peak-valley stresses of austenitic cast steel at 900 C.

the fatigue life. However, the Ramberg-Osgood equation is basically a rate-independent


form. It has limitations in handling high-temperature fatigue data. We propose a mech-
anism-based approach for constitutive modelling, that is, the integrated creep-fatigue
theory (ICFT) [10–13]. First, the ICFT recognizes the contribution from each deforma-
tion mechanism in the physical deformation decomposition as shown:
r 
e¼ þ ep þ ev (1)
E

where:
r ¼ the applied stress,
E ¼ the elastic modulus,
ep ¼ plastic strain, and
ev ¼ creep strain.
Rate-independent plasticity can be described by a power-law, consistent with
the Ramberg-Osgood equation (Eq 1):
r  r n
0
ep ¼ (2)
K

where:
r0 ¼ the elastic limit, and
K and n ¼ material constants.
WU ET AL., DOI 10.1520/STP159820160030 47

FIG. 10 Mid-life hysteresis behaviors of austenitic cast steel at 900 C, (a) at strain rate
of 0.002/s and (b) at strain rate of 0.0002/s.

(a)

(b)
48 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

Creep deformation can be further decomposed into many mechanism compo-


nents such as dislocation glide, dislocation climb, and diffusion as classified in the
Ashby’s deformation mechanism map [14]. In the present case (under relatively
fast strain-rate controlled loading conditions), the dominant creep mechanism
occurs by thermally activated dislocation glide in interaction with the solute atoms.
Dislocation glide with linear strain hardening can be formulated in Eq 3 as [15]:
!  
DG6¼
0 Vðr  Hev  r0 Þ
e_v ¼ 2e_0 exp  sinh (3)
kT kT

where:
e_0 ¼ the proportional constant,
DG6¼ 0 ¼ the activation energy,
V ¼ the activation volume,
H ¼ the work hardening coefficient,
k ¼ the Boltzmann constant, and
T ¼ the absolute temperature.

At low temperatures, plasticity—as represented by Eq 2—is the only contributing


mechanism in Eq 1. The description of Eq 2 for the cyclic stress-strain behaviors at
RT, 400 C, and 600 C are shown in Figs 2, 4, and 6, respectively. The effect of DSA
is reflected by an increase in the plastic drag stress K. These descriptions are in
good agreement with the experimental observations.
For deformation at 800 C and 900 C, both plasticity and creep mechanisms,
represented by Eqs 2 and 3, come into play in Eq 1, and the descriptions of the
cyclic stress-strain behaviors are shown in Figs. 8 and 10, respectively. Particularly, it
is more evident that, at the low strain rate of 900 C, both the cyclic stress-strain
and the hysteresis behaviors appear to be governed by dislocation-glide creep with
linear work hardening. Work hardening is believed to be associated with dislocation
glide because climb would not cause strain hardening but rather relieve it.

LCF Life
The LCF life versus cyclic inelastic strain range relationships are shown in Fig. 11 for
1.4848 cast austenitic steel at various temperatures; all strain-rate data are included.
The Coffin-Manson relationship provides the best linear regression for each tem-
perature condition but no revelation of the contribution from each individual
mechanism.
The life of holistic damage accumulation, consisting of nucleation and propaga-
tion of surface cracks in coalescence with internally distributed damage, has been
derived by ICFT and can be expressed as Eq 4 [10–13]:
 
1 1 h
¼D þ (4)
N N f ac
WU ET AL., DOI 10.1520/STP159820160030 49

FIG. 11 The inelastic strain range versus fatigue life for 1.4848 stainless steel.

where:
Nf ¼ the pure mechanical fatigue life,
pffiffiffiffiffiffiffiffiffiffiffi
h ¼ 2kox s; kox ¼ the oxidation constant,
s ¼ the cycle period, and
ac is the critical crack length at fracture as defined by Eq 5:
 
1 KIC 2
ac ¼ (5)
p Yrmax

Y ¼ the crack shape factor, and


D ¼ a factor considering all the effects of internally distributed damage/discon-
tinuities, as expressed by Eq 6:
!
X li
D¼ 1þ (6)
i
ki

where:
li is the size of the ith damage component and
ki is their interspacing.

First of all, at RT without contributions of creep and environmental effects, Eq 4


reduces to N ¼ Nf, where Nf is given by the Coffin-Manson relation as shown in Eq 7:

Dep
¼ e0f Nfc (7)
2

The best-fit to the RT data yields that c ¼ 0.53, which is very close to the the-
oretical value of 0.5 by Tanaka and Mura’s model [16]. Since the RT cyclic behavior
50 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

TABLE 2 Parameter values for Eq 4.

Mechanism Fatigue DSA Creep Oxidation: kox ¼ k0exp(Q/RT)


0 2 2
Parameter ef c a (m ) q0 (m ) b k0 (m2/s) Q (kJ) KIC/Y
(unit) (MPaHm)
Value 0.22 0.5 2.52  1014 1.68  107 200 1.7 231 100

is stabilized and Masing, the parameters thus obtained are truly attributes of me-
chanical fatigue by rate-independent plasticity.
At 400 C and 600 C, however, the material’s cyclic behavior becomes cyclically
non-stabilized, especially more significant at low strain rate as influenced by DSA.

FIG. 12 Striations on room-temperature fatigue fracture specimen with 0.7 % strain


range.
WU ET AL., DOI 10.1520/STP159820160030 51

FIG. 13 Fracture surface of specimen fatigued with 0.7 % strain range at 0.002/s,
400 C.

Slip Band

In this case, Eq 7 does not hold true because Dep keeps evolving during the entire
fatigue process. For generality, we propose a modified form of Eq 6 as shown in Eq 8:

Dep;eq 1
¼ e0f Nf 2 (8)
2

where Dep,eq is defined as shown in Eq 9:


vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
uN
uP 2
u De p;i
t
Dep;eq ¼ i¼1 _ (9)
N

It is straightforward to prove that Eq 7 is actually Tanaka and Mura’s model under


variable strain amplitude conditions and that it is, in principle, reinstating the Miner’s rule.
52 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 14 Fracture surface of specimen fatigued with 0.7 % strain range at 0.002/s,
600 C.

For austenitic stainless steels, DSA in the medium temperature range, 400 C
to 600 C, leads to slip inhomogeneity with severe dislocation pile-ups within the con-
centrated slip bands or dislocation walls, as observed in type 316L stainless steel [17].
Imaging dislocation pile-ups are embryos of cracks; the internal damage by DSA may
be formulated to be proportional to the excess forest dislocation density. Therefore, we
may formulate the D-factor for austenitic stainless steels as shown in Eq 10:
"  #
DrH 2
D¼1þa  q0 þ bev (10)
lb

where:
DrH ¼ the amplitude of cyclic hardening (maximum attainable peak stress mi-
nus the peak stress of the first cycle),
WU ET AL., DOI 10.1520/STP159820160030 53

FIG. 15 Fracture surface of specimen fatigued with 0.7 % strain range at 0.002/s,
800 C.

l ¼ the shear modulus,


b ¼ the Burgers vector,
q0 ¼ the dislocation density level below which there is no instantaneous crack
nucleation, and
a and b ¼ proportional constants for dislocation density crack nucleation and
creep cavity nucleation respectively.

Eqs 4–10 are used to predict the LCF life with the parameter values given in Table 2.
The predictions are shown in Fig. 11 as lines.
It is seen from Fig. 11 that, under the conditions that promote DSA (i.e., at low
strain rate and high strain amplitudes), at 400 C to 600 C, Eq 10 takes effect to re-
duce the fatigue life significantly by DSA with nearly zero creep strain. As the tem-
perature increases to 800 C and 900 C, the DSA term in Eq 10 becomes negligible
54 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

because cyclic hardening is balanced by creep softening, but the creep damage term
remains in Eq 10. At high temperatures, oxidation also plays an important role. As
seen from Fig. 11, the combination of creep and oxidation does reduce the LCF life
significantly in comparison with the RT-LCF life. On one hand, creep deformation
reduces the LCF life by a factor of D; but, on the other hand, creep lowers the peak
stress, which allows the specimen to tolerate a longer crack, thus increasing the life.
Similarly, lowering the controlling strain rate allows more creep deformation and
stress relaxation, but it also increases the time of oxidation per cycle. All such con-
tradicting effects result in a rather frequency-insensitive behavior of LCF life at high
temperatures, as described by the model curves for 800 C and 900 C at different
strain rates.

FIG. 16 Fracture surface of specimen fatigued with 0.7 % strain range at 0.002/s,
900 C.
WU ET AL., DOI 10.1520/STP159820160030 55

Fractographic Observation
Using SEM, it was observed that the RT fracture surface contained typical stria-
tions that indicate transgranular fatigue fracture, as shown in Fig. 12. Fatigue
fracture at 400 C to 600 C was also transgranular; however, DSA led to a con-
centrated slip band pattern, as shown in Figs. 13 and 14. At 800 C, the fatigue
specimen failed predominantly by an intergranular fracture mode, as shown in
Fig. 15. Most of the fracture surface at 800 C was covered with oxides, except
some isolated white spots containing high nickel (Ni) content. On the fracture
surface of the specimen that failed at 900 C, extensive oxidation and formation
of voids or cavities can be observed, which was presumably due to creep effect,
as shown in Fig. 16. These fractographic observations corroborate with the model
explanation.

Conclusion
LCF behaviors of 1.4848 austenitic steel were studied through strain-controlled
cyclic tests at strain rates of 0.02/s, 0.002/s, and 0.0002/s in the temperature range
of RT to 900 C. At RT, the material exhibited a cyclically stable, rate-independent,
and Masing behavior. The material exhibited a non-Masing behavior with continu-
ous cyclic hardening due to DSA at intermediate temperatures, 400 C to 600 C. At
high temperatures, 800 C and 900 C, the positive rate-dependence of deformation
was manifested via thermally-activated dislocation glide. The cyclic deformation
behaviors of 1.4848 cast austenitic steel are well-described by the ICFT with contri-
butions of rate-independent plasticity, DSA, and thermally activated dislocation
glide quantitatively identified.
According to the ICFT, the aforementioned mechanisms also participate in the
holistic damage accumulation involving nucleation and propagation of surface fa-
tigue cracks in coalescence with internally distributed (DSA and creep) damage.
Therefore, as the limiting case, RT-LCF life is controlled by rate-independent plas-
ticity. At intermediate temperatures, 400 C to 600 C, fatigue life is reduced by DSA
creating excess dislocation densities. At high temperatures, 800 C and 900 C, oxi-
dation comes into play with creep; their combined effect is related to the ratio of ox-
idation penetration per cycle to the critical crack length.
In conclusion, the ICFT provides a mechanism-based delineation of cyclic de-
formation and life behaviors of 1.4848 cast austenitic steel. Such a mechanism de-
lineation approach may provide guidance to material and component design for
optimal performance.

ACKNOWLEDGMENTS
The work was carried out as a contract between the National Research Council of
Canada and Wescast Industries, Inc. The authors would also like to acknowledge the
technical assistance of Ryan MacNeil in material testing.
56 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

References

[1] Alain, R., Violan, P., and Mendez, J., “Low Cycle Fatigue Behavior in Vacuum of a Type
316 Austenitic Stainless Steel between 20 and 600 C—Part I: Fatigue Resistance and
Cyclic Behavior,” Mat. Sci. Eng. A, Vol. 229, Nos. 1–2, 1997, pp. 87–94.

[2] Hong, S-G., Lee, S-B., and Byun, T-S., “Temperature Effect on the Low-Cycle Fatigue
Behavior of Type 316L Stainless Steel: Cyclic Non-Stabilization and an Invariable Fatigue
Parameter,” Mat. Sci. Eng. A, Vol. 457, Nos. 1–2, 2007, pp. 139–147.

[3] Coffin, Jr., L. F., “A Study of the Effects of Cyclic Thermal Stresses on a Ductile Metal,”
Trans. ASME, Vol. 76, No. 6, 1954, pp. 931–950.

[4] Morrow, J. D., “Cyclic Plastic Strain Energy and Fatigue of Metals, Internal Friction,
Damping, and Cyclic Plasticity,” Internal Friction, Damping, and Cyclic Plasticity, ASTM
STP378, B. J. Lazan, Ed., ASTM International, West Conshohocken, PA, 1965, pp. 45–84,
http://dx.doi.org/10.1520/STP378-EB

[5] Kim, Y-J., Jang, H., and Oh, Y-J., “High Temperature Low Cycle Fatigue Properties of a
HF30-Type Cast Austenitic Stainless Steel,” Mat. Sci. Eng. A, Vol. 526, Nos. 1–2, 2009,
pp. 244–249.

[6] Kim, Y-J. and Jang, H., “High Temperature Fatigue Resistance of an ACI HH50-Type
Cast Austenitic Stainless Steel,” Mat. Sci. Eng. A, Vol. 527, Nos. 21–22, 2010,
pp. 5415–5420.

[7] Seifert, T., Schweizer, C., Schlesinger, M., Moser, M., and Eibl, M., “Thermomechanical
Fatigue of 1.4849 Cast Steel—Experiment and Life Prediction Using a Fracture Mechan-
ics Approach,” Int. J. Mat. Res., Vol. 101, No. 8, 2010, pp. 942–950.

[8] Nagy, E., Mertinger, V., Tranta, F., and Sólyom, J., “Deformation Induced Martensitic
Transformation in Stainless Steels,” Mat. Sci. Eng. A, Vol. 378, Nos. 1–2, 2004,
pp. 308–313.

[9] Rodriguez, P., “Serrated Plastic Flow,” Bull. Mater. Sci., Vol. 6, No. 4, 1984, pp. 653–663.

[10] Wu, X. J., “A Model of Nonlinear Fatigue-Creep (Dwell) Interactions,” Trans. ASME, J.
Eng. Gas Turbines Power, Vol. 131, No. 3, 2009, http://dx.doi.org/doi:10.1115/1.2982152

[11] Wu, X. J., “An Integrated Creep-Fatigue Theory for Material Damage Modeling,” Key
Eng. Mat., Vol. 627, 2015, pp. 341–344.

[12] Wu, X. J., Quan, G., MacNeil, R., Zhang, Z., and Sloss, C., “Failure Mechanisms and Dam-
age Model of Ductile Cast Iron under Low-Cycle Fatigue Conditions,” Metall. Mater.
Trans. A, Vol. 45, No. 11, pp. 5088–5097.

[13] Wu, X. J., Quan, G., MacNeil, R., Zhang, Z., Liu, X., and Sloss, C., “Thermomechanical
Fatigue of Ductile Cast Iron and Its Life Prediction,” Metall. Mater. Trans. A, Vol. 46A,
No. 6, 2014, pp. 2530–2543.

[14] Frost, H. J. and Ashby, M. F., Deformation Mechanism Maps, Pergamon, Oxford, UK, 1982.

[15] Wu, X. J. and Krausz, A. K., “A Kinetics Formulation for Low-Temperature Plasticity,”
J. Mater. Eng. Perform., Vol. 3, 1994, pp. 169–177.
WU ET AL., DOI 10.1520/STP159820160030 57

[16] Tanaka, K. and Mura, T., “A Dislocation Model for Fatigue Crack Initiation,” J. Appl.
Mech., Vol. 48, No. 1, 1981, pp. 97–103.

[17] Hong, S.-G. and Lee, S.-B., “Mechanism of Dynamic Strain Aging and Characterization
of Its Effect on the Low-Cycle Fatigue Behavior in Type 316L Stainless Steel,” J. Nucl.
Mater., Vol. 340, Nos. 2–3, 2005, pp. 307–314.
58 FATIGUE AND FRACTURE TEST PLANNING, TEST DATA ACQUISITIONS AND ANALYSIS

STP 1598, 2017 / available online at www.astm.org / doi: 10.1520/STP159820160086

Martin Bjurman,1 Björn Forssgren,2 and Pål Efsing3

Fracture Mechanical Testing of


In-Service Thermally Aged Cast
Stainless Steel
Citation
Bjurman, M., Forssgren, B., and Efsing, P., “Fracture Mechanical Testing of In-Service Thermally
Aged Cast Stainless Steel,” Fatigue and Fracture Test Planning, Test Data Acquisitions and
Analysis, ASTM STP1598, Z. Wei, K. Nikbin, P. C. McKeighan, and D. G. Harlow, Eds., ASTM
International, West Conshohocken, PA, 2017, pp. 58–80, http://dx.doi.org/10.1520/
STP1598201600864

ABSTRACT
Embrittlement of duplex stainless steels by thermal aging shortens the service
life of structural components in light water reactors (LWRs). This is an important
issue when life-extension programs are aiming at 60 to 80 years in service
because avoidance of brittle failure is a design prerequisite. Cast and welded
austenitic stainless steels, which contain some ferrite, are known to be affected
by thermal aging. Historically, many LWR components of complex geometry
have been cast in the molybdenum-containing grade, CF8M. Aging is mainly
attributed to two types of phase transformations occurring within the minor
ferritic phase: (1) demixing of the ferrite by spinodal decomposition into
chromium-rich a´ and iron-rich a regions and (2) precipitation of G-phase,
carbides, and other secondary phases. Two in-service aged pipe bend castings
from the pressurized water reactor (PWR) Ringhals 2 steam generator are
tested. These components are large castings of stainless-steel quality CF8M.
The manufacturing process produces a nonuniform microstructure with coarse
ferrite and a high degree of directionality affecting properties as well as

Manuscript received April 4, 2016; accepted for publication September 9, 2016.


1
Studsvik Nuclear AB and Royal Institute of Technology, Stockholm, S-100 44, Sweden http://orcid.org/
0000-0003-2034-5391
2
Ringhals AB, Väröbacka, SE-430 22, Sweden
3
Ringhals AB and Royal Institute of Technology, Stockholm, S-100 44, Sweden http://orcid.org/
0000-0003-1498-5691
4
ASTM Symposium on Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis on May 4–5,
2016 in Grand Hyatt, San Antonio, TX.

Copyright V
C 2017 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.
BJURMAN ET AL., DOI 10.1520/STP159820160086 59

the methodology for testing. The materials are exposed to primary circuit PWR
water for 72 kh at 291°C and 325°C, followed by 22 kh at a reduced service
temperature. The testing approach and matrix, consisting of microstructural
characterization, mechanical testing, and modeling, is discussed. Results of
fracture mechanical evaluation using the J-integral resistance curve technique
and instrumented Charpy tests are provided. Effects of large microstructural
heterogeneity and anisotropy from the casting and heat-treating processes are
seen. The effect on aging embrittlement and fracture mechanisms within each
phase as well as phase interaction are also discussed.

Keywords
thermal aging, cast austenitic stainless steel (CASS), fracture mechanical testing,
spinodal decomposition

Introduction
Thermal aging embrittlement of cast and welded “austenitic” stainless steels for
nuclear power plant (NPP) applications was extensively studied during the 1980s
and early 1990s. Thermal aging of cast austenitic stainless steel (CASS) leads to an
increase in hardness and tensile strength and a decrease in ductility, impact
strength, and fracture toughness (FT) as well as corrosion resistance. A renewed
interest in these issues has arisen in recent years, driven by life-extension programs
and the difficulty in predicting the mechanical behavior and microstructure after
60 to 80 years at reactor temperatures.
The grades used in NPPs typically are of the CF-type.5 The grades and typical
compositions are presented in Table 1. These materials have a duplex solidification
microstructure consisting of austenite and dferrite phases. The main aging phe-
nomenon is the thermal diffusion-driven decomposition of the ferrite into iron-rich
a-phase and chromium-rich a´-phase due to the miscibility gap seen in the iron-
chromium (Fe-Cr) phase diagram [1–3]. A maximum rate of decomposition occurs
at 4758C, hence the frequently used name 4758C-embrittlement. The thermal aging
phenomenon of ferritic grades from spinodal decomposition in high-temperature
applications of 280 C to 500 C is widely known [4–9]. Research has focused on
investigating binary alloys and the commercial grades used in NPPs. An additional
contribution to embrittlement comes from the precipitation of G-phase, rich in nickel
(Ni), silicon (Si), manganese (Mn), molybdenum (Mo), and titanium (Ti)—ideally
a Ni16Ti6Si7 type composition, forming on the spinodal boundaries or on disloca-
tions in the dferrite. Also, precipitation and preferential growth of carbides and
nitrides at the ferrite/austenite interface occur commonly at approximately 4008C or
above. The differences in aging behavior between the grades mainly attributes to that
CF3 alloys contain low concentrations of carbon (C), Mo and Si; hence, the amounts
of carbide and G-phase are very low, while the CF8(M) grades have higher Si (and

5
Indicates that the grade is corrosion resistant (C) with 17 % to 21 % chromium and 8 % to 12 % nickel (F).
60 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

TABLE 1 Typical composition in weight percent of standard cast stainless steels in nuclear
applications.

Grade C Mn P S Ni Cr Mo

CF3 0.03 0.60 0.003 0.002 9 18 <0.5


CF8 0.057 0.62 0.003 0.002 8.5 20 <0.5
CF8M 0.074 1.21 0.032 0.024 9.6 18.7 2.7

Note: C ¼ carbon; Mn ¼ manganese; P ¼ phosphorus; S ¼ sulfur; Ni ¼ nickel; Cr ¼ chromium;


Mo ¼ molybdenum.

Mo) contents and form more carbides, G-phase, and are more prone to spinodal de-
composition [10]; for example, Mo is seen to increase the rate of decomposition.
The current materials are of grade CF8M; hence, they are expected to undergo
fast spinodal decomposition and are prone to G-phase and carbide precipitation.
The transformation of the microstructure does affect the mechanical properties,
though consensus has not been fully reached on the magnitude of the effects. Because
only ferrite is significantly aged, complex variations in microstructures and composi-
tions are partial causes of these varying results. Studies on cast alloys have revealed
that phase boundary precipitates also play a significant role in thermal embrittlement
[9]. Properties seen to change with aging are FT, impact toughness, hardness, cyclic
hardening, stress corrosion cracking rate, corrosion properties, and so forth.

Experimental Basis
MATERIALS
Test materials are extracted from the NPP Ringhals 2 steam generator loop 2 inlet
(hot, HL) and crossover (outlet, CL) to reactor coolant pump (cold) elbows during a
component replacement project. The cast material is CF8M that complies with
ASTM A351, Standard Specification for Castings, Austenitic, for Pressure-Containing
Parts, with a chemical composition according to Table 2. Only small compositional
differences are seen between the heats, but a slightly higher Cr and copper (Cu)

TABLE 2 Chemical composition and Schaeffler ferrite content of Ringhals 2 cast SG-elbows6 and
connecting weld material in weight percent.

C Si Mn P S Cr Ni Mo Ti Cu Co N Ferrite

HL 0.037 1.03 0.77 0.022 0.008 20.0 10.6 2.09 0.004 0.17 0.040 0.044 20.1
CL 0.039 1.11 0.82 0.020 0.012 19.6 10.5 2.08 0.004 0.08 0.035 0.037 19.8
Weld 0.02 0.8 1.7 20.0 10.2

6
Material certificate from S. A. Usines Emile Henricot, Heat 2009, Charge 4594 and 5854.
BJURMAN ET AL., DOI 10.1520/STP159820160086 61

content is seen for the HL samples. Tabulated ferrite contents are calculated from
the composition using Schaeffler diagrams. Heat treatment of the cast elbows was
originally conducted at 10508C for 24 h, followed by rapid cooling in accordance
with ASTM A351. The castings are large, approximately 40 tons, causing significant
cooling-rate gradients during solidification and heat treatment. The microstructure
consists of regions with both columnar and equiaxed solidification structures.
The dferrite contents were previously measured in these elbows by a ferrite-
scope to vary from 1.5 to 22.5 volume percent [11]. Most contents measured during
this investigation used a larger activated volume, which resulted in average ferrite
contents of 11 6 1 volume percent.
The ferrite structure exhibits large ferrite areas from slow cooling of these large
castings. The microstructure in the chosen regions is fairly equiaxed, with ferrite
spacing of approximately 100 lm with 10 to 20 lm ferrite arm diameter and some
carbide precipitates present in the cross-sections. No large microstructural differ-
ences among castings are seen.
The multipass weld is Böhlers EAS-2 IG (Si), a gas metal arc welding solid wire
with typical composition according to Table 2.

AGING CONDITION
Aging times and temperatures are presented in Table 3, with two exposure periods
at different temperatures due to the reduction of allowable full power rating of the
plant and therefore reduction of temperature during the last 22 kWh of service.
The mechanism of spinodal decomposition growth is diffusion-driven and
aging by decomposition; therefore, it follows the Arrhenius behavior, that is,
t1 Q 1 1
¼ e  R ðT 1  T 2 Þ (1)
t2

where:
Q [J/mole] ¼ the activation energy,
t [s] ¼ time,
T [K] ¼ the absolute temperature,
Subscripts 1 and 2 ¼ the two respective temperatures, and
R [J/(Kmole)] ¼ the gas constant.

TABLE 3 Full power temperature exposure of Ringhals 2 SG-elbows.

Full Power Time 70,000 h 22,000 h Equivalent Aging Time at 325°C

HL 325°C 303°C 74 kWh


CL 291°C 274°C 4 kWh
Temp. difference DT 34°C 29°C
Total Full Power time 92,000 h
62 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

By using the Arrhenius equation and the various times at different respective
temperatures given in Table 3, the aging state can be recalculated to equivalent times
at 3258C. The activation energy for spinodal decomposition in CF8M has previous-
ly been reported to be approximately 240 kJ/mole [12] and has been shown by
atom probe tomography (APT) to be reasonable for the material [13]. Using the ac-
tivation energy in Eq 1, the results are given in Table 3, and a spinodal decomposi-
tion or aging quotient of 18 between the two aging states is calculated. Lower
activation energies have been reported for weld materials [14], hence faster aging,
but only slightly less compared to the scatter in activation energies published.

Testing Approach and Testing Matrix


The purpose of the tests is to assess evolution of the mechanical properties due to
thermal aging of the Ringhals 2 operationally aged steam generator loop 2 hot- and
crossover-legs and to compare these to simulated nonaged material (heat-treated).
Relevant data are needed for assessment of long-term structural integrity, calcula-
tion of acceptable flaw sizes, verification of sufficient FT, and the leak before break
principle of these NPP pressure boundary components. Data are also used as input
for modeling efforts of mechanical properties of the two-phase structures, with the
intention of understanding the full range of properties with various microstructures
and the rate dependency of these effects.
The testing approach is based on the need to verify fracture mechanical data
and address the significant variations in material properties while using a limited
test matrix. Because the cast components are large, regions of low fracture resis-
tance that are much smaller than any dimension in actual components will only af-
fect the components’ fracture resistance marginally.
Prior to the fracture mechanical testing and as a base line, several measuring
techniques have been used to quantify aging parameters that affect fracture me-
chanical properties.
APT was used to measure the nanoscale microstructural evolution of ferrite
from spinodal decomposition and G-phase precipitation [13].
Micro- and nano-indentation were used to measure the hardness evolution
with aging in each phase of the cast materials. For weld specimens, only overall
microhardness was measured.
Variations in microstructure are documented using light optical microscopy
(LOM), scanning electron microscopy (SEM), and ferritescope. Electron back scat-
ter diffraction (EBSD) measurements of as-received and strained tensile specimens
are used to measure strain between phases and strain-localization within the
microstructure.
Testing temperatures RT and 3008C are chosen to cover the component tem-
perature range from cold shutdown and maintenance at RT to operating tempera-
tures. The 300 C temperature, between the HL state of 325 C and CL state of
291 C, is chosen to represent both states for comparability purposes. The FT test
BJURMAN ET AL., DOI 10.1520/STP159820160086 63

TABLE 4 Testing matrix for FT at RT.

State RT 300°C

HL L-S, R-S L-S, R-S


CL L-S, R-S R-S
Reference state R-S R-S
HL_W* R-S R-S
CL_W* R-S R-S

Note: Crack propagation is from the pipe’s water-exposed inner surface toward the outer surface
in the radial-short and longitudinal-short directions. Duplicate specimens were used for each case.
*18 mm maximum width because of limited material size.

matrix is presented in Table 4. Both relevant cracking directions simulating axial


pipe split and transverse pipe failure (L-S and R-S, respectively) crack orientation
directions are tested for both cast aging states. The L-S direction is not tested in
weld material due to weld dimension restrictions. The states are named HL and CL
for cast materials, the reference state is the heat treated casting, and HL_W and
CL_W are the weld material. Charpy-V testing with double and triple specimens is
conducted for both directions in cast material and the L-S direction for weld mate-
rial (see Table 5). A U-notch reference test, Table 6, is conducted for comparison
with the as-cast testing from the original quality control.
The combination of J-integral resistance curve (J-R) testing and Charpy testing
is conducted for several reasons. The J-R curves of each state are needed for the
structural integrity assessment at the plant. The Charpy testing is conducted for
ductile to brittle transition temperature (DBTT) evaluation, upper shelf energy
(USE)/lower shelf energy (LSE) to monitor the aging progress by the DBTT and
USE/LSE shifts. Dynamically instrumented Charpy is used here with the intention
of limiting the need for extensive J-R testing by replacing them with smaller impact
specimens. Dynamic instrumentation data will be used to calculate plane strain FT
J1c and dynamic FT J1d and other crack extension resistance JR-values, based on rea-
sonably small specimens. Another positive aspect of high rate testing is that results
are more decisive when evaluating hardening of the ferrite phase.

TABLE 5 Test matrix for impact testing KCV.

Temperature [°C] 2196 250 RT 100 150 200 250 300 400

CL L-S 2 3 3 3 3 2
CL R-S 2 3 3 3 3 2
HL L-S 2 3 3 3 3 2
HL R-S 2 3 3 3 3 2
CL_W L-S 2 3 3 3 3 2
HL_W L-S 2 3 3 3 3 2
64 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

TABLE 6 Test matrix for impact testing KCU.

Temperature [°C] RT

CL L-S 1
CL R-S 1
HL L-S 1
HL R-S 1

Small specimens introduce larger scatter in test results, but in this case, where
large heterogeneity of the microstructure generates the scatter, they also allow for
the use of statistically based analysis of ferrite and austenite behavior.
The basis for the test plan and choice of testing locations followed a specific
procedure:
• Microstructural evaluation to choose specimens from microstructurally repre-
sentative locations with regions of fairly equiaxed ferrite having ferrite con-
tents of 10 % to 15 %
• Specimen fabrication
• Microstructural evaluation by LOM of selected specimens
• Verifying ferrite content of selected test specimens by ferritescope
• Precracking at RT
• Testing
• Evaluation of results and post-fractography

The Charpy testing is designed as a complementary and comparative test of pre-


vious results [11] from other parts of the same components. Testing temperatures
and the number of specimens at each temperature are determined prior to testing
and are based on these previous results. Two types of specimens are tested, V-notch
(KCV) and U-notch (KCU). The latter is also tested during manufacturing of the
components. In-depth analysis of impact tests is made from dynamic instrumenta-
tion results to increase the understanding of crack propagation behavior. The two-
phase structure has shown indications of strain rate sensitivity; therefore, this is part
of the analysis. The force-displacement results can separate possible cracking phases,
elastic initiation phase, plastic growth, brittle propagation, and plastic final fracture.
Force, displacement, and fracture energies according to ASTM E2298, Standard Test
Method for Instrumented Impact Testing of Metallic Materials [15], were measured.

MECHANICAL TESTING
Fracture mechanical J-R testing is conducted on an Instron 8500 servohydraulic
testing machine according to ASTM E1820, Standard Test Method for Measurement
of Fracture Toughness [16]. Tensile testing is conducted according to the SS-EN
6892-1:2009 standard for testing at RT and SS-EN 10002-5 at elevated temperature.
A clip-gage is used in all cases—except for high-temperature tensile testing, where a
linear variable differential transformer is used. A water-cooled furnace is used for
the elevated temperature tests.
BJURMAN ET AL., DOI 10.1520/STP159820160086 65

Impact testing is based on the SS EN-10045-1 [17] standard and is conducted


on Losenhausen pendulum impact testing equipment. Specimens used are a stan-
dard Charpy U- and V-notch type with the dimensions 10 by 10 by 55 mm.
Dynamically instrumented Charpy testing is also conducted on cast material in HL
and CL conditions using Charpy V-notch specimens. An INSTRON CEAST 9350
drop tower equipped with an instrumented tup (striker) is used for these tests. The
load range of the piezoelectric force sensor used in the tup is 0 to 22 kN.
FT-testing specimens are 18- to 25-mm-wide compact tension-type-specimens
with a chevron starter notch. Precracking is conducted at room temperature with
an initial load of Po ¼ 6 kN and a range of stress intensity DK ¼ 15 MPaHm and
R ¼ 0.1 at a frequency of f ¼ 18 Hz. The heterogeneity of these materials makes it
difficult to fatigue precrack specimens with crack fronts fulfilling all requirements
in the ASTM E1820 standard. To achieve cracks with reasonably straight and
in-plane crack fronts, a long notch 23 mm in length and 1.5 mm fatigue precracks
are used, followed by 10 % side grooves on each side to minimize the remaining
thumbnail crack-shape and to guide the crack propagation during testing. The 10 %
side grooving is chosen because this has been frequently used in similar previously
conducted tests.
A micro Vickers hardness indenter is used on polished metallography speci-
mens to measure the hardness of each phase at HV 50 p when measuring ferrite
and austenite. For bulk hardness measurements, HV 500 p is used.

MICROSCOPY
The microscopes used are a tungsten filament SEM, JEOL 6300 with a Thermo-
Fisher NS-7. Energy dispersive spectroscopy (EDS) is used for elemental analysis of
ferrite and austenite phases. EDS analyses are carried out at 20 keV and averaged
over several grains. A Zeiss Auriga Cross Beam field emission gun (FEG)-SEM
with a Hikari EBSD camera with software supplied by EDAX is also used. Strains
are evaluated using EBSD. Twinning, overall dislocation structure, and local phase-
structure and proportions are also measured using EBSD.
LOM examinations are conducted in a Reichert Polyvart MET (presently Leica)
microscope, equipped with a Leica DFC digital camera.

Results
MICROSTRUCTURE
The microstructures of the cast materials examined by LOM are seen in Fig. 1a–c
and those of the weld material in Fig. 1d–f. The castings exhibit a coarse ferrite struc-
ture originating from the slow cooling of these large castings. The variational span
of microstructures is illustrated in Fig. 1. In the regions chosen for testing, the
microstructure is fairly equiaxed. For cast materials, the ferrite spacing is approxi-
mately 100 lm with a 10- to 20-lm ferrite arm diameter as well as some carbide
precipitates present in the cross-sections. The austenite grain size is approximately
66 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 1 LOM of cast microstructure HL (a) and some details from CL (b,c). Weld
microstructure CL (d) and some details from HL (e,f).

(a) (d)

(b) (c) (e) (f)

2 mm as seen from the orientation of ferrite regions in Fig. 1a, verified by EBSD.
Weld ferrite spacing is typically 10 lm and approximately 1-lm dendrite arm
diameter. No large microstructural differences are seen among the castings.
For cast materials, dferrite contents in these elbows were measured previously
by ferritescope and vary from 1.5 to 22.5 volume percent [11] with an average of
10 volume percent. The dferrite contents measured during this investigation in
the tested specimens result in average ferrite contents of 11 6 1 volume percent.
The weld bead size is measured to 3 mm and with a ferrite content of 10 6 0.5 volume
percent using a ferritescope.
The microscale deformation behavior of the two-phase structure is shown in
Fig. 2, where an inverse pole figure plot of fractured ferrite (blue) in an austenite
matrix shows the localization of strains in the thinner ferrite regions and around
the fracture surfaces. In the austenite strain, localization is mainly seen around the
crack tips of ferrite and close to the phase boundary.

HARDNESS
Micro hardness is measured on both cast and weld material. Ferrite and austenite
hardness results are measured at 50 gf and are only possible for cast material be-
cause the dimensions of ferrite in the weld material are too small. The results pre-
sented in Table 7 show significant scatter due to the difficulty of measuring each
BJURMAN ET AL., DOI 10.1520/STP159820160086 67

FIG. 2 EBSD-plot of HL casting strained with 103s1 strain rate, and at 3008C, EBSD-
inverse pole figure (IPF) map of strained ferrite (blue) in an austenite matrix.

phase exclusively. A further issue is that the thickness of the ferrite tested is not
known. Average hardness is measured at 500 gF on both cast and weld material. In
the cast material, indentations are the same size as those of the microstructure, and
the average of randomly positioned indents is used as overall hardness. A signifi-
cantly higher hardness is seen for the weld material.

FRACTURE TOUGHNESS TESTING


The results of FT testing at RT is shown in Table 8. Cast CT specimens are 25 mm
wide and weld specimens are 18 mm wide. All specimens have 10 % side grooves
machined prior to testing. The CL specimens and one of the HL specimens blunts
in J-R testing. Blunting of the HL specimen is most likely due to the precrack locally
extending out of the intended fracture plane. The HL specimens show well-behaving
stable crack growth, as seen in the J-R curve of HL in the L-S orientation with the

TABLE 7 Vickers hardness measurement of each phase and macro hardness for cast and welded
material.

HV Overall Ferrite Austenite

CL Cast 190 291 183


HL Cast 207 452 189
CL Weld 260
HL Weld 285
68 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

TABLE 8 Testing results of FT for each tested specimen given in JQ (kN/m); tests are conducted at
RT with B indicating blunting of the crack front.

State RT

L-S R-S

[kN/m] Low High Low High

HL 326 B 347 450


CL B* B B* B
Ref. state B B B B
HL_W 166 196 - -
CL_W 287 314 - -

Note: *FT-test at increased strain rate.

qualification J-resistance JQ ¼ 326, Fig. 3a. JQ is used instead of J1C when not all of
the blunting crack of the CL specimen in R-S direction is seen in Fig. 3b. All RT
weld specimens show stable crack growth and are considered representative for FT
of weld, even though the crack front straightness criteria are not met. J-R curves for
weld are seen in Fig. 4.
In order to assess the strain rate sensitivity and achieve JQ for the CL-state,
specimens are loaded at an increased rate of 20 s and 10 s to maximum load Pmax,
compared to 100 s, equal to approximately K ¼ 2.5MPam1/2/s for the other speci-
mens. No significant difference in FT behavior compared with the lower strain rate
is noted.
Examining the fracture surfaces in SEM, fractography of cast material shows
that the CL state, Fig. 5, exhibits only ductile dimple fracture throughout, while the
HL state, Fig. 6, also exhibits mainly ductile dimple type fracture but with some per-
centage of cleavage fracture. There are also some secondary cracks out of plane as
seen in Fig. 6b. These cracks appear to propagate in the ferrite phase or the ferrite-
austenite boundaries.
The weld material fractograph of HL weld, Figs. 7a and 7b, predominantly
shows ductile failure with a dimple size much smaller than for cast material. No
splits are present in the cross-sections, seen in Fig. 7.

IMPACT TESTING
Dynamically instrumented impact tests are conducted at RT with 3.8 m/s
impact speed. Results are presented in Table 9. The HL specimens are both in
the transition region with mixed mode behavior deduced from the plot of Fig. 8.
The CL state is well into the ductile region as seen in the plots in Figs. 9 and 10.
Note that only CL2 is conducted at sufficient impact energy to produce valid
results.
Results of the ordinary impact testing of both weld and cast material using both
U- and V-notch Charpy specimens at RT are shown in Fig. 11.
BJURMAN ET AL., DOI 10.1520/STP159820160086 69

FIG. 3 J-R curves for RT-tests of cast materials (a) HL L-S and (b) CL R-S. Only the HL
specimens resulted in crack propagation. Lines are defined as: (1) 0-line with
slope two times the flow stress, (2) 0.15-mm lower exclusion line, (3) 0.5-mm
intersect line, and (4) 1.5-mm upper exclusion line.

800
JQ = 326.769 kN/m

700

600

500
J / kN/m

400

300

200

100

0
-0.5 0 0.5 1 1.5 2 2.5 3

a / mm

(a)
1,200
JQ = kN/m

1,000

800
J / kN/m

600

400

200

0
0 0.5 1 1.5 2 2.5 3

a / mm

(b)
70 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 4 J-R curves for RT-tests of weld materials.

General Discussion
From the equivalent aging times calculated in the “Aging Condition” section,
the equivalent aging time of HL is seen to be 18 times longer than that of CL.

FIG. 5 Cast CL fracture surfaces from FT test of in LS direction, centrally close to


prefatigue to J-R transition line: (a) 30  magnification and (b) 1000 
magnification.

(a) (b)
BJURMAN ET AL., DOI 10.1520/STP159820160086 71

FIG. 6 Cast HL fracture surfaces from RT FT test, centrally close to prefatigue to J-R
transition line: (a,b) at 100  and (c,d) at 1000  magnification.

(a) (b)

(c) (d)

The hardness results in Table 7 show a significantly larger increase in hardness


achieved in the HL d-ferrite compared to CL. Although aging significantly changes
the d-ferrite mechanical properties, change in the austenitic phase remains very
limited. The mechanical response to aging will therefore be highly sensitive to the
local ferrite content, morphology, and size of ferrite structures. Mechanical proper-
ties of ferrite and austenite phases in the nonaged state are similar. Hence, the local
stress and strain fields during deformation are of similar size in the two phases.
With aging, the differences between phases increase and the sensitivity to phase
structure variations increases.
The ferrite contents of weld and castings, measured with a ferritescope and cal-
culated from the area fraction in light optical micrographs, are similar. Aging is
quantified by APT for the cast components but not for weld material. On the other
hand, the components have been exposed to the same time and temperatures and
can therefore be seen as equally aged. In this comparable aging state, welds are
much less tough, with JQ reduced by at least a factor of three compared to the cast
material. Several significant differences affect this result. From the standpoint of mi-
crostructural differences, the sizes of ferrite-austenite regions are a decade larger
72 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 7 Weld fracture surfaces from RT FT-test, (a,b) HL and (c,d) CL.

(a) (b)

(c) (d)

and less elongated for the cast material than for the weld material, which predomi-
nantly is caused by the high cooling rate of the weld. This also causes higher inter-
nal stresses affecting the stress state, hence, FT and tensile properties.
This coarse structure causes the cast material, and to a limited extent weld ma-
terial, to suffer from highly irregular precrack fronts. The crack surface can deviate
locally from the intended plane by several millimetres, even though corrective mea-
sures are taken. Several specimens are therfore discarded during precracking. When
the cause of irregular precrack fronts is from microstructural heterogenity and

TABLE 9 Results of instrumented Charpy test for RT specimens.

Peak Force Peak Energy Total Energy Peak Displ. Total Displ.
[J] [J] [J] [mm] [mm]

HL1 14.688 38.1 84 3.41 11.2


HL2 15.431 36.7 59 3.18 6.98
CL1 14.988 103.6 182 8.42 14.0
CL2 14.595 115.2 238 9.51 22.7
BJURMAN ET AL., DOI 10.1520/STP159820160086 73

FIG. 8 Instrumented time versus force impact graph for HL specimens HL1 (thin curve,
left specimen in picture) and HL2 (thick curve, right specimen in picture).

anisotropy, discarding specimens can bias the test result and great care is therefore
taken to minimize the number of discarded specimens.

FRACTURE MODE DISCUSSION


The two phases are inherently different with respect to fracture mode. The austenite
is prone to ductile failure, with void formation mainly on ferrite and other precipi-
tate positions. This is a process initiating based on both stress and strain. By using,

FIG. 9 Instrumented time versus force impact graph for CL2 and specimen after
testing.
74 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 10 Instrumented time versus force impact graph for CL1 and specimen after
testing.

for example, Argon’s continuum mechanics-based void nucleation model [18], crit-
ical stress rc when voids form due to de-bonding of cylindrical particles is approxi-
mately the sum of effective (von Mises) and mean stress.
Void nucleation is often seen to occur more often with large particles (e.g., with
ferrite in castings). This is possibly due to the presence of defects.

FIG. 11 Results of impact testing of cast and weld at RT using U- and V-notch
specimens.

250

200
Impact energy [J]

KCV1
150
KCV2
KCV3
100
KCU

50

0
HL L-S HL R-S CL L-S CL R-S Weld HL RS Weld CL RS
BJURMAN ET AL., DOI 10.1520/STP159820160086 75

The ferrite, on the other hand, cracks by cleavage, which is mainly a stress-driven
process with scatter based on critical defect size. Any stress-increasing process such
as triaxial stress state, lower temperature, high strain rate, strain aging, and so on,
therefore increases the probability of fracture in the ferrite.
The ferrite is sufficiently large in the welds that a cracked ferrite arm causes sig-
nificant stresses, strain, and blunting of austenite but is small enough to withstand
significant stress from a tensile test. Ferrite areas in the weld are also more elongat-
ed compared to castings, which increases sensitivity to bending forces from local-
ized strain.
For a spherical particle in ferrite, producing a penny crack, the failure stress is
proportional to Co1/2 where Co is the particle or defect diameter. These effects
contribute to the total tendency for fracture. By assuming the power law for stress-
strain in Eq 2 [19,20],
 n
e r r
¼ þa (2)
e0 r0 r0

With stress r, reference stress r0 usually being yield stress, and strains e and
reference strain e0 ¼ r0/E; a is a constant without dimension, and n is the strain
hardening exponent.
The stress and strains ahead of the tip can be written as Eq 3
 nþ1
1  nþ1
n
J J
rij ¼ k1 ; eij ¼ k2 (3)
r r

Hence, with increasing n, stresses are decreasing while strains are increasing in
front of a crack tip, where cleavage and void nucleation occur. Higher stresses will
increase the likelihood of cleavage fracture of the ferrite.
Hardening exponents are calculated from tensile results previously pre-
sented [21]. Higher hardening exponents are seen for weld specimens (n > 2)
than for cast specimens (n ¼ 1.7) in agreement with the FT results, Fig. 12. As
expected, a temperature effect is seen with a lower average strain hardening ex-
ponent n (e.g., for cast material n ¼ 1.1 at 3008C compared to RT where
n ¼ 1.7). This is also consistent with expected FT previously measured [11]. On
the other hand, only very limited and not statistically significant differences in
hardening are seen between the aging states. Increasing exponents are linked to
the reduction of FT results.
It is also seen from the fractographs that weld specimens exhibit much smaller
dimples than cast specimens. This is in line with the lower FT of weld material also
seen in other stainless steels [22]. Higher density of defects in the ferrite, and possi-
bly a higher local stress triaxiality, contribute to this.
From the plot of conducted FT tests in Fig. 13, it is seen that JQ are high. A signifi-
cant difference in FT between weld and cast material of the same aging and ferrite
76 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 12 Correlation between the effect of hardening coefficient and FT at RT.

contents is also seen. An effect of aging is also apparent for both materials. Fracture
toughnesses appears higher in the current test when compared to Jansson [11].
Charpy testing results from the impact tests show higher impact energies com-
pared to previous results, Figs. 14 and 15, but they are in reasonable agreement.
From the dynamically instrumented results, it is seen that both CL and HL states

FIG. 13 Fracture toughness results of cast and weld material and comparison to results
from Jansson [11]. Arrows indicate that the specimens are blunted and that JQ
are high.
BJURMAN ET AL., DOI 10.1520/STP159820160086 77

FIG. 14 Results of ordinary and dynamically instrumented impact testing of the


CL-casting as well as average results from Jansson [11].

250

200
CL L-S ref. 11
Impact Energy [J]

150 CL R-S ref. 11

CL Instrumented
100
Heat Treated

50 CL L-S

CL R-S
0
-250 -150 -50 50 150 250
Temperature [°C]

are well into the DBT-region, also verified by dynamic data in Figs. 8–10. These
figures also show that the material behaves well enough in impact testing for evalua-
tion of dynamic J. The dynamic impact testing could therefore be used to reduce

FIG. 15 Results of ordinary and dynamically instrumented impact testing of the


HL-casting as well as average results from Jansson [11].

120

100
HL L-S ref 11
80
Impact Energy [J]

HL R-S ref 11
60
HL Instrumented
40

HL L-S
20

HL R-S
0
-250 -150 -50 50 150 250
Temperature [°C]
78 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

the material needs of testing compared to the J-R approach, while still having a frac-
ture mechanically relevant measure of aging evolution.
It is seen in Fig. 11 that impact results at RT weld show lower impact toughness
compared to cast material, which is consistent with the J-R results, and that
U-notch specimens are lower in toughness compared to V-notch specimens. This is
due to the increased activated volume compared to V-notch specimens. The use of
U-notch specimens also allows for a more natural fracture route, and these results
can therefore be regarded as more relevant to component failure.

Summary
The testing methodology being developed is based on the need to obtain relevant
fracture mechanical results of aged cast and welded stainless ferritic-austenitic
structures with a large microstructural heterogeneity, using limited material and
resources.
• The approach uses microstructural evaluation as the basis of test specimen se-
lection with representative microstructures to reduce scatter from large com-
ponent heterogeneity.
• These very large castings generate very coarse microstructure and, though spi-
nodal decomposition is similar, fracture mechanically age very different com-
pared to the fast solidified welds.
• A correlation between hardening coefficients in tensile tests and the FT is seen
between weld and cast materials and with temperature. The correlation is also
corroborated by the dimple size of fracture surfaces. On the other hand, the
correlation with aging state is limited.
• Charpy U-notch specimens show consistently lower impact toughness ener-
gies compared to V-notch specimens due to the larger volume of activated ma-
terial during testing. The use of U-notch specimens is also more relevant for a
large component typical in a power plant.
• Impact tests require smaller specimen sizes compared to J-R tests and the ma-
terial results in dynamically instrumented Charpy show that these can be used
to calculate J-values. The instrumented tests therefore show potential to be
used as quick mechanically sound tests of thermal aging using small material
volumes.

ACKNOWLEDGMENTS
The authors wish to express their sincere gratitude to Irene Linares at KTH for con-
ducting much of the fracture mechanical testing, the PhD project reference group con-
sisting of Peter Ekström (Swedish Radiation Safety Authority), Massimo Cocco
(Forsmarks Kraftgrupp), Mattias Coudret (Oskarshamns Kraftgrupp), Björn For-
ssgren (Ringhals), and Anders Jenssen (Studsvik) for fruitful discussions. Thanks also
to Alessandro Tomaiuolo at Instron for conducting the instrumented impact testing.
The work was funded by Ringhals AB, Swedish Radiation Safety Authority, Fors-
marks Kraftgrupp AB, Oskarshamns Kraft AB, and Studsvik Nuclear AB.
BJURMAN ET AL., DOI 10.1520/STP159820160086 79

References

[1] Sahu, J. K., Krupp, U., Ghosh, R. N., and Christ, H. J., “Effect of 475°C Embrittlement on
the Mechanical Properties of Duplex Stainless Steel,” Mater. Sci. Eng. A, Vol. 508, Nos.
1–2, 2009, pp. 1–14.

[2] Fisher, R. M., Dulis, E. J, and Carrol, K. G., “Identification of the Precipitate Accompanying
885°F Embrittlement in Chromium Steels,” Trans. AIME, Vol. 197, 1953, pp. 690–695.

[3] Chandra, D. and Schwartz, L. H, “Mössbauer Effect Study of the 475°C Decomposition of
Fe-Cr,” Metall. Trans., Vol. 2, 1971, pp. 511–519.

[4] Calonne, V., Berdin, C., Saint-Germain, B., and Jayet-Gendrot, S., “Damage and Dynamic
Strain Aging in a Thermal Aged Cast Duplex Stainless Steel,” J. Nucl. Mater., Vol. 327,
Nos. 2–3, 2004, pp. 202–210.

[5] Notten, G., “Application of Duplex Stainless Steel in the Chemical Process Industry,” Pro-
ceedings of 5th World Conference, Duplex Stainless Steel, Maastricht, The Netherlands,
October 21–23, 1997, KCI Publishing, Toronto, 1997, pp. 9–16.

[6] El-Batahgy, A. and Zaghloul, B., “Fatigue Failure of an Offshore Condensate Recycle
Line in a Natural Gas Production Field,” Mater. Charact., Vol. 54, No. 3, 2005,
pp. 246–253.

[7] Reidrich, G. and Loib, F., “Embrittlement of High Chromium Steels within Temperature
Range of 570–1100°F,” Arch. Eisenhüttenwesen, Vol. 15, 1941, pp. 175–182.

[8] Blackburn, J. M. and Nutting, J. J., “Metallography of an Iron-21 % Chromium Alloy Sub-
ject to 475°C Embrittlement,” J. Iron Steel Inst., Vol. 202, 1964, pp. 610–613.

[9] Cortie, M. B. and Pollak, H., “Embrittlement and Aging at 475°C in an Experimental Fer-
ritic Stainless Steel Containing 38 wt.% Chromium,” Mater. Sci. Eng. A, Vol. 199, No. 2,
1995, pp. 153–163.

[10] Brown, J. E. and Smith, G. D. W., “Atom Probe Studies of Spinodal Processes in Duplex
Stainless Steels and Single- and Dual-Phase Fe-Cr-Ni Alloys,” Surf. Sci., Vol. 246,
Nos. 1–3, 1991, pp. 285–291.

[11] Jansson, C., “Degradation of Cast Stainless Steel Elbows after 15 Years in Service in a
PWR,” presented at Fontevraud II International Symposium, Avignon, France, Septem-
ber 10–14, 1991.

[12] Pareige, C., Novy, S., Saillet, S., and Pareige, P., “Study of Phase Transformation and
Mechanical Properties Evolution of Duplex Stainless Steels after Long Term Thermal
Ageing (>20 Years),” J. Nucl. Mater., Vol. 411, Nos. 1–3, 2011, pp. 90–96.

[13] Bjurman, M., Thuvander, M., Liu, F., and Efsing, P., “Phase Separation Study of In-Service
Thermally Aged Cast Stainless Steel—Atom Probe Tomography,” presented at the 17th
International Conference on Environmental Degradation of Materials in Nuclear Power
Systems—Water Reactors, Ottawa, Ontario, August 9–12, 2015.

[14] Chandra, K., Kain, V., Bhutani, V., Raja, V. S., Tewari, R., Dey, G. K., and Chakravartty, J. K.,
“Low Temperature Thermal Aging of Austenitic Stainless Steel Welds: Kinetics and
Effects on Mechanical Properties,” Mater. Sci. Eng. A, Vol. 534, 2012, pp. 163–175.
80 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

[15] ASTM E2298-15, Standard Test Method for Instrumented Impact Testing of Metallic Mate-
rials, ASTM International, West Conshohocken, PA, 2015, www.astm.org, http://
dx.doi.org/10.1520/E2298-15

[16] ASTM E1820-13, Standard Test Method for Measurement of Fracture Toughness, ASTM
International, West Conshohocken, PA, 2013, www.astm.org, http://dx.doi.org/10.1520/
E1820

[17] SS-EN 10045-1, Metallic Materials—Charpy Impact Test—Part 1: Test Method, Swedish
Standards Institute, Stockholm, 1990.

[18] Argon, A. S., Im, J., and Safoglu, R., “Cavity Formation from Inclusions in Ductile
Fracture,” Metall. Trans. A, Vol. 6, 1975, pp. 825–837.

[19] Rice, J. R. and Rosengren, G. F., “Plane Strain Deformation near a Crack Tip in a Power
Law Hardening Material,” J. Mechan. Phys. Solids, Vol. 16, No. 1, 1968, pp. 1–12.

[20] Hutchinson, J. W., “Singular Behaviour at the End of a Tensile Crack Tip in a Hardening
Material,” J. Mechan. Phys. Solids, Vol. 16, No. 1, 1968, pp. 13–31.

[21] Bjurman, M. and Efsing, P., “Localized Deformation Behaviour of Thermally Aged Stain-
less Steel Castings,” presented at Fontevraud II International Symposium, Avignon,
France, September 15–18, 2014.

[22] Das, A. and Tarafder, S., “Geometry of Dimples and Its Correlation with Mechanical Prop-
erties in Austenitic Stainless Steel,” Scr. Mater., Vol. 59, No. 9, 2008, pp. 1014–1017.
FATIGUE AND FRACTURE TEST PLANNING, TEST DATA ACQUISITIONS AND ANALYSIS 81

STP 1598, 2017 / available online at www.astm.org / doi: 10.1520/STP159820160053

Noushin Torabian,1 Véronique Favier,2 Saeed Ziaei-Rad,3


Justin Dirrenberger,2 Frédéric Adamski,2 and Nicolas Ranc2

Calorimetric Studies and


Self-Heating Measurements
for a Dual-Phase Steel Under
Ultrasonic Fatigue Loading
Citation
Torabian, N., Favier, V., Ziaei-Rad, S., Dirrenberger, J., Adamski, F., and Ranc, N., “Calorimetric
Studies and Self-Heating Measurements for a Dual-Phase Steel Under Ultrasonic Fatigue
Loading,” Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis, ASTM
STP1598, Z. Wei, K. Nikbin, P. C. McKeighan, and D. G. Harlow, Eds., ASTM International, West
Conshohocken, PA, 2017, pp. 81–93, http://dx.doi.org/10.1520/STP1598201600534

ABSTRACT
The objective of the present research is to study the self-heating behavior of a
dual-phase (DP) steel under ultrasonic fatigue loading and to investigate the
effect of frequency on intrinsic heat dissipation of the material. The steel studied
in this work is DP600 commercial DP steel. Fatigue tests were conducted using
an ultrasonic fatigue machine at a testing frequency of 20 kHz with flat
specimens. An infrared camera was used to measure the mean temperature
evolution during the tests. A specific form of heat diffusion equation was
adopted in this work to calculate the heat dissipation per cycle from temperature
measurements. The variation of this dissipation versus stress amplitude in cyclic
loading was also studied.

Manuscript received March 1, 2016; accepted for publication October 17, 2016.
1
Laboratoire PIMM, Arts et Métiers-Paris Tech, CNRS, Cnam, 151 Boulevard de l’Hôpital, Paris 75013, France
http://orcid.org/0000-0002-4410-8203
2
Laboratoire PIMM, Arts et Métiers-Paris Tech, CNRS, Cnam, 151 Boulevard de l’Hôpital, Paris 75013, France
3
Isfahan University of Technology, Dept. of Mechanical Engineering, Daneshgah e Sanati Highway, Isfahan
84156-83111, Iran
4
ASTM Symposium on Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis on May 4–5,
2016 in Grand Hyatt, San Antonio, TX.

Copyright V
C 2017 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.
82 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

Keywords
ultrasonic fatigue, infrared thermography, dual-phase steel

Introduction
Since the first relevant conference in Paris in 1998 [1], the fatigue behavior of metallic
materials in the very high cycle fatigue (VHCF) regime (Nf  107) has been investigated
by various research groups. Ultrasonic fatigue testing [2,3] is an effective tool to carry out
VHCF tests. Due to the extremely high loading frequency of 20 kHz, an ultrasonic fatigue
machine considerably reduces the testing time and makes it possible to investigate the
VHCF properties of different high-strength steels in a reasonable time. For instance, it
takes 14 h to reach 109 cycles with ultrasonic loading, while using a conventional fa-
tigue machine with a working frequency of 100 Hz, a time period of four months is re-
quired to go up to 109 cycles. With regard to the application of the ultrasonic testing
frequency, most researchers believe that frequency effect itself is small in most cases at
low stress amplitudes and does not change the essence of fatigue [4–6]. However, the
frequency effect is not clear for all metallic materials such as ferritic-martensitic dual-
phase (DP) steels. Moreover, the thermal effect induced by loading with ultrasonic fre-
quency is still questionable. The aim of the present work is to conduct a thermography
study on a DP steel under ultrasonic fatigue loading at low stress amplitudes.
DP steels are a group of advanced steels that contain hard martensite islands
dispersed in a soft ferrite matrix. Due to its characteristics of high strength, good
ductility, and high initial work hardening rates, DP steel has a broad application in
the automotive industry.
In several research works, the dissipated energy was deduced from self-heating meas-
urements of DP steels during low-frequency fatigue loadings [7–11]. With respect to ul-
trasonic fatigue testing (20 kHz), Blanche et al. [12] as well as Ranc et al. [13] developed
methods to identify dissipative fields from infrared (IR) thermography measurements.
The method developed by Blanche et al. [14] was also applied to compare the dissipative
response during cycling of three single-phase metals: copper, alpha-brass, and alpha-iron.
The present paper studies the thermal response of DP600 commercial steel
under ultrasonic fatigue loading. Successive fatigue tests with increasing stress
amplitude are carried out by means of an ultrasonic fatigue machine with the work-
ing frequency of 20 kHz. IR thermography is used to record the mean temperature
on the surface of the specimen. Self-heating diagrams are developed for this material
based on mean temperature. The mean dissipated energy per cycle is estimated as a
function of the stress amplitude.

Material and Experimental Procedure


MATERIAL CHARACTERIZATION
The material used in this study is DP600 steel that was received as sheets 3 mm thick.
This commercial ferritic-martensitic DP steel contains 15 wt % martensite. Fig. 1 shows
TORABIAN ET AL., DOI 10.1520/STP159820160053 83

FIG. 1 SEM micrograph of the DP600 sample. (Bright grains are martensite and dark
grains are ferrite.)

the microstructure of the material obtained from scanning electron microscope (SEM)
observations. Table 1 presents the chemical composition of the material. The mechanical
properties of DP600 in the transverse direction are presented in Table 2.

ULTRASONIC FATIGUE LOADINGS


Fatigue tests are conducted with flat specimens using an ultrasonic fatigue machine
at a testing frequency of 20 kHz and by applying a sinusoidal displacement wave.
The specimen dimensions are calculated so that the free resonant frequency of the
specimen in the first longitudinal mode is 20 kHz. The specimens are machined
in the transverse direction. All specimens are mechanically and then electrolytical-
ly polished to remove all hardened layers on the specimen surface and, conse-
quently, to release the residual stresses. Fig. 2 illustrates the geometry of the
fatigue specimens.
In all of the tests, the loading ratio is R ¼ 1. During the tests, an IR camera
with a spatial resolution of 0.024 mm per pixel is used to monitor the temperature
field on the specimen surface. The acquisition frequency of the camera is 10 Hz.
The specimen surface is painted in matte black to have a uniform surface emissivity
close to 1. From the temperature measurements, the intrinsic dissipation is

TABLE 1 DP600 chemical composition [15].

Alloying Element Mn P Si Cr C S Nb

% weight 0.93 0.04 0.213 0.727 0.0748 0.001 0.0425


84 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

TABLE 2 Mechanical properties of DP600 steel [15].

Young Modulus, GPa Yield Strength, MPa Ultimate Tensile Strength, MPa Elongation, %

210 420 610 20

determined using a heat diffusion model as explained in the following section. Fig. 3
shows the ultrasonic equipment and the camera.
Successive steps of fatigue tests are conducted by increasing the stress ampli-
tude from 57 MPa to 241 MPa, as shown schematically in Fig. 4. The tests are limit-
ed to low stress amplitudes (i.e., the stress values below the conventional fatigue
limit, which equals 250 MPa) [15]. At each stress amplitude, the fatigue test is car-
ried out up to 107 cycles, and the mean temperature evolutions are registered during
the tests. At the end of each step, the testing machine is stopped and temperature
measurements are continued for 2 min to record the cooling of the specimen after
unloading. Between the two steps that follow, there is a time gap of around 10 min
to restart at equilibrium. In all cases, the temperature is measured at the center of
the gage part of the specimen.

DETERMINATION OF THE DISSIPATED ENERGY


In this study, the specific form of heat diffusion equation proposed by Boulanger
et al. [7] is adopted to calculate the intrinsic dissipation from temperature meas-
urements. Assuming there is no coupling between microstructure and temperature,
and neglecting the convective terms, the heat diffusion equation (Eq 1) is written:
 
qC T_  div k grad ðT Þ ¼ sthe þ d1 þ r (1)

where:
q ¼ the mass density,
C ¼ the specific heat,

FIG. 2 Fatigue test specimen geometry (all dimensions are in mm).


TORABIAN ET AL., DOI 10.1520/STP159820160053 85

FIG. 3 Ultrasonic fatigue testing equipment.

Ultrasonic Horn

Infrared Camera

Flat Specimen

T_ ¼ the time derivative of temperature,


k ¼ the conduction coefficient,
sthe ¼ the thermoelastic source,
d1 ¼ the intrinsic heat dissipation, and
r ¼ the external volume heat supply.
For the DP600 steel, q ¼ 7800 kg=m3 and C ¼ 460 J=kg= C.

FIG. 4 Successive fatigue loadings (each block consists of 107 cycles).


86 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

It is assumed that the volume heat source is time independent so it is expressed


in Eq 2 as:
r ¼ kT0 (2)

in which T0 is the initial equilibrium temperature. By assuming the material parame-


ters q, C, and k to be constant and by introducing the temperature increase
h ¼ T  T0 , the heat diffusion equation (Eq 3) can be rewritten as:

qC h_  kDh ¼ sthe þ d1 (3)

where D is the Laplace operator.


As suggested by Boulanger et al. [7], for symmetric boundary conditions and
initial conditions corresponding to an uniform temperature field, it can be assumed
that heat losses are linear with respect to the temperature variation. Therefore, it
can be assumed that kDh ¼ qC hs, where s is a time constant describing the ther-
mal exchanges between the specimen and its environment. The heat diffusion equa-
tion is therefore rewritten as:
h
qCh_ þ qC ¼ sthe þ d1 : (4)
s

The thermal boundary conditions are not symmetrical in the present case be-
cause one end of the specimen is fixed to the horn and the other end is free, as dis-
cussed in Blanche et al. [12]. However, the thermographic images show that the
specimen temperature field is symmetrical, as will be presented in the following sec-
tion. That is why Eq 4 is considered here.
The time constant s can be determined at the end of the test, just after the
unloading when there is no applied stress; at that moment, the thermoelastic and
intrinsic dissipation heat sources are zero, while the temperature increase is not null
(h 6¼ 0). Therefore, the heat equation is reduced to:
h
qCh_ þ qC ¼ 0: (5)
s

Solving Eq 5, the theoretical relation for temperature increase is obtained as:


t 
hðt Þ ¼ hf exp (6)
s

where hf is the temperature increase measured when the loading stops. Thus, s is
estimated by fitting the experimental data with the theoretical evolution of h.
Moreover because the loading frequency is high, due to the temporal inertia, it
is not possible to measure the instantaneous value of h but only its average value
per cycle determined over numerous loading cycles, therefore, by denoting:
ð 1
tþnT ð1
tþT
1 f
u~ ¼ udt ¼ udt (7)
nT1 n
t t
TORABIAN ET AL., DOI 10.1520/STP159820160053 87

where T1 is the period of the loading, f is the loading frequency, and n a number of
cycles (Eq 7). The heat diffusion equation after the integration explained earlier is re-
written in Eq 8 as:
h~
qC h~_ þqC ¼ ~sthe þ d~1 : (8)
s

As stated by Boulanger et al. [7], the sum of the thermoelastic power over one
loading cycle is null ( ~sthe ¼ 0); thus, the final form of the heat diffusion equation is
obtained in Eq 9:

_ h~
qC h~ þqC ¼ d~1 : (9)
s
_
Therefore, in the stabilized regime, when h~ ¼ 0, the average dissipation per
cycle is easily determined from Eq 10 as:

h~
qC ¼ d~1 : (10)
s

The mean dissipated energy per cycle can be obtained as d~1 =f , where f is the
loading frequency.

Results
Fig. 5 shows the evolution of the temperature for some of the loading steps. This fig-
ure shows that, for each loading step, the temperature increases suddenly in the

FIG. 5 Evolution of the mean temperature during several series of ultrasonic cyclic
loadings.
88 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

early stages of loading (after around 106 cycles) and then gradually reaches a steady
state. In fact, the stabilization of the temperature corresponds to a balance between
the mechanical energy dissipated into heat and the energy lost by convection and
radiation at the specimen surface and by conduction inside the specimen. The
higher the stress amplitude, the higher is the rate of temperature increase. The
mean steady-state temperature elevation for each loading step defined as
hf ¼ Tsteady  Tinitial is plotted versus stress amplitude in Fig. 6. From this figure, it
is clear that, by increasing the stress amplitude, the temperature elevation increases;
for stress amplitudes lower than 60 MPa, the temperature increase remains lower
than 10  C. However, it reaches around 70  C for higher stress amplitudes. As an
instance, the specimen steady-state temperature field obtained from IR thermogra-
phy is depicted in Fig. 7 for the stress amplitude of 102 Mpa. This figure shows that
the temperature distribution is symmetrical.
Fig. 8 shows the values of the time constant s obtained by matching Eq 6 and
the evolution curve of temperature with time after the stop of the fatigue test for the
various stress amplitudes displayed in Fig. 5. By increasing the stress amplitude up to
127 MPa, s decreases from 20.5 s to 8 s; however, after this point, it reaches a plateau
and remains approximately constant by increasing the stress. Taking into account
the value of s for the different stress amplitudes, the mean dissipated energy per cy-
cle is calculated from the steady-state temperature. Fig. 9 depicts the change in dissi-
pated energy per cycle as a function of stress amplitude. From this figure, it is clear
that the higher the stress amplitude, the higher is the dissipated energy. Moreover,
the dissipated energy per cycle is a quadratic function of stress amplitude.

FIG. 6 Temperature increase versus stress amplitude for DP600 steel under ultrasonic
fatigue loading.
TORABIAN ET AL., DOI 10.1520/STP159820160053 89

FIG. 7 Temperature field on the specimen surface for ra ¼ 102 Mpa at N ¼ 8  106 cycles.

FIG. 8 The time constant, s, versus stress amplitude.

25

20

15
τ (s)

10

0
0 50 100 150 200 250 300
Stress Amplitude (MPa)
90 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 9 Mean dissipated energy per cycle versus stress amplitude for the DP600 steel
under ultrasonic loading.

Discussion
For low stress amplitudes, below 250 MPa, the change in self-heating and dissi-
pated energy per cycle with increasing temperature displays a quadratic form;
therefore, a gradual increase in the slope of the curve is observed. Considering
that the dissipated energy is due to an anelastic or inelastic material behavior
characterized by constant properties leads to the following expressions for the dis-
sipated energy [16]; assuming a prescribed sinusoidal stress with the amplitude ra
and a Kelvin-Voigt model (spring and dashpot in parallel), the dissipated energy
per cycle is written as [16]:

d~1 1 2gðpra Þ2
¼ 2 (11)
f fl
þ 4p2 g2
f2

where l and g are the elastic and viscous moduli, respectively.


In the case of a pure inelastic behavior and a zero mean stress, the dissipated
energy per cycle is written as [16]:

d~1 1 ra 2
¼ : (12)
f f 2g

In both cases, in the case of constant material properties (l and g), the dissipated
energy per cycle is a quadratic function of the stress amplitude. In this work, as the
TORABIAN ET AL., DOI 10.1520/STP159820160053 91

dissipated energy per cycle is found to be a quadratic function of the stress amplitude
for low stress amplitudes, it can be assumed that the material internal state remains
nearly the same during cyclic loading. Because the hardness of the martensite
is much higher than ferrite, dislocations are assumed to move only in the fer-
ritic phase, which has a body-centered cubic (bcc) structure. In bcc metals,
the flow stress strongly depends on temperature and strain rate at low tem-
peratures. Mughrabi, Herz, and Stark [17] describe the temperature-dependent de-
formation of bcc metals by introducing two deformation modes based on a transition
temperature, T0. In the low-temperature (or thermally activated) mode (T < T0), the
screw dislocations are immobile and the edge dislocations move to and fro in a non-
hardening quasi-recoverable manner. On the other hand, in the athermal mode
(T > T0), mobilities of the screw and edge dislocations are comparable, and screw
dislocations can cross slip easily. In this mode, strain localization can occur on slip
bands, and more energy can be dissipated because of high dislocation mobilities.
The transition temperature, T0, largely depends on strain rate and can be shifted to
higher values at high strain rates. For instance, according to the flow behavior dia-
gram reported by Campbell and Ferguson [18] for a 12 wt % carbon mild steel, T0
would shift from 25  C to 100  C by increasing the strain rate from 0.01 s1 to 1 s1.
For ultrasonic fatigue loadings at low stress amplitudes from 60 to 247 MPa, the
maximum strain rate ranges from 34 s1 to 141 s1. Therefore, for such high
strain rates, the transition temperature for the DP600 should be higher than
room temperature. As a result, and as suggested by Favier et al. [14] for alpha-
iron, the thermally activated mode, which is typical of a bcc structure, prevails at
room temperature for a 20-kHz cyclic loading at low stress amplitudes. There-
fore, the dissipated energy probably results from the to-and-fro motion of dislo-
cations that result in slight changes in the material internal state and are
consistent with Eqs 11 and 12.
In addition, the estimated values for dissipated energy per cycle are in agree-
ment with results reported in the literature for alpha-iron. For instance, for the
stress amplitude of 200 MPa, which is equal to 30 % of the ultimate stress ampli-
tude, the dissipated energy per cycle is 800 J=m3=cycle. This is consistent with the
results obtained by Favier et al. for alpha-iron [14].

Conclusions
In this work, ultrasonic fatigue tests along with in situ IR thermography are con-
ducted on flat specimens of DP600 steel. The self-heating diagrams are devel-
oped under ultrasonic loading for low stress amplitudes, and the dissipated
energy per cycle is also determined. It is observed that, at low stress amplitudes,
the dissipated energy per cycle is a quadratic function of the stress amplitude.
The to-and-fro glide of the edge dislocations is considered to be the main dissi-
pative mechanism. This motion is quasi-recoverable and leaves a quasi-constant
dislocation structure (no cyclic hardening).
92 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

References

[1] Bathias, C., “There Is No Infinite Fatigue Life in Metallic Materials,” Fatigue Fract. Eng.
Mater. Struct., Vol. 22, No. 7, 1999, pp. 559–565.

[2] Wu, T. and Bathias, C., “Application of Fracture Mechanics Concept to Ultrasonic
Fatigue,” Eng. Fract. Mech., Vol. 47, No. 5, 1994, pp. 683–690.

[3] Stanzl-Tschegg, S. E., Mayer, H. R., and Tschegg, E. K., “High Frequency Method for
Torsion Fatigue Testing,” Ultrasonics, Vol. 31, No. 4, 1993, pp. 275–280.

[4] Wang, Q. Y., Berard, J. Y., Dubarre, A., Baudry, G., Rathery, S., and Bathias, C., “Gigacycle
Fatigue of Ferrous Alloys,” Fatigue Fract. Eng. M., Vol. 22, No. 8, 1999, pp. 667–672.

[5] Furuya, Y., Matsuoka, S., Abe, T., and Yamaguchi, K., “Gigacycle Fatigue Properties for
High-Strength Low-Alloy Steel at 100 Hz, 600 Hz, and 20 kHz,” Scripta Mater., Vol. 46,
No. 2, 2002, pp. 157–162.

[6] Marines, I., Dominguez, G., Baudry, G., Vittori, J. F., Rathery, S., and Doucet, J. P.,
“Ultrasonic Fatigue Tests on Bearing Steel AISI-SAE 52100 at Frequency of 20 and 30
kHz,” Int. J. Fatigue, Vol. 25, No. 9, 2003, pp. 1037–1046.

[7] Boulanger, T., Chrysochoos, A., Mabruand, C., and Galtier, A., “Calorimetric Analysis of
Dissipative and Thermoelastic Effects Associated with the Fatigue Behavior of Steels,”
Int. J. Fatigue, Vol. 26, No. 3, 2004, pp. 221–229.

[8] Munier, R., Doudard, C., Calloch, S., and Weberb, B., “Towards a Faster Determination of
High Cycle Fatigue Properties Taking into Account the Influence of a Plastic Pre-Strain
from Self-Heating Measurements,” Procedia Eng., Vol. 2, No. 1, 2010, pp. 1741–1750.

[9] Doudard, C., Calloch, S., Hild, F., and Roux, S., “Identification of Heat Source Fields from
Infrared Thermography: Determination of ‘Self-Heating’ in a Dual-Phase Steel by Using
a Dog Bone Sample,” Mech. Mater., Vol. 42, No. 1, 2010, pp. 55–62.

[10] Doudard, C. and Calloch, S., “Influence of Hardening Type on Self-Heating of Metallic
Materials under Cyclic Loadings at Low Amplitude,” Eur. J. Mech. A-Solids, Vol. 28, No. 2,
2009, pp. 233–240.

[11] Doudard, C., Calloch, S., Hild, F., Cugy, P., and Galtier, A., “Identification of the Scatter in
High Cycle Fatigue from Temperature Measurements,” C. R. Mecanique, Vol. 332, No. 10,
2004, pp. 795–801.

[12] Blanche, A., Chrysochoos, A., Ranc, N., and Favier, V., “Dissipation Assessments during
Dynamic Very High Cycle Fatigue Tests,” Exp. Mech., Vol. 55, No. 4, 2015, pp. 699–709.

[13] Ranc, N., Blanche, A., Ryckelynck, D., and Chrysochoos, A., “POD Preprocessing of IR
Thermal Data to Assess Heat Source Distributions,” Exp. Mech., Vol. 55, No. 4, 2015,
pp. 725–739.

[14] Favier, V., Blanche, A., Wang, C., Phung, N. L., Ranc, N., Wagner, D., Bathias, C., Chryso-
choos, A., and Mughrabi, H., “Very High Cycle Fatigue for Single Phase Ductile Materials:
Comparison between a-Iron, Copper, and a-Brass Polycrystals,” Int. J. Fatigue, Vol. 93,
No. 2, 2016, pp. 326–338.
TORABIAN ET AL., DOI 10.1520/STP159820160053 93

[15] Munier, R., Doudard, C., Calloch, S., and Weber, B., “Determination of High Cycle Fatigue
Properties of a Wide Range of Steel Sheet Grades from Self-Heating Measurements,”
Int. J. Fatigue, Vol. 63, 2014, pp. 46–61.

[16] Mareau, C., Favier, V., Weber, B., and Galtier, A., “Influence of the Free Surface and the
Mean Stress on the Heat Dissipation in Steels under Cyclic Loading,” Int. J. Fatigue,
Vol. 31, Nos. 8–9, 2009, pp. 1407–1412.

[17] Mughrabi, H., Herz, K., and Stark, X., “Cyclic Deformation and Fatigue Behavior of a-Iron
Mono- and Polycrystals,” Int. J. Fract. Vol. 17, No. 2, 1981, pp. 193–220.

[18] Campbell, J. D. and Ferguson, W. G., “The Temperature and Strain-Rate Dependence of
the Shear Strength of Mild Steel,” Philos. Mag., Vol. 21, No. 169, 1970, pp. 63–82.
94 FATIGUE AND FRACTURE TEST PLANNING, TEST DATA ACQUISITIONS AND ANALYSIS

STP 1598, 2017 / available online at www.astm.org / doi: 10.1520/STP159820160094

Raghu V. Prakash,1 Mathew John,2 Deepika Sudevan,3


Andrea Gianneo,4 and Michele Carboni5

Fatigue Studies on Impacted and


Unimpacted CFRP Laminates
Citation
Prakash, R. V., John, M., Sudevan, D., Gianneo, A., and Carboni, M., “Fatigue Studies on
Impacted and Unimpacted CFRP Laminates,” Fatigue and Fracture Test Planning, Test Data
Acquisitions and Analysis, ASTM STP1598, Z. Wei, K. Nikbin, P. C. McKeighan, and D. G. Harlow,
Eds., ASTM International, West Conshohocken, PA, 2017, pp. 94–118, http://dx.doi.org/
10.1520/STP1598201600946

ABSTRACT
Carbon fiber-reinforced polymer (CFRP) composites are widely used in safety-
critical aerospace structures and experience low-velocity impacts, which
deteriorate their residual strength and residual life. This paper presents the results
of damage accumulation during fatigue loading, estimated through stiffness
degradation under constant amplitude and the programmed version of FALSTAFF
spectrum loading (Prog-FALSTAFF) of [0/90]8 and quasi-isotropic CFRP
laminates. The laminates are prepared by a hand lay-up technique; specimens of
size 250 mm by 45 mm by 4.5 mm thickness are extracted from the laminate for
testing. Some of the specimens are subjected to drop impact at different input
impact energies (23J, 35J, and 51J), and the fatigue response after impact loading
is studied. The rate of degradation of stiffness for impacted CFRP specimens
under a Prog-FALSTAFF spectrum loading is less compared to constant

Manuscript received April 20, 2016; accepted for publication October 5, 2016.
1
Indian Institute of Technology, Dept. of Mechanical Engineering, Chennai 600 036, India http://orcid.org/
0000-0002-8888-022X
2
Indian Institute of Technology Madras, Dept. of Mechanical Engineering, Chennai 600 036, India http://
orcid.org/0000-0001-8504-7016
3
Indian Institute of Technology Madras, Dept. of Mechanical Engineering, Chennai 600 036, India http://
orcid.org/0000-0003-0704-8712
4
Politecnico di Milano, Dept. of Mechanical Engineering, Via La Masa 1, Milano, 20156, Italy http://orcid.org/
0000-0002-6102-6381
5
Politecnico di Milano, Dept. of Mechanical Engineering, Via La Masa 1, Milano, 20156, Italy http://orcid.org/
0000-0003-0828-9067
6
ASTM Symposium on Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis on May 4–5,
2016 in Grand Hyatt, San Antonio, TX.

Copyright V
C 2017 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.
PRAKASH ET AL., DOI 10.1520/STP159820160094 95

amplitude loading. The sequence of intermediate block cycles of constant


ampltiude in Prog-FALSTAFF is organized either in an ascending order of
magnitude, low to high (Lo-Hi), or in a descending order of magnitude (Hi-Lo)
to study the effect of load sequencing on damage accumulation. Lo-Hi
sequencing results in greater damage compared to Hi-Lo sequencing for all the
orientations of CFRP laminates studied. In the case of unimpacted specimens,
there is a marginal increase in stiffness in the early stages, followed by a drop in
stiffness; the stiffness degradation rate in unimpacted specimens is much less
compared to impacted specimens. Computed X-ray tomography has been
carried out on specimens under the following conditions: pristine; after fatigue
cycling; after impact damage; and postimpact, postfatigue cycling to
understand the damage progression. As an exploratory study, investigation of
damage through active infrared thermography is carried out. The presence of
delamination could be identified through the variable specimen cooling curves
of CFRP specimens using the active thermography technique.

Keywords
woven carbon fiber-reinforced polymer specimens, stiffness, fatigue damage,
impact damage, computed tomography, infrared thermography

Introduction
Understanding the fatigue response of fiber-reinforced composite materials is
important for the design of safety-critical, lightweight, energy-efficient aircraft
structures. During service, these composite structures are subjected to variable am-
plitude (VA) loadings, which consist of several high (Hi) and low (Lo) loadings;
these VA loadings are very different from constant amplitude loadings that are typi-
cally used for research studies. As a consequence, the way fatigue damage accumu-
lates in these materials is different from that of constant amplitude loading, which
could be further accentuated by the sequencing of loading (Lo loads followed by Hi
loads or vice versa). The classical Palmgren-Miner method has been found to be
satisfactory in predicting the lifetime of metallic materials subjected to VA loadings
[1]; it is based on the assumption that the total relative lifetime is a linear sum of
relative lifetimes at various amplitudes of loading.
In the case of polymer composite materials, the failure modes are very different
from that of metals, and these exhibit failures due to various reasons such as [2]: fi-
ber breakage, fiber-matrix debonding, and matrix yielding at the microscopic and
ply levels, which could be failure due to transverse cracking, shear failure, or delam-
ination (or any combination thereof). As a consequence, the Palmgren-Miner
method, which is based on the assumption that the total relative lifetime is a linear
sum of relative lifetimes at various amplitudes of loading, has been known to give
unsatisfactory predictions for composite materials. Even in the case of metals, it is
well-documented that sequencing effects can alter fatigue damage [3]. In the case of
composites, the literature suggests that a Lo-Hi sequence can result in shorter
96 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

lifetimes compared to a Hi-Lo sequence [4], although a few studies suggest that a
Hi-Lo sequence results in shorter lifetimes [5,6]. A general statement cannot be
made with regard to the load sequencing effects of polymer composites. Eftekhari,
Fatemi, and Khosrovaneh [7] have shown that, in the case of short fiber composites,
both Lo-Hi and Hi-Lo sequences have shown a strengthening effect.
Most often, composite structures are subjected to low-velocity impact loading
due to accidental dropping of tools, hail storms, bird strikes, and so on that intro-
duce damage to these materials, causing a drop in residual strength and residual life
of built-up laminate structures. In some cases, the damage is barely visible (referred
to as BVID) and, as a consequence, these structures may be operated without realiz-
ing the presence of damage in the laminates, which may result in accelerated dam-
age progression in the composite material. Damage detection using nondestructive
testing (NDT) methods such as ultrasonic bulk waves, guided waves, specialized
bonding techniques as pitch-catch (PC), mechanical impedance analysis (MIA),
and resonance, as well as active thermography have been attempted in aircraft
structures; an exaustive literature review is given in Aerospace NDT [8].
Damage progression in composite materials is widely studied through ultrason-
ic C-scan imaging, which identifies the extent of damage in the composite material
by means of appropriate gated information. However, the use of C-scan for practi-
cal applications is typically time-consuming; as a result, often the damage response
is studied using alternate methods such as stiffness degradation, frequency response
to an excitation, and through NDT methods such as infrared (IR) thermography—a
qualitative characterization technique at the present time. In general, the ultrasonic
C-scan method is exhaustive and requires adequate specimen preparation to ensure
the true damage is identified, without noise, especially when there are changes in
sections as well as when there is a curved member. Recently, computed tomography
(CT) has been used as an effective tool for studying damage in materials [8].
The purpose of this research is to examine the fatigue response of carbon fiber-
reinforced plastic (CFRP) laminates in both unimpacted (pristine) and impacted
conditions under constant amplitude as well as under VA loading. A programmed
version of a European standard Fighter Aircraft Loading STAndard for Fatigue
(FALSTAFF) [9] was used to study the response of CFRP laminates under tension–
tension loading conditions. Typically, a full FALSTAFF spectrum consists of 17,983
cycles (35,966 reversals) representing 200 flights [10]. A Mini-FALSTAFF was
derived earlier [11] from the FALSTAFF spectrum by omitting small amplitude cycles
that do not cause significant damage. The Mini-FALSTAFF spectrum consists of
1424 cycles. Programmed FALSTAFF is derived from Mini-FALSTAFF by retaining
the 18 major cycles of the FALSTAFF spectrum and reorganizing the intermediate
cycles’ ranges as blocks of constant amplitude cycles. The Prog-FALSTAFF consists
of 1242 cycles in total, which includes the 18 major cycles. The balance, 1224 cycles,
are divided into 18 identical blocks, each having 68 cycles. Equivalence of fatigue
damage between Mini-FALSTAFF and Prog-FALSTAFF was established earlier on
metallic materials [9]. The primary purpose of programming the FALSTAFF as a
PRAKASH ET AL., DOI 10.1520/STP159820160094 97

block of constant amplitude loads interspersed with major cycles was to punch the
beach marks on the fracture surface of metallic materials by the first author along
with other coauthors as part of notch root crack growth rate studies [12]. The same
programmed version of FALSTAFF spectrum is used in this study to understand the
damage progression in composites. It is possible that the damage mechanics in com-
posites could be very different from that of metallic materials, but the purpose of this
study is to identify the variability in damage, if any, due to Lo-Hi and Hi-Lo load
sequencing.
Because damage in general leads to changes in stiffness, a stiffness degradation
method is used as a primary tool in this study to understand the fatigue damage ac-
cumulation in unimpacted and impacted woven-mat CFRP specimens. Exploratory
work has been carried out to understand the damage progression through IR ther-
mography. There are different methods of using IR thermography: active and pas-
sive. In the case of passive thermography, the thermal imaging is done online and
the radiation component of heat emitted by the composite through self-heating (or
hysteretic heating response) is captured. In the case of active thermography, the ra-
diation response to heating of the composite laminate during the cooling stage can
be used to characterize the damage state. The present work uses active thermogra-
phy to identify the fatigue damage response at intermediate stages. In addition to
the aforementioned NDT methods, CT image reconstruction of CFRP specimens is
carried out in: (a) the pristine condition, (b) after fatigue cycling, (c) after impact
damage, and (d) after post-impact fatigue damage to understand the various stages
of damage progression in the laminate.

LITERATURE REVIEW
In a review of fatigue analysis of composites, Filis, Farrow, and Bond [13] have indi-
cated that the fatigue life prediction capability of CFRP laminates is dependent on
the lay-up sequence. Because carbon fibers have low strain to fracture, and high
strength and modulus, fatigue response for fibers along the loading direction pro-
vides unconservative estimates of life. In the case of angle-ply fastener specimens
under tensile FALSTAFF loading, standard fatigue analysis results are found to be
unconservative; whereas, under compressive FALSTAFF loading, fatigue analysis
results are reasonably accurate [13]. In the case of woven composites, fatigue dam-
age is dependent on the geometrical arrangement of fibers [14].
Hoover, Kujawski, and Ellyin [15] correlated the stiffness reduction during a
tensile test with the crack density in a composite laminate through a linear relation-
ship. Three stages of damage progression were reported: the first stage—where there
is no transverse cracking, the second stage—where there is a linear stiffness degra-
dation where almost all transverse cracks had initiated, and the final stage—where
there is a significant drop in stiffness due to large-scale delamination. Salkind [16]
suggested that changes in stiffness can be used to characterize damage in composites;
however, the stiffness measurement cannot identify the different micromechanics of
98 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

failure in composites. Whitworth [17] noted that the rate of stiffness reduction is in-
versely proportional to the number of cycles to failure.
Gamstedt and Sjögren [4] observed that, in a cross-ply CFRP laminate dur-
ing fatigue loading, transverse cracks and delamination growth occur due to ini-
tiation and progression mechanisms, respectively, where transverse cracks
dominate the delamination growth at high amplitudes and vice versa at low
amplitudes. Thus, it was concluded that the Hi-Lo loading amplitude sequence is
more damaging than the Lo-Hi sequence, though many researchers [5,6] have
observed the reverse order could be more damaging. Many of these observations
relate to unimpacted CFRP systems. Gamstedt and Sjögren [4] have indicated
that composite materials have a memory, and the development of damage is de-
pendent on the order of the various stress levels. Kang and Kim [18] studied the
post-impact fatigue behavior of CFRP under a two-stage tensile loading and in-
troduced a strength reduction concept based on Broutman and Sahu’s model
[19] under three-stage loading. Behesty and Harris [20] studied the fatigue re-
sponse of CFRP laminates after low-velocity impact and observed that the con-
stant life model proposed by Harris et al. [21] can be used to predict fatigue life
for low energy impact–damaged specimens.
Clark and van Blaricum [22] conducted spectrum loading fatigue studies on
low energy impact–damaged CFRP specimens and observed that the major cycles
contribute to fatigue damage compared to several small amplitude cycles in the
spectrum. This would imply that any changes in the sequence of minor cycles in a
spectrum loading should not affect the outcome of fatigue damage. In other words,
Hi-Lo or Lo-Hi of minor cycles need not influence fatigue damage accumulation
during spectrum loading.
Toubal, Karama, and Lorrain [14] evaluated the fatigue damage evolution in
woven CFRP using IR thermography. The observed increase in specimen tempera-
ture was an index of the damage evolution during cycling. The temperature evolu-
tion was observed to be in three stages—similar to damage progression curves
based on stiffness degradation. The initial heating is due to friction between the
fibers and the fiber and matrix; the steady-state response in the second stage is due
to the balance of mechanisms of self-heating of specimens, and the rapid increase
in final stages is due to extensive delamination prior to final failure of fiber compo-
sites. Steinberger et al. [23] have used the pulse thermography technique to deter-
mine the heat conductivity in the thickness direction for CFRP laminates and
observed that IR thermography can be used to characterize damage due to fatigue
in composite specimens.

Test Method/Overview
SPECIMEN PREPARATION
Woven carbon fabric (HCT 502) of 500 gsm, 0.42 mm thick, sourced from Hindoo-
stan Technical Fabrics, is used as the reinforcement in an epoxy resin matrix system
PRAKASH ET AL., DOI 10.1520/STP159820160094 99

consisting of AraldyteV R LY 556 as resin and AradurV R HY 951 as the hardener.

Laminates are prepared using the hand lay-up technique, by laying up eight layers
of [0/90] fabric with epoxy between the layers. Similarly, quasi-isotropic laminates
with a stacking sequence of [0/90/645/0/90/645]s are prepared for understand-
ing fatigue damage progression. The laminates are cured at 80 C for 3 h; the finish
size of laminates was approximately 350 by 350 mm with a final thickness of
4.5 mm, which includes the thickness of woven fabric and epoxy. The fiber volume
fraction is found to be approximately 0.5. The test specimens are then sliced on a
computerized numerical controlled (CNC) cutting machine to a finish size of 250 by
45 by 4.5 mm in thickness (Fig. 1a). The dimensions and orientation of the [0/90]8
specimens are the same as reported by Sudevan, Prakash, and Kamaraj [24].
Because the test matrix involves testing some unimpacted specimens, fatigue dam-
age can occur at any location between the gripping portions of the test specimen,
which will make the experimental observations of damage progression difficult. To
overcome this, some quasi-isotropic laminate specimens are machined to an hour-
glass shape (Fig. 1b) to ensure damage concentration at the minimum width sec-
tion. The narrowed-down width of the hourglass specimen is approximately
35 mm, while the width at the grip region is maintained at 45 mm. The typical
stress concentration for an hourglass specimen using isotropic material constants
is observed to be 1.39.

DROP IMPACT TESTING


An in-house-developed drop-impact tester as shown in Fig. 2, with an impactor of
mass 5.2 kg and a hemispherical tip of approximately 16 mm in diameter, is used

FIG. 1 (a) Pristine, unimpacted specimen (similar to those used by Sudevan, Prakash,
and Kamaraj [24]) and (b) hourglass specimen.

(a)

(b)
100 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 2 Photograph of drop impact testing machine.

for causing an impact damage at three levels (23J, 35J, and 51J of energy) on the
composite laminate. The energy levels are calculated based on the mass of impact
as well as on the height of drop. Typically, the impact velocities below 5 m/s cause
low-velocity impact damage on the specimens. The load during impact is measured
using a load cell and the displacement is measured using an LVDT that is mounted
beneath the drop impactor. Care is taken to ensure that the impactor makes only a
single impact (i.e., without re-bounce) by designing a stopper for arresting multiple
impacts on the specimen.
The load-displacement data are recorded online and used for calculating the
energy absorbed by the specimen during drop impact testing by taking into consid-
eration the area under the load-displacement curve. Fig. 3a presents the load-
displacement data obtained during one of the drop impact tests on [0/90]8 CFRP
laminates. Typical load-displacement records during drop impact on quasi-isotropic
laminates at different levels of input impact energy (23J, 35J, and 51J) are shown in
Fig. 3b. It can be observed that higher impact energy levels result in an increase in
force as well as displacement of the specimen. This translates into an increase of
energy absorption by the specimen, which is calculated from the load-displacement
PRAKASH ET AL., DOI 10.1520/STP159820160094 101

FIG. 3 Load-displacement response during drop impact: (a) [0/90]8 laminate during
51J impact [24] and (b) quasi-isotropic laminate at various impact energies.

(a) Load vs. Displacement for [0/90]8


500
400
Load, N
300
200
100
0
0 2 4 6 8 10 12
Displacement, mm

(b) Load vs. Displacement - Quasi-Isotropic


600

400
Load, N

200 23J Hi-Lo


51J CA
35J L-1
0
0 5 10 15
Displacement, mm

records. Fig. 4 shows the energy absorption levels of the quasi-isotropic laminates for
the various levels of input impact energy.
The failure zone is observed to be visually concentrated over the first few layers
of the laminate (Fig. 5). The specimens are further examined using immersion-type

FIG. 4 Impact energies absorbed by quasi-isotropic CFRP laminates.


Energy Absorbed by Specimen, J

Impact Energy Absorbed - Quasi Isotropic CFRP Laminate


6
5
y = 0.001x2 + 0.051x - 0.2807
4
R2 = 1
3
2
1
0
0 10 20 30 40 50 60
Input Impact Energy, J
102 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 5 Quasi-isotropic laminate after: (a) 23J impact energy and (b) 51J impact
energy.

(a) (b)

ultrasonic testing. A representative C-scan image of an [0/90]8 impacted specimen


obtained by the pulse-echo method is shown in Fig. 6a. The ultrasonic imaging
parameters are as follows:
• Transducer: Probe frequency of 2.25 MHz and focal length approximately
63 mm (from Panametrics NDT)
• Trigger source: DPR 300 Pulser-Receiver from JSR Ultrasonics
• Scan length: 252 mm
• Scan resolution: 0.1 mm
• Index length: 50 mm
• Index resolution: 0.2 mm
• Voltage: 100 V
• Energy level: 2
• Pulse repetition frequency (PRF): 800 Hz
• Gain: 38 dB
• Damping: 52
• Average sample thickness: 4.2 mm
• Couplant: Water

An IR thermal image of the impacted specimen is obtained using a Micro-Epsilon


TIM 160 IR camera in the passive mode (i.e., measurement of specimen temperature
during mechanical loading without any external thermal excitation of the laminate)
and the same is shown in Fig. 6b. Typically, the hot spot representing the impact-
damaged region shows a temperature of 31.4 C compared to a cold spot tempera-
ture of 27.9 C. In the same figure, the temperature profiles for the regions away from
the impact damaged zone (i.e., the bottom side of the impact zone and the top side of
the impact zone) are shown, where the temperature is not as high as in the encircled
impact zone. Thus, the impact damaged area could be identified during mechanical
loading of the specimen through the passive thermography technique.

TENSILE AND FATIGUE TESTING


Tensile tests are conducted on unimpacted and impacted laminate specimens using
a 100 kN-MTS 810 servohydraulic testing machine under displacement control
mode at a loading rate of 0.1 mm/min. Typically, unimpacted [0/90]8 specimens
PRAKASH ET AL., DOI 10.1520/STP159820160094 103

FIG. 6 (a) C-scan and A-scan image of impact damage of a [0/90]8 laminate [24] and
(b) passive thermal image of the [0/90]8 specimen during mechanical loading,
showing the region of impact damage. The hot spot had a temperature of
31.4 C compared to the cold spot temperature of 27.9 C.

(a)

(b)

45 mm

have an ultimate tensile strength of 750 MPa, and the quasi-isotropic specimens
have an ultimate tensile strength of 313 MPa, while 51J-impacted [0/90]8 speci-
mens have a typical ultimate tensile strength of 202 MPa (approximately 27 % of
unimpacted ultimate tensile strength). Quasi-isotropic laminates subjected to 23J
impact have a tensile strength of 112 MPa (approximately 35 % of unimpacted
ultimate tensile strength). Fig. 7 shows the typical tensile test results for post-impacted
specimens with [0/90]8 [24] and quasi-isotropic configurations. It is observed that the
tensile strength of quasi-isotropic laminates are lower than the tensile strength values
obtained for [0/90]8 laminates. This is taken into consideration while deciding the load
levels for fatigue cycling. This means that the peak load applied on quasi-isotropic
laminates during fatigue loading is lower compared to [0/90]8 laminates.
104 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 7 Tension test data: (a) [0/90]8 laminate after 51J impact [24] and (b) quasi-
isotropic after 35J impact.

(a) Load vs. Displacement - Tension Test (b) Load vs. Displacement - Tensile Test
50
25
40 20

Force, kN
Force, kN

30 15
20 10
10 5
0
0
0 0.5 1 1.5 2 2.5 3 0 1 2
Displacement, mm Displacement, mm

Fatigue tests are conducted on an MTS 810 servohydraulic machine under con-
stant amplitude (CA) sinusoidal waveform cycling at a stress ratio (R) of 0.1 and at
a cyclic frequency of 5 Hz as per the test matrix shown in Table 1. The load-displace-
ment data during the fatigue test are continuously monitored for stiffness evalua-
tion. In the case of constant amplitude fatigue testing, the test is stopped after a
run-out criterion of 106 cycles.
The programmed FALSTAFF spectrum as detailed in Mitchenko, Prakash, and
Sunder [9] is applied to evaluate the fatigue performance under spectrum loading.

TABLE 1 Test parameters for fatigue cycling.

Sl. No. Specimen Condition Load Range (kN) Stress Range (MPa) Test Condition

1 Unimpacted [0/90]a 6.0–60 29.63–296.3 CA


2 Unimpacted [0/90]b 0–60 0–296.3 Prog-FALSTAFF Hi-Lo
3 Unimpacted [0/90] 0–60 0–296.3 Prog-FALSTAFF Lo-Hi
4 51J Impacted [0/90]a 2.4–24 11.85–118.5 CA
5 51J Impacted [0/90]b 0–24 0–118.5 Prog-FALSTAFF Hi-Lo
6 51J Impacted [0/90]a 0–24 0–118.5 Prog-FALSTAFF Lo-Hi
7 Unimpacted (Quasi) 4.3–43, 21.23–212.3, CA
2.25–22.5 10.4–104
8 23J Impacted (Quasi) 0.882–8.82 5.6–56 CA
9 23J Impacted (Quasi) 0–8.82 0–56 Prog-FALSTAFF Hi-Lo
10 23J Impacted (Quasi) 0–8.82 0–56 Prog-FALSTAFF Lo-Hi
11 35J Impacted (Quasi) 0.882–8.82 5.6–56 CA
12 35J Impacted (Quasi) 0–8.82 0–56 Prog-FALSTAFF Hi-Lo
13 35J Impacted (Quasi) 0–8.82 0–56 Prog-FALSTAFF Lo-Hi
14 51J Impacted (Quasi) 0.882–8.82 5.6–56 CA
15 51J Impacted (Quasi) 0–8.82 0–56 Prog-FALSTAFF Hi-Lo
16 51J Impacted (Quasi) 0–8.82 0–56 Prog-FALSTAFF Lo-Hi
a
Note: Stiffness data from this experiment also reported in Sudevan, Prakash, and Kamaraj [24].
b
Stiffness data also reported in Prakash and Sudevan [25]. CA ¼ constant amplitude.
PRAKASH ET AL., DOI 10.1520/STP159820160094 105

The 18 major loads of FALSTAFF spectrum are retained in their original order and
all minor loads are converted as equivalent blocks of marker loads as per Table 2;
the same is applied after every major load. The Prog-FALSTAFF spectrum has 18
major cycles, and between major cycles there are 5 intermediate blocks comprised
of 68 cycles of constant amplitude loading. Thus, the total number of cycles for the
Prog-FALSTAFF that represents 200 flights is 1242. Because the marker block in
Lo-Hi and Hi-Lo format always is attached to the rising half of the major cycle, one
would expect reduced influence of mean stresses on the fatigue damage compared
to the case where the marker block in Lo-Hi format is attached on the rising half
and Hi-Lo format is attached on the falling half of the major cycle.
For cases where the minor block of loads is applied in the Lo-Hi range, the
spectrum is designated as Prog-FALSTAFF Lo-Hi, and in cases where the minor
block loads are arranged in the order of Hi-Lo ranges, it is designated as Prog-
FALSTAFF Hi-Lo (Fig. 8). All negative loads are truncated to zero for the Prog-
FALSTAFF spectrum loading. This is necessary because the specimens are tested
without anti-buckling guides; the laminate strength and stiffness in compression
loading are not adequate; hence, all FALSTAFF experiments are conducted
under tension-tension loading. It is noted that a thorough spectrum load dam-
age analysis is required for this class of composite material because the damage
progression in composites is different from that of metals; this is being pursued
as a separate study.
Active thermography of fatigue-cycled CFRP specimens at intermediate stage
of fatigue cycling is studied for unimpacted as well as impacted specimens. The
thermal excitation of specimens is done using a filament lamp heating until they
attain a stable temperature in the range of 80 C to 100 C, and their cooling
response is monitored. A Micro-Epsilon IR camera (Model TIM 160) with a resolu-
tion of 160 by 120 pixels and a noise equivalent of temperature difference (NETD)
of 0.1 K is used for temperature measurements.
CFRP specimens are then inspected using a North Star Imaging 3D X-Ray
CT X25 tomographic system at PoliMi, Milano, Italy, with a voltage of 76 kV
and a current of 40 lA at the X-ray source. The detector used is a Dexela 1512 with
303 ms of integration time; the final voxel size is 26.1 lm.

TABLE 2 Marker block loading among 18 major loads of FALSTAFF spectrum (Programmed
FALSTAFF) [9].

Sl. No. Pmax Pmin No. of Cycles

1 83 % 0 2
2 75.5 % 0 4
3 70 % 0 10
4 65.5 % 4.21 % 17
5 58 % 8.15 % 35
106 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 8 Typical programmed block of FALSTAFF: (a) Major cycles of Prog-FALSTAFF


[12], (b) part of Lo-Hi sequence showing marker blocks inserted between major
cycles, and (c) part of Hi-Lo sequence showing marker blocks inserted between
major cycles. The start of the marker block is always from the trough (minimum
point) of the major cycle and begins with the peak load of the first marker block.

(a)

(b)

(c)

Results and Discussion


The unloading stiffness of pristine (unimpacted) as well as impacted specimens is
estimated after the first block of loading (in the case of Prog-FALSTAFF spectrum)
and after the first few cycles of loading for CA-loading specimens; this stiffness is
used as a reference for each stress level of fatigue testing, which is labeled as the
PRAKASH ET AL., DOI 10.1520/STP159820160094 107

initial stiffness of the specimen (K0). Stiffness data at periodic intervals of cycles are
estimated from the unloading segment of the load-displacement data over a win-
dow of 50 % to 90 % of maximum stress applied during fatigue cycling. Similarly,
loading stiffness is estimated from the loading segment of load-displacement data
over the window of 10 % to 50 % of maximum stress applied during fatigue cycling.
In the case of spectrum loading, stiffness data are estimated from load-displacement
data obtained from three constant amplitude cycles that are interspersed with
FALSTAFF loading at the end of every block (200 flights’ equivalent) of FALSTAFF
loading. It could be argued that measurement of stiffness from actuator displace-
ment could lead to errors in actual stiffness values exhibited by the CFRP speci-
mens. To address this, the specimens are all mounted at the same cross-head and
actuator displacement position. Independent of this, the stiffness value from a stiff
specimen is used to estimate the load-train compliance at similar maximum loads
applied during testing. The load-train has a typical stiffness of 80–100 kN/mm,
which is much higher compared to the specimen stiffness values. It may be noted
that, from a strict viewpoint of estimating specimen compliance, this method of es-
timating the stiffness may not provide accurate estimates of specimen stiffness; but,
if we consider that the total displacement of the actuator consists of specimen dis-
placement and load-train displacement, the load-train displacement is not likely to
change with fatigue cycles, whereas the specimen displacement would increase with
damage progression. For a given maximum load applied on the specimen (together
in series with load train), the stiffness is a function of the ratio of specimen displace-
ment to load-train displacement. The purpose of this study is to understand the
trend of damage progression through stiffness reduction, which can be satisfied with
the current approach of taking total actuator displacement for estimating stiffness.
Figs. 9 and 10 show the stiffness (K) versus the number of cycles for the unim-
pacted and impacted CFRP specimens tested under constant amplitude and pro-
grammed FALSTAFF spectrum loading. The stiffness of specimens is found to vary
over a small range for both impacted and unimpacted specimens, as shown in
Fig. 11. Higher stiffness values are found for fatigue tests with an unimpacted config-
uration compared to impacted specimens. Quasi-isotropic unimpacted specimens
show less stiffness compared to [0/90]8 unimpacted samples; the initial stiffness is
closer to 51J impacted [0/90]8 specimens. The loss of stiffness under constant am-
plitude loading—either loading stiffness, measured over a window of 10 % to 50 %
of max load, or unloading stiffness, measured over 90 % to 50 % of maximum load,
for quasi-isotropic lamina for 23J impact—is marginal (approximately 5 % to 9 %
of unimpacted specimens); it is about 15 % to 19 % for 35J impacted specimens and
about 20 % to 34 % for 51J-impacted specimens after the first block of spectrum
loading. Loss of stiffness is greater for programmed Lo-Hi FALSTAFF compared to
programmed Hi-Lo FALSTAFF for all impact energies (Table 3). However, constant
amplitude loading showed the greatest loss of stiffness for a given number of cycles,
followed by Lo-Hi FALSTAFF and Hi-Lo FALSTAFF loading. The last row in the
same table provides information about the loading and unloading stiffness loss
108 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 9 Stiffness versus number of cycles for [0/90]8 lay-up of woven CFRP laminates;
in the same graph, data are shown from a representative quasi-isotropic
laminate.

Stiffness vs. No. of cycles


Hi-Lo-UNIMP Lo-Hi-UNIMP Hi-Lo-51J-IMP
Lo-Hi-51J IMP CA-UNIMP CA-51J-IMP
QUASI-CA-UNIMPACTED

60

50
Stiffness, K, kN/mm

40

30

20

10

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 5.5 6 6.5 7 7.5
Cycles x 100,000

values for [0/90]8 laminates. Because the intrinsic stiffness of [0/90]8 laminates are
higher than the quasi-isotropic laminates, a gain (negative loss) in stiffness value is
reported.
With regard to our observation of greater damage under Prog-FALSTAFF Lo-
Hi compared to Prog-FALSTAFF Hi-Lo, there are claims and counterclaims in the
literature as to which of the two sequences could be more damaging. In our
opinion, it depends on the visco-elastic nature of the material as well as on the fiber
lay-up sequence. The literature suggests that, in the case of woven mat composites,
friction at the interface during cyclic loading (even at low amplitudes) can increase
the temperature at the interface, close to the glass transition temperature (GTT) of
the material. If the volume fraction of resin is high, the GTT would also be affected
for the composite material, and it can affect the failure mode. Alternately, if the
visco-elastic response of the material is influencing its mechanical loading response,
application of Hi loads could cause greater damage due to shear strain rate and the
fact that the Lo-amplitude cycles that follow the Hi loads could be under this shad-
ow; thus, the damage does not progress.
It may be argued that the number of cycles of the same maximum and mini-
mum applied in the programmed FALSTAFF spectrum is much less compared to a
case where pure constant amplitude cycles are applied and that, as a consequence,
there is a significant drop in stiffness for the case of pure constant amplitude load
PRAKASH ET AL., DOI 10.1520/STP159820160094 109

FIG. 10 Stiffness versus number of cycles for quasi-isotropic lay-up of woven CFRP
laminates; in the same figure, data are reported from [0/90]8 laminate tested
under Prog-FALSTAFF Hi-Lo after 51J of drop impact.

Sffness vs. No. of Cycles


35

30
Sffness, K, kN/mm

25

20

15

10

0
0 100,000 200,000 300,000 400,000 500,000 600,000 700,000 800,000
No. of Cycles

23J CA 23J Hi-Lo 23J Lo-Hi HG eq 35J CA


35J Hi-Lo 35J Lo-Hi 51J CA 51J Hi-Lo
51J Lo-Hi 51J Hi-Lo (0/90) Un Impacted

FIG. 11 Stiffness variation for quasi-isotropic CFRP impacted and unimpacted


specimens.
110 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

TABLE 3 Percentage loss of stiffness for impacted specimens after first cycle of loading.

CA Lo-Hi FALSTAFF Hi-Lo FALSTAFF

Loading Unloading Loading Unloading Loading Unloading

Unimpacted 0 0 0 0 0 0
23J 9.35 5.00 6.84 4.97 3.19 0.95
35J 18.57 14.94 23.06 12.00 16.38 7.57
51J 33.78 20.73 44.95 36.38 41.25 26.14
51J ([0/90]8 81.57a 48.43a
a
Note: Negative values imply stiffness gain.

cycling. Because the S-N data and damage accumulation details for woven CFRP
laminates are not available at present, it is not possible to make a fair comparison
of damage potential of programmed spectrum loading as a fraction of damage in-
duced by extreme load constant amplitude cycles.
To address concerns regarding the use of limited test data to make the
aforementioned conclusions, the results of stiffness are non-dimensionalized with
reference to the stiffness data obtained after the first block of loading. Fig. 12
presents the normalized stiffness (K/K0) versus normalized cycles (N/Nf) data plot
for the [0/90]8 impacted specimens. K0 refers to the unloading stiffness from the
first cycle of fatigue measured over the window of 50 % to 90 % of peak load, and K
refers to the unloading stiffness for the periodic cycles over which the data are
collected. Fig. 13 presents the normalized stiffness (K/K0) against normalized cycles
(N/Nf) plot for quasi-isotropic laminates. The stiffness degradation has two distinct
regimes: initial rapid decrease followed by steady state; possibly the final drop in
stiffness is not captured because the specimens did not fail for the number of cycles

FIG. 12 Normalized specimen stiffness versus normalized cycles for [0/90]8 laminates.

Normalized impacted specimen sffness vs. Normalized cycles

Lo-Hi-51J-IMP Hi-Lo-51J-IMP CA-51J-IMP


1

0.8

0.6
K/Ko

0.4

0.2

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
N/Nf
PRAKASH ET AL., DOI 10.1520/STP159820160094 111

FIG. 13 Normalized stiffness variation as a function of normalized life for quasi-isotropic


laminates.

Normalized Stiffness vs No. of Cycles


1.2

1
Normalized stiffness, K/Ko

0.8

0.6

0.4

0.2

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Normalized life, N/Nf

23J CA 23J Lo-Hi 23J Hi-Lo 35J CA


35J Lo-Hi 35J Hi-Lo 51J CA 51J Lo-Hi
51J Hi-Lo UnImpacted 51J Hi-Lo(0-90)

they were subjected to. Determination of residual strength, post-impact, post-fatigue is


one of the objectives of the program; hence, all tests are not continued until failure in
fatigue. In such cases, 106 cycles are treated as Nf for the purpose of this graph. It is
found that the Hi-Lo load sequence of FALSTAFF results in less damage compared to
the Lo-Hi sequence. On the contrary, the damage progression in constant amplitude
loading is most severe in the cases considered.
From Figs. 12 and 13, it can be observed that the stiffness degradation is high in
the case of [0/90]8 laminates compared to quasi-isotropic laminates. Nearly 40 % loss
in stiffness after 20 % of life is noticed during constant amplitude loading for [0/90]8
orientation specimens impacted at 51J, when compared with quasi-isotropic lami-
nates with similar impact energies (where the drop is approximately 25 %). This
trend is observed consistently in all quasi-isotropic laminates. This implies that
though the initial strength and stiffness of quasi-isotropic laminates is lower than the
corresponding values for [0/90]8 laminates, their post-impact fatigue performance is
much superior. The extent of damage in [0/90]8 laminates as well as quasi-isotropic
laminates after a 51J impact is observed to be closer to 60 % (loss in stiffness) after
one million cycles. However, in the case of programmed FALSTAFF load spectrum
loading, quasi-isotropic laminates showed about a 20 % drop in normalized stiffness
compared to a 40 % drop in stiffness for [0/90]8 laminates.
112 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

To achieve a deeper insight about stiffness degradation because of fatigue load-


ing conditions, specimen status (impacted or pristine)—the behavior previously
highlighted—is investigated in a preliminary way by means of X-ray CT recon-
structions. Fig. 14 shows the side views from a volumetric CT reconstruction
for: (a) a pristine (unimpacted) specimen, (b) a 35J impacted specimen, (c) an

FIG. 14 CT volumetric reconstruction of (a) pristine (unimpacted), (b) 35J impacted,


(c) unimpacted, constant amplitude fatigue-loaded, and (d) 35J impacted and
constant amplitude fatigue-loaded, quasi-isotropic woven CFRP specimens.

(a) (b)

(c) (d)
PRAKASH ET AL., DOI 10.1520/STP159820160094 113

unimpacted, CA fatigue-cycled specimen, and (d) a 35J impacted and fatigue-


cycled specimen.
In Fig. 15, x-y-z CT slices for pristine (unimpacted), impact-damaged, fatigue-
damaged, and post-impact fatigue damaged specimens are presented to understand
the mechanism of damage progression and consequently the stiffness reduction in
woven composites. Fig. 15a suggests the presence of small pockets of resin-rich areas
and some places of air gap; but, the presence of any delamination or defect cannot
be seen from this image. From Fig. 15b, which corresponds to a 35J impacted CFRP
specimen, the following can be inferred: the side views suggest the incidence of an
impact damage in the specimen to an approximate depth of three to four layers of
laminate. Beyond this depth, the effect of impact damage is found to be negligible.
Discontinuities or delamination cracks in the specimen appear to be traversing
across the border of the specimen by cutting through the layers and flowing around
the fibers.

FIG. 15 CT slices along three cutting planes of (a) pristine (unimpacted), (b) 35J
impacted, (c) constant amplitude fatigue-loaded, and (d) 35J impacted and
constant amplitude fatigue-loaded, quasi-isotropic woven CFRP specimens.

(a) (b)

(c) (d)
114 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

Fig. 15c points out CT slices of an unimpacted quasi-isotropic CFRP specimen


subjected to CA fatigue-loading up to 600,000 cycles: it is worth noting that there is
very feeble delamination or crack growth. This behavior agrees with the experimen-
tal evidence of marginal stiffness reduction for an unimpacted specimen.
Lastly, Fig. 15d presents the x, y, and z plane CT slices for a 35J impacted and
post-impact fatigue-cycled (under constant amplitude loading up to 100,000 cycles)
quasi-isotropic specimen. It highlights significant damage extension and progres-
sion within this specimen, compared to CT images taken prior to fatigue cycling—
as shown in Figs. 15a and 15b as well as those images dealing with CA fatigue dam-
age (Fig. 15c). The mode of fracture appears to be three-dimensional where the
cracks weave around the thickness layers. Onset of delamination cracking can be
observed at intermediate depths from the impact region. Further studies are cur-
rently being carried out in order to quantify the damage zones in terms of damage
area fraction at different layer depths and correlate this information with loss of
stiffness, with the final aim of understanding the effect of load sequencing on dam-
age progression. Ideally, one would expect that the loss of stiffness could be due to
the reduced cross section of the laminate that participates in the load bearing. How-
ever, such a quick estimation does not appear to correlate with the experimental
data obtained through this study.
Fig. 16 presents the active IR thermograph of cooling for pristine sample and
post-impact fatigue under programmed FALSTAFF Hi-Lo specimens. The speci-
mens are heated using a halogen lamp until a stable temperature of about 90 C is
reached. Two combinations of heating (heating from the front face, which is the
same as the field of view for the IR camera, and heating from the back) and the
placement of the defective region in the front and backside of the field of view of
the IR camera are considered. Fig. 16 shows the results of a cooling curve for defect

FIG. 16 Cooling curve comparison between pristine (ND) and FALSTAFF-damaged


[0/90]8 (DB) specimen.

Cooling Curves-[0/90]8 Prisne and Post Impact FALSTAFF fague specimen

95
DB-HFF ND-HFF
85
Temperature, °C

75
65
55
45
35
25
0 200 400 600 800 1000 1200 1400
Time, s
PRAKASH ET AL., DOI 10.1520/STP159820160094 115

FIG. 17 Cooling curve comparison between quasi-isotropic pristine (ND), CA fatigue-


damaged specimens (DB) with [0/90]8 pristine (ND), and FALSTAFF-fatigued
specimens (DB).

Cooling Curves for Quasi-Isotropic, [0/90]8 Specimens


110
100 QUASI-HFF-DB
QUASI-HFF-ND
90
Temperature, °C

0/90- HFF-DB
80 0/90-HFF-ND
70
60
50
40
30
0 100 200 300 400 500 600 700
Time, s

in the back (DB) for [0/90]8 a post-impact fatigue-damaged specimen subjected to


FALSTAFF loading, which is compared with data for a no-defect (ND) specimen
with heating on the front face. It can be observed that the rate of cooling for the de-
fective specimen is higher than for the no-defect specimen, even though the defect
is actually at the back face of the specimen. This implies that, in the case of the ND
specimen, part of the heat supplied is conducted through the thickness of the speci-
men; hence, the radiation component of heat on the front face is less, which
explains the slower rate of cooling for the ND specimen.
Fig. 17 presents the cooling curve for quasi-isotropic as well as [0/90]8 lay-up
specimens obtained from active thermography for specimens tested under constant
amplitude loading conditions. Comparison is made with data obtained from un-
damaged (ND) specimens of both lay-up sequences. It is observed that there is a
distinct difference in the rates of cooling for quasi-isotropic and [0/90]8 composite
laminates; also, there is a distinct difference in cooling rates for fatigue-damaged
specimens versus undamaged specimens. This information can serve as a basis for
detection of the extent of damage in composite laminates once additional data are
available and an empirical fit of cooling rate versus damage fraction is established.
Further studies are in progress to link the location of damage with the IR imaging
to quantify the fatigue damage location in CFRP composites.

Conclusions
This paper presents the results of stiffness degradation due to fatigue loading under
constant, programmed FALSTAFF spectrum loading on woven-mat CFRP speci-
mens. The stiffness degradation for constant amplitude loading is observed to be
the highest, followed by the Lo-Hi version of programmed FALSTAFF; the smallest
116 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

amount of stiffness reduction is found in the Hi-Lo version of FALSTAFF spectrum


loading. To support the aforementioned results, NDT evaluations are performed by
means of X-ray CT volumetric image reconstruction as well as active thermogra-
phy. In particular, CT is an affordable NDT technique to bring about damage initia-
tion and extension within CFRP laminates; further studies are being carried out to
understand the effect of load sequencing on damage evolution. Active thermogra-
phy of undamaged as well as impact- and fatigue-damaged specimens suggests that
the cooling rates are distinctly different for each of these cases; one can qualitatively
see the difference in cooling rates based on the lay-up as well as the location of
damage in the CFRP laminates. Further work is in progress to generate additional
data which would help in developing an empirical model of defect location versus
thermal data.

ACKNOWLEDGMENTS
The authors would like to thank several individuals in the Mechanical Engineering,
Aerospace Engineering, Metallurgy, and Materials Engineering departments at the In-
dian Institute of Technology, Madras, for their support. The support received from
Ms. Hema Bhat at ADE, Bangalore, and Professor R. V. Mohan, NC A&T, Greens-
boro, NC, is gratefully acknowledged. X-ray CT volumetric image reconstruction anal-
yses were carried out in the AMALA Interdisciplinary Laboratory, Milano, Italy. The
first author acknowledges the support received from the Erasmus-Mundus Heritage
Action 2 program and thanks Professor Monica Bordegoni and Professor Stefano
Beretta of PoliMi, Italy, for their support and continued encouragement related to this
collaborative study.

References

[1] Miner, M. A., “Cumulative Damage in Fatigue,” Trans. ASME, J. Appl. Mech., Vol. 67, 1945,
pp. A159–A164.

[2] Mallick, P. K., Fiber-Reinforced Composites: Materials, Manufacturing, and Design, 3rd
ed., CRC Press, Boca Raton, FL, 2007.

[3] Prakash, R. V. and Sunder, R., “Fatigue Life Under Random Load History Derived from
Exceedance Curves Using Different Algorithms,” Fatigue Fract. Eng. M., Vol. 16, No. 7,
1993, pp. 707–721.

[4] Gamstedt, E. K. and Sjögren, B. A., “An Experimental Investigation of the Sequence
Effect in Block Amplitude Loading of Cross-Ply Composite Laminates,” Int. J. Fatigue,
Vol. 24, Nos. 2–4, 2002, pp. 437–446.

[5] Yang, J. N. and Jones, D. L., “Load Sequence Effects on Graphite/Epoxy [635]2s Lami-
nates,” Long-Term Behaviour of Composites, ASTM STP 813, T. K. O’Brien, Ed., ASTM
International, West Conshohocken, PA, 1983, pp. 246–262, http://dx.doi.org/10.1520/
STP31826S
PRAKASH ET AL., DOI 10.1520/STP159820160094 117

[6] Schaff, J. R. and Davidson, B. D., “Life Prediction Methodology for Composite Structures.
Part I. Constant Amplitude and Two-Step Level Fatigue,” J. Compos. Mater., Vol. 31,
No. 2, 1997, pp. 128–157.

[7] Eftekhari, M., Fatemi, A., and Khosrovaneh, A., “Fatigue Behavior of Neat and Short
Glass Fiber Reinforced Polymers Under Two-Step Loadings and Periodic Overloads,”
SAE Int. J. Mater. Manf., Vol. 9, No. 3, 2016, pp. 585–593.

[8] The American Society for Nondestructive Testing, ASNT Industry Handbook: Aerospace
NDT, ASNT, Columbus, OH, 2014.

[9] Mitchenko, E. I., Prakash, R. V., and Sunder, R., “Fatigue Crack Growth Under an Equivalent
FALSTAFF Spectrum,” Fatigue Fract. Eng. Mater. Struct., Vol. 18, No. 5, 1995, pp. 583–595.

[10] Van Dijk, G. M. and de Jonge, J. B., “Introduction to FALSTAFF,” Proceedings of the
Eighth ICAF Symposium, Lausanne, Switzerland, June 2–5, 1975, J. Branger, and
F. Berger, Eds., International Committee on Aeronautical Fatigue, Swiss Federal Aircraft
Establishment, Emmen, The Netherlands, 1975.

[11] Mitchenko, E. I., Prakash, R. V., and Sunder, R., “Fatigue Crack Growth Under Pro-
grammed Equivalent of FALSTAFF Spectrum,” NAL Project Document PD ST 9233,
National Aeronautical Laboratory, Bangalore, India, 1992.

[12] Sunder, R., Prakash, R. V., and Mitchenko, E. I., “Fractographic Study of Notch Fatigue
Crack Closure and Growth Rates,” Fractography of Modern Engineering Materials: Com-
posites and Metals, Second Volume, ASTM STP1203, J. E. Masters, and L. N. Gilbertson,
Eds., ASTM International, West Conshohocken, PA, 1993, pp. 113–131, http://dx.doi.org/
10.1520/STP14902S

[13] Filis, P. A., Farrow, I. R., and Bond, I. P., “Classical Fatigue Analysis and Load Cycle Mix-
Event Damage Accumulation in Fiber Reinforced Laminates,” Int. J. Fatigue, Vol. 26,
No. 6, 2004, pp. 565–573.

[14] Toubal, L., Karama, M., and Lorrain, B., “Damage Evolution and Infrared Thermography in
Woven Composite Laminates Under Fatigue Loading,” Int. J. Fatigue, Vol. 28, No. 12,
2006, pp. 1867–1872.

[15] Hoover, J. W., Kujawski, D., and Ellyin, F., “Transverse Cracking of Symmetric and Unsym-
metric Glass-Fibre/Epoxy-Resin Laminates,” Compos. Sci. Technol., Vol. 57, No. 11, 1997,
pp. 1513–1526.

[16] Salkind, M. J., “Fatigue of Composites,” Composite Materials: Testing and Design (2nd
Conference), ASTM STP497, H. T. Corten, Ed., ASTM International West Conshohocken,
PA, 1972, pp. 143–169, http://dx.doi.org/10.1520/STP27745S

[17] Whitworth, H. A., “A Stiffness Degradation Model for Composite Laminates Under
Fatigue Loading,” Compos. Struct., Vol. 40, No. 2, 1998, pp. 95–101.

[18] Kang, K.-W. and Kim, J.-K., “Fatigue Life Prediction for Impacted Carbon/Epoxy Lami-
nates Under 2-Stage Loading,” Compos. Pt. A, Vol. 37, No. 9, 2006, pp. 1451–1457.

[19] Broutman, L. J. and Sahu, S., “A New Theory to Predict Cumulative Fatigue Damage in
Fiberglass Reinforced Plastics,” Composite Materials: Testing and Design (2nd Confer-
ence), ASTM STP497, H. T. Corten, Ed., ASTM International West Conshohocken, PA,
1972, pp. 170–188, http://dx.doi.org/10.1520/STP27746S
118 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

[20] Behesty, M. H. and Harris, B., “A Constant-Life Model of Fatigue Behaviour for Carbon-
Fibre Composites: The Effect of Impact Damage,” Compos. Sci. Technol., Vol. 58, No. 1,
1998, pp. 9–18.

[21] Harris, B., Gathercole, N., Lee, J. A., Reiter, H., and Adam, T., “Life Prediction for Constant-
Stress Fatigue in Carbon Fibre Composites,” Philos. T. R. Soc. A., Vol. 355, No. 1727,
1997, pp. 1259–1294.

[22] Clark, G. and van Blaricum, T. J., “Load Spectrum Modification Effects on Fatigue of
Impact-Damaged Carbon-Fibre Composite Coupons,” Composites, Vol. 18, No. 3, 1987,
pp. 243–251.

[23] Steinberger, R., Valadas Leitao, T. I., Ladstatter, E., Pinter, G. H., Billinger, W., and Lang, R. W.,
“Infrared Thermographic Techniques for Non-Destructive Damage Characterization of Car-
bon Fibre Reinforced Polymers During Tensile Fatigue Testing,” Int. J. Fatigue, Vol. 28,
No. 10, 2006, pp. 1340–1347.

[24] Sudevan, D., Prakash, R. V., and Kamaraj, M., “Post-Impact Fatigue Response of CFRP
Laminates Under Constant Amplitude and Programmed FALSTAFF Spectrum Loading,”
Procedia Eng., Vol. 101, 2015, pp. 395–403.

[25] Prakash, R. V. and Sudevan, D., “Understanding Fatigue Damage Progression in Low
Velocity Impacted CFRP Laminates Through Stiffness Measurements,” Mater. Sci. Forum,
Vols. 830–831, 2015, pp. 413–416.
FATIGUE AND FRACTURE TEST PLANNING, TEST DATA ACQUISITIONS AND ANALYSIS 119

STP 1598, 2017 / available online at www.astm.org / doi: 10.1520/STP159820160077

Gongyao Wang,1 Kimberly Maciejewski,2 and Mark James3

Application of Kresidual
Measurements to Fracture
Toughness Evaluations
Citation
Wang, G., Maciejewski, K., and James, M., “Application of Kresidual Measurements to Fracture
Toughness Evaluations,” Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis,
ASTM STP1598, Z. Wei, K. Nikbin, P. C. McKeighan, and D. G. Harlow, Eds., ASTM International,
West Conshohocken, PA, 2017, pp. 119–132, http://dx.doi.org/10.1520/STP1598201600774

ABSTRACT
Residual stresses are self-equilibrating internal stresses that exist in materials
and structures due to a number of factors including thermal-mechanical
processing, machining, forming, or welding. These stresses are of critical
importance, particularly in structural materials for fatigue-crack growth and
fracture design, where residual stresses can bias the material property
measurements, leading to either unconservative or overconservative design
data. The driving force due to residual stress is referred to as the residual stress-
intensity factor (Kresidual) and can be used to understand and prevent bias during
material characterization for material expected to contain residual stresses. The
cut-compliance method can be employed to measure the Kresidual based on
Schindler’s method, where new crack surface area is created by electrical
discharge machining (EDM). However, the cut-compliance test is destructive and,
as a result, residual stress effects on a property such as fracture toughness would
require two tests, one for residual stress and one for fracture toughness, along
with an assumption that the residual stress is the same for both measurements.

Manuscript received March 21, 2016; accepted for publication September 15, 2016.
1
Arconic Technology Center, 100 Technical Dr., New Kensington, PA 15069 http://orcid.org/
0000-0002-2427-5389
2
Arconic Technology Center, 100 Technical Dr., New Kensington, PA 15069 http://orcid.org/
0000-0002-2880-9843
3
Arconic Technology Center, 100 Technical Dr., New Kensington, PA 15069 http://orcid.org/
0000-0002-1816-8712
4
ASTM Symposium on Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis on May 4–5,
2016 in Grand Hyatt, San Antonio, TX.

Copyright V
C 2017 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.
120 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

The new crack-compliance method is similar to the cut-compliance method, but


the new surface area is created through fatigue-crack growth due to cyclic loading.
The advantage of the crack-compliance method is to produce Kresidual data on the
same specimen as the property data. This crack-compliance method is based on
the same standard compliance data used to calculate crack length, with a simple
extension of the method for calculating the Kresidual value. The method has been
demonstrated in the past with good success for fatigue-crack growth, and this
paper will extend the concept to fracture toughness. The approach is to apply the
crack-compliance method during the precrack phase of a fracture toughness test
to estimate Kresidual at the end of the precrack. Then a simple superposition
model is used to correct measured fracture toughness.

Keywords
Kresidual, crack compliance method, fatigue crack growth (FCG), fracture toughness

Introduction
Residual stresses are self-equilibrating internal stresses that are locked in materials
and structures even without external loading. Residual stresses can have a number
of causes related to manufacturing and fabricating processes and can be an un-
avoidable result of manufacturing for some materials, such as die-forged aluminum
alloys, where cold work minimizes—but does not completely eliminate—residual
stress in the material. These stresses are of critical importance, particularly in struc-
tural materials for fatigue crack growth (FCG) and fracture design, where residual
stresses can bias the material property measurements, leading to either unconserva-
tive or overconservative design data.
Fracture mechanics-related material properties (FCG and fracture toughness) are
important for material selections for lightweight aerospace structural applications [1,2].
Residual stress influences the driving force for fracture and constraint conditions at the
crack tip. Residual stress can cause an entire crack front of macroscopic dimensions to
be in a state of tension or compression before the application of any external load. In
general, tension residual stress will increase the crack opening and compression residual
stress will decrease the crack opening. For fracture toughness, these factors will result in
the change of the fracture load of the specimens during fracture-toughness testing.
Therefore, residual stress present in specimens for fracture-toughness testing will affect
the experimental determination of fracture toughness [3,4]. The driving force due to re-
sidual stress is referred to as the residual stress-intensity factor (Kresidual) and can be
used to understand and prevent bias during material characterization for material
expected to contain residual stresses.
The most common residual stress measurement methods, such as hole drilling,
ring core, cut-compliance, layer removal, and sectioning, are destructive proce-
dures [5]. The two primary nondestructive techniques, X-ray diffraction (XRD)
and neutron diffraction (ND), have several significant limitations. XRD can only
WANG ET AL., DOI 10.1520/STP159820160077 121

measure residual stresses on the surface or near surface in crystalline alloys. ND


can measure residual stress to depths of many centimetres [5], but it requires a
large and expensive neutron source and is usually used for research purposes.
Moreover, a typical stress depth profile obtained using high energy XRD or ND
methods may take days to a week with significant lead time, planning, and expense
often required. Destructive methods such as cut-compliance [6,7] can be complet-
ed in less than a day, but the destructive nature of the method limits its applicabili-
ty for fracture mechanics property determination. Due to these limitations, it is
difficult to evaluate the effects of the residual stresses on the mechanical properties,
especially FCG and fracture properties, because it is difficult to obtain residual
stress and fatigue and fracture data from the same specimen.
Recently, a new method was developed to make it possible to partition residual
stress effects from FCG rate (FCGR) data [8]. The online crack-compliance method
was developed to measure the stress-intensity factor due to residual stress in a coupon
in real time during an FCG test [8] and has recently been applied for FCGR characteri-
zation [2]. This method is similar to the cut-compliance method (also sometimes
referred to as the slitting method), but new surface area is created by FCG during cyclic
loading rather than electrical discharge machining (EDM) slitting. A principal advan-
tage for the online crack-compliance method is that the FCGR data and Kresidual data
will be obtained from the same specimen. Thus, the effects of residual stresses can be
evaluated to produce property data unbiased by residual stress.
The main objective of this paper is to develop a method to correct for residual stress
that lies in fracture-toughness data, with primary application to plane-strain fracture
toughness. The approach is to use the online crack-compliance method during the fa-
tigue precracking procedure to estimate Kresidual at the end of the precrack (i.e., for the
start of the fracture-toughness test). Then the single value of measured Kresidual is used in
a simple linear superposition model to correct for the residual stress bias. In this paper,
the online crack-compliance method was used to compare with the results that were
obtained from the cut-compliance method. The Kresidual obtained from the online crack-
compliance method was employed to correct the fracture-toughness results, which
showed the bias due to the different residual stress present in the tested specimens.

Crack-Compliance Method
The crack-compliance method was developed by Lados, Apelian, and Donald [8]
using linear elastic fracture mechanics (LEFM) concepts to calculate the stress-
intensity factor due to residual stress (Kresidual) at the crack tip and to extend the
cut-compliance method developed by Schindler [6,7] for application to fatigue
cracks. A schematic of a load versus displacement curve that represents the compli-
ance change during crack growth is shown in Fig. 1. In order to simplify the calcula-
tions for demonstration purposes, nonlinear effects (i.e., crack closure) are
neglected. If there is no residual stress in the specimen, the compliance increases
with crack growth and the displacement always returns to the same value at zero
122 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 1 Schematic of load-displacement data used for the crack-compliance method to


determine Kresidual [9] (tensile residual stress case shown).

Load, P Ddmax Ddresidual

Pmax

Previous
Crack
Length New Crack Length
(With Residual Stress)

New Crack Length


Pmin (No Residual Stress)

P0
Crack Opening Displacement, d
Ddresidual

load, as demonstrated in Fig. 1. When residual stress does exist in the specimen, the
body load is redistributed due to residual stress as the crack grows. This redistribu-
tion causes a shift in the displacement at zero load. Thus, the crack compliance method
applies an incremental approach that uses the maximum displacement, Ddmax,
displacement at zero load, Ddresidual, and the maximum applied stress-intensity factor,
Kmax, to calculate Kresidual, as shown in Eq 1 [8,9]:
 
Ddresidual
Kresidual ¼ Kmax : (1)
Ddmax

In order to obtain an accurate calculation of Kresidual, it is necessary to have very


stable load-displacement signals for the crack-compliance method. Acceptable slope vari-
ation may affect crack length slightly but can affect the intercept significantly. The Kresidual
equation for the crack-compliance method is surprisingly simple and easy to use without
the need of influence functions necessary to apply the cut-compliance method.

Experimental Approaches
Two materials (2024-T351 and 7050-T7651 plate) were employed in this work. The
2024-T351 plate (25.4 mm thick and 1219.2 mm wide) was friction stir welded
(FSW) to induce residual stresses and then machined into standard T-L orientation
(loading in the material transverse direction and crack growth in the material long
[rolling] direction) compact tension, C(T), FCG specimens (width, W ¼ 101.6 mm,
and thickness, B ¼ 6.35 mm) [10]. Specimens excised with the friction stir region at
WANG ET AL., DOI 10.1520/STP159820160077 123

the back edge of the specimen (Fig. 2a) had a tensile-dominated Kresidual, while
specimens excised with the friction stir region inboard of the back face (Fig. 2b) had
a compression-dominated Kresidual. Fig. 2c shows the standard C(T) specimen with a
machined notch. The 7050-T7651 thick plate (149.86 mm thick and 1308.1 mm
wide) was machined to standard L-S oriented (loading in the material long [rolling]
direction and crack growth in the material short transverse direction) C(T) FCG speci-
mens without the notch (W ¼ 101.6 mm and B ¼ 6.35 mm), as shown in Fig. 2d. The
L-S orientation positions the specimen for crack growth through the thickness of the
plate material, using the very low—but measurable—residual stress in the plate.

FIG. 2 (a) C(T) FCG specimen configuration with dominated tensile residual stress; (b)
C(T) FCG specimen configuration with dominated compressive residual stress;
(c) C(T) FCG specimen; (d) C(T) FCG specimen without machining notch.

(a) (b)

(c) (d)
124 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

Residual stress in these specimens was due to the thermal-mechanical processing of the
plate during normal production. This series of specimens was designed to examine the
influence of pinholes on the residual stress distribution and on the transition from cut-
compliance to crack-compliance subsequent testing.
Constant load amplitude (DP) and constant DK FCG tests were performed on
computer-controlled servohydraulic testing machines using a Fracture Technology
Associates (FTA) control system [11]. A crack-opening displacement (COD) gage
was used to monitor the crack length using the compliance method. FCG tests were
carried out at a frequency of 10 Hz and load ratios, R of 0.5 and 0.1, were employed to
compressive residual stress-dominated specimens and tension residual stress-dominated
specimens, respectively. In addition, selected specimens were also subjected to
precracking followed by fracture-toughness testing. Details of specimen identification
(ID), loading conditions, and testing procedures are outlined in Table 1.
For the cut-compliance method, a strain gage was glued on the back face of the
C(T) specimen. An EDM machine was employed to do incremental cutting. Strain
readings were taken by a strain indicator and the Schindler equation was used to
calculate the Kresidual from the measured strains [7].

TABLE 1 Specimen IDs, residual stress conditions, and testing procedures.

Specimen
Alloy ID Specimen Condition Testing Procedure

2024-T351 T4 With tensile-dominated Cut-compliance measurement


field
C4 With Cut-compliance measurement
compressive-dominated
field
T15 With tensile-dominated Precracking to a ¼ 30.48 mm and then running
field FCG test to a ¼ 50.8 mm at constant DK ¼ 11
MPam0.5 and R ¼ 0.1, cut-compliance to
the edge
C15 With Precracking to a ¼ 30.48 mm and then running
compressive-dominated FCG test to a ¼ 50.8 mm at constant DP ¼ 6.9
field kN and R ¼ 0.5
T16 With Precracking to a ¼ 40.64 mm at constant
tensile-dominated field DK ¼ 8.2 MPam0.5 and R ¼ 0.1 and then fracture
toughness test
C16 With Precracking to a ¼ 40.64 mm at constant
compressive-dominated DP ¼ 6.9 kN and R ¼ 0.5 and then fracture
field toughness test
7050-T7651 L-S-3 With pinholes Cut compliance to a ¼ 20.32 mm and then FCG
test to a ¼ 88.9 mm at constant DK ¼ 11
MPam0.5 and R ¼ 0.1
L-S-5 With pinholes Cut-compliance measurement
L-S-20 Without pinholes Cut-compliance measurement
WANG ET AL., DOI 10.1520/STP159820160077 125

For the crack-compliance method, standard C(T) specimens were machined. In


general, a constant applied DK was used to precrack the specimens following
ASTM E647, Standard Test Method for Measurement of Fatigue Crack Growth
Rates, and ASTM E399, Standard Test Method for Linear-Elastic Plane-Strain Frac-
ture Toughness Klc of Metallic Materials. However, for some specimens with high
compressive residual stress, a constant load amplitude to precrack was used, where
the increasing K for loading helped offset the decreasing K due to residual stress.
FTA software was used to obtain Kresidual data [11]. The force-shed procedure of
precracking outlined in ASTM E399 was not investigated.

Results and Discussion


Focusing on alloy 2024-T351 first, two residual stress fields (tensile- and compressive-
dominated) were considered, and two test methods (cut- and crack-compliance) were
used in conjunction with various test procedures for FCG and fracture-toughness
testing, as detailed in Table 1. The cut-compliance method was employed on Specimens
T1, T4, and C4, with tensile- and compressive-dominated residual stress fields, first to
obtain baseline Kresidual profiles, as shown in Fig. 3. These curves demonstrate that there
is a distinct difference in Kresidual profiles along the crack-growth direction for the
compressive versus tensile residual stress state. The difference between T1 and T4 is
likely due to the slower transient in the FSW process at the edge of the plate.

FIG. 3 Kresidual versus crack length curves from cut-compliance and crack-compliance
methods.
126 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

The next step is to compare data generated from FCG with the cut-compliance
data. Specimens T15 and C15, with tensile- and compressive-dominated residual
stress fields, were precracked and subjected to FCG testing. Specimen T15 received
a two-step test procedure: This specimen was fatigue precracked for a prescribed
distance. Then T15 was removed from the fatigue test frame and introduced to the
EDM machine for slitting from the end of the fatigue crack to the edge of the
specimen to obtain Kresidual data using the cut-compliance method. Fig. 3 shows
that the measured Kresidual data, as a function of crack length obtained from the
crack-compliance and cut-compliance methods for Specimen T15, are consistent
with each other and with the baseline Kresidual profile for Specimen T4. This shows
that redistribution effects associated with interrupting the fatigue test and moving
to the EDM procedure are correctly captured by the incremental analysis
approaches for the crack-compliance method and Schindler’s method.
Similarly, for Specimen C15, Kresidual from the crack-compliance method was
plotted in Fig. 3 and exhibits a profile consistent with the baseline Kresidual profile
for Specimen C4. Additionally, Kresidual profiles for the precracking sequence of
Specimens T16 (with a tensile-dominated residual stress field) and C16 (with a
compressive-dominated residual stress field) are shown in Fig. 3. It seems that
Specimen C16 exhibited similar Kresidual profiles to Specimens C4 and C15, al-
though Specimen T16 had similar Kresidual profiles to Specimen T1 and slightly
small Kresidual compared to T4 and T15.
After Specimens T16 and C16 were precracked, fracture-toughness testing was
conducted on both specimens, and applied KQ was obtained, as shown in Fig. 4.
The applied KQ is measured following the standard fracture-toughness analysis

FIG. 4 The applied and corrected KQ for tensile- and compressive-dominated residual
stress fields, in Specimens T16 and C16, respectively.

Difference: 12.75 (34 %)


60.00 Difference: 2.83 (6 %)
50.66
50.00 45.57
42.74
40.00 37.91
KQ (MPa.m0.5)

T16
C16
30.00

20.00

10.00

0.00
Applied Corrected
WANG ET AL., DOI 10.1520/STP159820160077 127

using ASTM E399. The focus here is on the precracking procedure and application
as a correction procedure (the authors acknowledge that the specimens are too thin
to produce valid plane-strain toughness values). The difference in applied KQ be-
tween T16 and C16 is about 12.75 MPam0.5 (34 %). This large variance is because
the local K at the crack tip in T16 is very different from C16. For T16, the Kresidual
at the crack tip is about 7.66 MPam0.5, which was obtained from the crack-
compliance method during precracking using FTA software. However, for C16, the
Kresidual at the crack tip is about 7.91 MPam0.5. If we consider the Kresidual effects,
the corrected K is equal to (KQ þ Kresidual), which is 45.57 MPam0.5 at the precrack-
ing tip for T16, while corrected K is 42.74 MPam0.5 at the precracking tip for C16,
as shown in Fig. 4. Thus, the difference of corrected K at the precracking tip is about
2.82 MPam0.5 (6 %). These results show that the same material with differing
amounts of residual stress (T16 versus C16) exhibit a bias in fracture-toughness
measurements. The variation in corrected K is minimized by considering the influ-
ence of Kresidual on the applied KQ.
Turning attention to alloy 7050-T7651, cut-compliance measurements were
conducted on Specimens L-S-5 and L-S-20, which are specimens with and without
pinholes, respectively, and their Kresidual profiles are shown in Fig. 5. These results
show that pinholes caused the residual stress redistribution and, thus, Kresidual
profiles changed. As a result, subsequent testing was performed only on specimens
with holes machined.

FIG. 5 Kresidual versus crack length for 7050-T7651 alloy obtained from cut-compliance
method.
128 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

Next, cut-compliance measurements were performed on Specimen L-S-3 with


pinholes and without the machined notch, as shown in Fig. 6. The cut-compliance
measurement was stopped at a designated notch length, and the specimen was
moved to a fatigue-testing machine where a crack-compliance Kresidual measure-
ment was performed during subsequent FCG testing, as shown in Fig. 6. For the
7050-T7651 alloy, the residual stresses are relatively small and, thus, the resultant
Kresidual is small (Fig. 6), which results in small displacement at zero load, Ddresidual.
Therefore, the difficulty in measuring small displacement at zero load could cause
the large data scattering seen with the crack-compliance method. In general, the
comparison between the L-S-5 specimen (cut-compliance for the full width of the
specimen) and L-S-3 is good, and the small variation in Kresidual (less than 1 MPam0.5)
between Specimens L-S-3 and L-S-5 is believed to be due to the natural variation in the
residual stress field. Focusing on the L-S-3 specimen results, the transition from the
cut-compliance to the crack-compliance procedure shows generally good agreement,
demonstrating that the crack-compliance method can be used directly from a
premachined notch in a specimen. This was believed to be the case because the
formulation of the crack-compliance method is incremental and relies on the previ-
ous history of the strain and stress field only through the numerical derivative cal-
culation in Eq 1. In this way, the calculation of Kresidual for a given crack length is
dependent on the choice of increment size for the numerical derivative and is not
dependent on the full previous residual stress history.

FIG. 6 Kresidual versus crack length for 7050-T7651 alloy obtained from cut-compliance
and crack-compliance methods.
WANG ET AL., DOI 10.1520/STP159820160077 129

In order to reduce the data scatter for the crack-compliance result for Specimen
L-S-3, a rolling average smoothing (RA factor) was performed with a window size
of 10 data points for the Kresidual calculation, which was applied to Specimen L-S-3
data analysis, as shown in Fig. 7. The area circled in Fig. 7, at the transition from
cut-compliance to crack-compliance, highlights an artifact of the smoothing meth-
od applied to this data set. At the start of the crack-compliance test, a new crack
must form at the EDM notch, and some amount of crack growth is needed to grow
away from the stress concentration of the notch and to stabilize the plastic wake
behind the growing crack tip. A review of the FCGR data collected during the test
indicated a transient in the FCGR, as well as in the Kresidual, at the start of the FCG
portion of the test. The FCGR and Kresidual stabilized after about 1 mm of crack
growth from the EDM notch, which corresponds to the first five data points. Such a
transient was not noticeable in the previous 2024-T351 data set—for instance, for
Specimen T15, which also had a similar transition between test methods. However,
the magnitude of Kresidual was larger and, as a result, a small transient such as the
one for L-S-3 might not be noticeable for T15.
Other transients in data from other testing efforts not reported on in detail here
have been observed where high-tensile stresses or compressive stresses require differ-
ing amounts of crack growth to stabilize FCG rates and closure measurements and, as
a result, a relatively conservative procedure was developed for measuring the Kresidual
during precracking. The procedure combines the EDM procedure to introduce the

FIG. 7 Kresidual versus crack length for 7050-T7651 alloy obtained from cut-compliance
and crack-compliance method (RA factor of 10 for the Kresidual calculation was
applied).
130 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

notch, with Kresidual measurement during the EDM using the cut-compliance method.
Then the crack-compliance method is used for precracking, with a generous amount
of crack growth required for the precrack. The EDM cut-compliance data can be used
as an indicator of high-tensile or compressive residual stress effects to adjust loads in
the precracking strategy. For instance, if significant compressive residual stress effects
are present, precracking at the normal loading level may require excessive loading
cycles, and higher precracking loads may be warranted.
Based on the results of this work, the following procedure is recommended:
1. C(T) specimens are machined with holes but no notch.
2. EDM cut-compliance is used to introduce the notch and to measure the Kresidual
to a notch length: a/W ¼ (0.5 – 6.35/W) (with W having units of mm).
3. Precracking is used to extend the notch by Da ¼ 6.35 mm to a total crack
length: a/W ¼ 0.5. Kresidual data is measured during precracking and the value
at the end of the precrack at a/W ¼ 0.5 is used in subsequent data corrections.
Precracking loads can be adjusted based on indicators from the EDM cut-
compliance data.
4. The ASTM E399 fracture test is performed as normal to calculate the uncor-
rected material KQ value.
5. The precracking FCG data should be reviewed for consistency. For instance,
the FCGR, closure, and Kresidual data should be plotted as a function of crack
length to evaluate initial transients for growth from the notch for consistency.
6. The cut-compliance and crack-compliance data should be plotted together
similar to Fig. 6 to evaluate the two data sets for consistency.
7. It is likely that some form of averaging will be necessary for the Kresidual data
points approaching the end of the precrack. Judgment and careful review of
the precracking data will be required until a robust approach can be recom-
mended to calculate the precracking Kresidual.
8. Calculate the corrected fracture toughness, KQrs ¼ KQ þ Kresidual.

After adequate data are collected over time, the need for the cut-compliance
test as verification for the crack-compliance test should be reevaluated.
There is not enough experience with this method to fully recommend a set of
validity requirements and to propose a fully qualified fracture toughness that is
residual-stress corrected. However, our current thinking is that generally any of the
validity requirements that depend on the actual stress-intensity factor at the crack
tip should be dependent on the corrected value of stress-intensity factor, KQrs.
Further work remains to fully develop the method before standardization can
be considered in ASTM E399.

Summary and Conclusions


A new approach was developed to measure the effects of residual stress on fracture-
toughness data based on the crack-compliance method of measuring the Kresidual
stress-intensity factor during the precracking phase of a fracture-toughness test.
The results were verified by performing a cut-compliance test using EDM to
WANG ET AL., DOI 10.1520/STP159820160077 131

introduce the notch into the specimen, which provides a direct comparison of
Kresidual at the transition from the EDM notch to the fatigue precrack. The Kresidual
measurements from the cut-compliance and crack-compliance methods were found
to be consistent. For two tests that had significant residual stress intentionally
introduced using friction stir welding, the linear superposition approach was effec-
tive in significantly reducing the scatter in the fracture-toughness results. Finally, a
procedure was outlined that combines the cut-compliance and crack-compliance
methods to provide verification of the Kresidual results while experience is developed
in the approach. Further work remains to fully develop the method before stan-
dardization can be considered in ASTM E399.

ACKNOWLEDGMENTS
The authors would like to thank Richard L. Brazill, Donald W. Millard, Ronald M.
Marcozzi, Kevin C. Honeck, and William Szekely for their help completing this work.

References

[1] Donne, C. D., Biallas, G., Ghidini, T., and Raimbeaux, G., “Effect of Weld Imperfections
and Residual Stresses on the Fatigue Crack Propagation in Friction Stir Welded Joints,”
presented at the Second International Symposium on Friction Stir Welding, Gothenburg,
Sweden, June 26–28, 2000.

[2] James, M., Maciejewski, K., Wang, G., Ball, D., and Bucci, R., “A Methodology for Parti-
tioning Residual Stress Effects From Fatigue Crack Growth Rate Test Data,” Materials
Performance and Characterization, Vol. 5, No. 3, 2016, pp. 119–132, http://dx.doi.org/
10.1520/MPC20150062

[3] Hill, M. R. and Panontin, T. L., “Effect of Residual Stress on Brittle Fracture Testing,” Fatigue
and Fracture Mechanics, ASTM STP1332, T. L. Panontin and S. D. Sheppard, Eds., ASTM
International, West Conshohocken, PA, 1998, http://dx.doi.org/10.1520/STP1332-EB

[4] Hill, M. R. and VanDalen, J. E., “Evaluation of Residual Stress Corrections to Fracture
Toughness Values,” J ASTM Intl, Vol. 5, No. 8, pp. 1–11, http://dx.doi.org/10.1520/
JAI101713

[5] Prime, M. B., “Residual Stress Measurement by Successive Extension of a Slot: The Crack
Compliance Method,” Appl Mech Rev, Vol. 52, No. 2, 1999, pp. 75–96.

[6] Schindler, H. J., “Determination of Residual Stress Distributions from Measured Stress In-
tensity Factor,” Int J Fracture, Vol. 74, No. 2, 1995, pp. R23–R30.

[7] Schindler, H. J., Cheng, W., and Finnie, I., “Experimental Determination of Stress Intensity
Factors Due to Residual Stresses,” Exp Mech, Vol. 37, No. 3, 1997, pp. 272–277.

[8] Lados, D. A., Apelian, D., and Donald, J. K., “Fracture Mechanics Analysis for Residual
Stress and Crack Closure Corrections,” Int J Fatigue, Vol. 29, No. 4, 2007, pp. 687–694.

[9] Donald, J. K. and Lados, D. A., “An Integrated Methodology for Separating Closure and
Residual Stress Effects from Fatigue Crack Growth Rate Data,” Fatigue & Fracture of
132 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

Engineering Materials & Structures, Vol. 30, 2007, pp. 223–230, http://dx.doi.org/10.1111/
j.1460-2695.2006.01081.x.

[10] Newman, J. A., Smith, S. W., Seshadri, B. R., James, M. A., Brazill, R. L., Schultz, R. W.,
Donald, J. K., and Blair, A., “Characterization of Residual Stress Effects on Fatigue Crack
Growth of a Friction Stir Welded Aluminum Alloy,” NASA/TM–2015–218685, NASA Lang-
ley Research Center, Hampton, VA, 2015.

[11] Donald, J. K. and Blair, A., Automated FCG Testing and Analysis—Series 2001, Version
3.09, Fracture Technology Associates, LLC, Bethlehem, PA, 2009.
FATIGUE AND FRACTURE TEST PLANNING, TEST DATA ACQUISITIONS AND ANALYSIS 133

STP 1598, 2017 / available online at www.astm.org / doi: 10.1520/STP159820160080

Stephen M. Graham1

Sensitivity Study on Parameters


that Influence Automated Slope
Determination
Citation
Graham, S. M., “Sensitivity Study on Parameters that Influence Automated Slope
Determination,” Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis,
ASTM STP1598, Z. Wei, K. Nikbin, P. C. McKeighan, and D. G. Harlow, Eds., ASTM International,
West Conshohocken, PA, 2017, pp. 133–150, http://dx.doi.org/10.1520/STP1598201600802

ABSTRACT
There are numerous ASTM standard test methods where force and displacement
are recorded and the data analysis requires that the slope of the force-
displacement record be determined. Demands for greater accuracy and the
availability of computers have led to the widespread use of simple linear
regression. However, computers are not good at determining what data to
include in the regression, so the analyst must manually select the upper and
lower limits of the regression region, thereby introducing subjectivity into the
analysis. Fixed fit ranges that are often used for linear regression can lead to
slope bias for data sets that exhibit curvature within the fixed range. This is
particularly true for data sets that have an initial curvature or that have a small
linear region. Two approaches that provide a powerful tool for examining a data
set to determine the linear region are reduced displacement and analysis of
residuals. The latter was incorporated into a fully automated algorithm for slope
determination by analysis of residuals. This study looked at how noise, digital
resolution, and sampling rate influence the determination of slope using this
algorithm. Twelve synthetically generated data sets were analyzed to provide
insight into how each of these data sets’ characteristics affected the resulting
slope. It was determined that slope error from linear regression is a complex
interaction between the shape of the data in the nonlinear regions and the data

Manuscript received March 22, 2016; accepted for publication August 5, 2016.
1
United States Naval Academy, Mechanical Engineering Dept., 590 Holloway Rd., Annapolis, MD 21402
http://orcid.org/0000-0002-4971-1271
2
ASTM Symposium on Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis on May 4–5,
2016 in Grand Hyatt, San Antonio, TX.

Copyright V
C 2017 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.
134 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

set characteristics. Noise has more effect on slope error than digital resolution
over the ranges considered. The algorithm proved robust in that, even with
typical noise and digital resolution, slope error in data sets with small linear
regions was less than about 2 %.

Keywords
linear regression, mechanical testing, analysis of residuals

Introduction
There are numerous ASTM standard test methods where force and displacement
are recorded and the data analysis requires that the slope of the force-displacement
record be determined [1–7]. What these test methods have in common is that the
specimen response is inherently linear over some region. There are other test meth-
ods where the response of a system or process being measured is inherently linear,
so there are many situations where the analyst is faced with the challenge of deter-
mining the slope of the linear region. Historically, the method for determining the
slope has not been explicitly described and, therefore, is left up to the analyst. In the
days before computerized data analysis, data records were created using pen plot-
ters and slopes were determined by laying a straight edge on the plot. There is some
subjectivity in visual determination of slope, but the human eye is very adept at de-
termining the linear region and fitting a line through the center of the plotter pen
trace. Demands for greater accuracy and the availability of computers have led to
the widespread use of simple linear regression. However, computers are not good at
determining what data to include in the regression, so the analyst must manually
select the upper and lower limits of the regression region, thereby introducing sub-
jectivity into the analysis. The objective of a standardized test method is that multi-
ple labs performing the same analysis should produce the same result. However,
round-robin test programs to develop precision statements reveal that this is a de-
ceptively difficult task [8].
Sometimes fixed ranges have been used for automated data analysis programs,
such as 20 % to 80 % of the maximum force. This can work when the specimen be-
havior is well-defined and consistent but does not work when behavior deviates
from what is expected. One example of this was an analytical round-robin con-
ducted by the ASTM task group for E399, Standard Test Method for Linear-Elastic
Plane-Strain Fracture Toughness KIc of Metallic Materials [4]. Fifteen fracture
toughness test records from three different materials were given to nine participants
with instructions to use ASTM E399 to determine PQ, which requires determining
the slope of the test record [9]. The results are presented in terms of the z-score in
Fig. 1. The z-score is the number of standard deviations a value is from the mean.
For each data set, a mean and standard deviation for PQ were calculated based on
results from all participants. If the slope used to determine PQ is high, it will result
in a low PQ value, and conversely if it is low, it will result in a high PQ. Participant
GRAHAM, DOI 10.1520/STP159820160080 135

FIG. 1 Results from ASTM E399 analytical round-robin to determine PQ for 15 data sets.

1
z-factor

-1

-2

-3
A B C D E F G H I
Analysis Participant ID

E used a fixed range of 10 % to 60 % of Pmax to determine PQ, while Participant H


used a fixed 20 % to 80 %. Many of the test records exhibited an initial region with
a decreasing slope leading into the nominal linear region. In all cases, the slope
decreased after the nominal linear region. Consequently, biasing the fit range to the
lower end results in a higher slope and a lower PQ, as can be seen in the four points
for Participant E that fall below a z-score of 1. On the other hand, 80 % is in the
upper decreasing slope region for many test records, so the results for Participant
H tend to be high with five points above a z-score of þ1 and all but one above 0.
For Participant C, the fit range was determined by manually examining each data
set. The upper end is around 50 % of Pmax for most tests; the lower end varies from
1 to 29 % but is generally around 10 to 20 % of Pmax, which explains the why Partic-
ipant C does not show the low values that Participant E does. Participant F utilized
a fully automated algorithm referred to as slope determination by analysis of resid-
uals (SDAR, to be explained later) to find the optimum fit range and obtained val-
ues that are within 61 z-score with only one exception.
Use of fixed ranges can lead to inaccurate results when the analyst accepts the pro-
gram output without checking a graph of the regression line superimposed on the
data record, as was the case for another round-robin conducted for ASTM E1290,
Standard Test Method for Crack-Tip Opening Displacement (CTOD) Fracture Tough-
ness Measurement [8]. Five labs were sent six identical steel specimens to be tested
according to ASTM E1290 at 196  C. The specimen response was expected to be very
linear at this temperature. Each lab conducted analysis of its own tests to determine
CTOD, which involves determining the slope in the initial part of the test record. All
tests were reanalyzed using a single method of slope determination to provide a bench-
mark. The percent difference between the lab and benchmark calculations of CTOD
136 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

(Lab1290/Benchmark1290 – 1) are shown in Fig. 2. Many of the tests that exhibited


large differences had an initial nonlinearity in the force-crack mouth opening displace-
ment (CMOD) record. Great care was taken in the benchmark to carefully select data
in the linear part of the record using analysis of residuals. Comparison of the lab anal-
ysis with the benchmark revealed that, for some labs, the initial nonlinearity was not
excluded from the range used in the lab analysis. In most cases, the slope started out
high and gradually decreased into the linear region. When this nonlinearity is not
accounted for in the data selection for regression, it results in a slope that is too high,
which results in a CTOD that is too high, as can be seen in Fig. 2.
Visual examination of the linear fit superimposed on the data is important as a
check on the regression analysis, but it is not sensitive enough to provide repeatable
results. Two approaches that have been used to reveal the linear region and devia-
tion from linearity are referred to as reduced displacement [10] and analysis of
residuals [11]. Reduced displacement looks at the horizontal distance (x-value) the
data points are from the regression line, and analysis of residuals looks at the verti-
cal distance (y-value). An example of reduced displacement is shown in Fig. 3, and
residuals for the same data set are shown in Fig. 4. This particular data set is from

FIG. 2 Comparison of lab analysis with benchmark from ASTM E1290 round-robin.
GRAHAM, DOI 10.1520/STP159820160080 137

FIG. 3 Example of reduced displacement from ASTM E399 round-robin, specimen M2


[10].

FIG. 4 Example of residuals from ASTM E399 round-robin, specimen M2.


138 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

an analytical round-robin for ASTM E399 and had a small linear region with a lot
of curvature. The power of these methods to reveal the character of the data relative
to the linear regression line is clear from comparison of the reduced displacement
and residuals plots with the load-displacement plot on the right in Fig. 3. The solid
line in the plot on the right is the linear fit, and the dashed line is the 95 % secant
used to determine PQ. These methods provide a very effective tool for examining
data and evaluating the quality of the linear regression fit.
In the ideal case, the specimen response is linear over a significant region, devi-
ation from linearity is abrupt, and the data acquisition system accurately captures
the data. Too often, the specimen response is less than ideal. The data may be linear
only over a small portion of the record and may deviate from linearity very gradual-
ly. Often there is an initial nonlinearity in the data before the linear region, or the
data do not have a true linear region. Assuming that the transducers have been
checked for linearity, deviations from theoretical linear behavior can result from in-
teraction between the transducer or load frame and the specimen due to interfaces
or fixtures or due to material behavior that is not ideal. An example of this was a
round-robin conducted by the ASTM task group for E9, Standard Test Methods of
Compression Testing of Metallic Materials at Room Temperature [12]. In this
round-robin, ten labs were sent seven identical specimens taken from a single alu-
minum plate and asked to conduct tests and analysis per ASTM E9 and report
values obtained for the compression modulus and yield strength. Calculating the
compression modulus requires determining slope of the test record. In this study,
there was a significant difference in the specimen response depending on whether
lubricant was used between the specimen and the loading platens and on whether
one or two displacement gages were used. Two examples of force (y) versus dis-
placement (x) data are shown in Fig. 5 and Fig. 6. In the first of these, the specimen
behavior is nonlinear over all but a small range. The slope starts out high and grad-
ually decreases through what would be expected to be the linear range before
decreasing more rapidly as compressive yielding occurs. Contrast this with Fig. 6
where the specimen response is linear up to yielding. The difference between these
two specimens is revealed more clearly by plotting the residuals, as shown in Fig. 7.
In this graph, the residuals are calculated after normalizing the data, as will be dis-
cussed in the next section. Specimens LT-55 and LT-85 had a very linear response
from the start of the test, so the residuals are near zero and gradually increase as
yielding occurs. The curve for LT-41 starts out with a higher slope than the regres-
sion line, so the data points are to the right of the line and the residuals are negative.
The curve for LT-63 starts out with a lower slope than the regression line, so the
data points are to the left and the residuals are positive. The residuals plots clearly
show the small linear regions for the latter two specimens. In this round-robin,
each lab conducted their own tests and analyzed their data to obtain the compres-
sion modulus. In an effort to separate the effects of specimen behavior from analysis
method on modulus determination, all tests were reanalyzed using a single auto-
mated method (SDAR) as a benchmark. Criteria were applied to evaluate the
GRAHAM, DOI 10.1520/STP159820160080 139

FIG. 5 Force-displacement record from ASTM E9 test LT-41 that exhibited significant
nonlinearity.

quality of the data and the resulting fit. One of those criteria was to compare the fit
range to the full range up to the “knee” in the curve. Good test records were deemed
to be those with a fit range of at least 40 % of the full range up to the “knee.” The
percent difference in compression modulus (Lab E/Overall Avg E – 1) is plotted
versus the relative fit range in Fig. 8, where the relative fit range is the fit range di-
vided by 0.4 such that a good fit has a relative fit range of greater than 1. Data that
passed all the quality checks in the SDAR algorithm (to be explained in the next
section) are labeled as “Pass” in the caption, whereas those that failed one or more
quality checks are labeled as “FAIL.” There is a clear trend for the percent difference
in modulus to decrease as the relative fit range increases. This is true for both the
lab analysis and the benchmark analysis, so nonlinearity in data creates significant
challenges to identifying the linear region and determining slope. When there is no
true linear region, the slope determination algorithm should alert the analyst that
the resulting slope should be examined carefully to determine validity.
140 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 6 Force-displacement record from ASTM E9 test LT-55 with large linear region.

The identification of the linear region is further complicated by noise, digital reso-
lution, and sampling by the signal conditioning and data acquisition system. In the
extreme, these factors can obscure the underlying specimen behavior. So another desir-
able trait for an automated data analysis program would be to determine if noise, digi-
tal resolution, or sampling rate could negatively influence the slope determination.

Description of the SDAR Algorithm


Graham and Adler developed an automated algorithm, SDAR [11], that can analyze
a test record and determine an accurate, repeatable slope without the need for sub-
jective operator input. Results from an analytical round-robin revealed that this al-
gorithm can be implemented in various coding languages by different analysts and
obtain the same results for a variety of data sets.3

3
Based on results from round-robin conducted by ASTM Task Group E08.03.05 in May 2016.
GRAHAM, DOI 10.1520/STP159820160080 141

FIG. 7 Residuals plots for selected tests from ASTM E9 round-robin showing deviation
from linearity.

FIG. 8 Comparison of lab analysis results for ASTM E9 round-robin with a benchmark
analysis showing the effect of nonlinearity on error in slope determination.

25
Percent Difference Between Lab and Overall Average

20

15

10

5
Modulus

-5

-10
LAB - Pass
-15 Benchmark - Pass
-20 LAB - FAIL
Overall Average E = 75.0 GPa Benchmark - FAIL
-25
0.20 0.40 0.60 0.80 1.00 1.20 1.40 1.60
Relative Fit Range
142 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

The SDAR algorithm provides a fully automated process for slope determination
and also provides metrics for evaluating the quality of the data and the quality of the
resulting fit. Criteria are included to evaluate the level of noise in the data and to deter-
mine if the resolution of the data acquisition system is adequate. The quality of the fit is
evaluated using analysis of residuals to examine the shape of the curve in the vicinity of
the fit window and the linear range. If the data set passes all of the quality checks, the pro-
cess is fully automated and no analyst intervention is required. If it fails one or more
quality checks, that does not mean that the resulting slope is wrong, but that the analyst
should examine the data and the fit carefully to determine if the slope is acceptable.
SDAR is intended for analyzing data records that are nominally monotonically
increasing in both x and y, and the early part of the record should have a region
where y is approximately proportional to x. The following description is provided of
a slightly modified version of the SDAR algorithm that was developed for this study.
In the first step of the analysis, the “knee” in the curve is found using the approach
recommended by Scibetta and Schuurmans [13]. Only data below the point of tan-
gency at the “knee” are used in finding the optimum fit. The truncated data set is
then normalized by dividing the y-variable by the tangency point y-value and the
x-variable by the tangency point x-value. The resulting data record then has a range
from 0 to 1 in both x and y. The data are normalized to facilitate the development of
fit criteria that are independent of the maximum values and units. Next, the noise
and digital resolution are checked, which are the metrics on quality of the data.
The method used to extract the noise from a data set is as follows.
1. Determine the number of points in the data set, N.
2. Calculate Dx ¼ xiþ1  xi for i ¼ 1 to N1.
3. Calculate the average of the Dx values, Dx.
4. Calculate the standard deviation of the ðDx  DxÞ values; this is the normalized
standard deviation of the noise. Multiply by the tangency point x-value to
unnormalize.
This process is repeated for the normalized y-values.
If the digital resolution is not sufficient, the data will exhibit highly discretized
behavior (steps). For a twelve-bit data acquisition system where the maximum value
is right at full scale, the normalized digital resolution, d, would be the maximum
value divided by the product of 212 times the value at the tangency point, 212yymax tangent
or
xmax
212 xtangent . The latter is required to normalize the resolution in the same way the data
are normalized. This is taken as the optimum resolution for the quality check. The
digital resolution of the data set is estimated as follows:
1. Calculate DDx ¼ jDxiþ1  Dxi j. This step removes the part of Dx that is due
to the steadily increasing x resulting from the test itself.
2. Create a histogram of DDx using bin start-points of d 2 þ j  d and bin end-
points of d2 þ j  d where j starts at 0.
3. The histogram will exhibit spikes at the multiples of the digital resolution. The
digital resolution is the average of the differences between the midpoints of
bins corresponding to the spikes.
GRAHAM, DOI 10.1520/STP159820160080 143

This process is also performed for the normalized y-variable. The next step is to
search the data set for the most linear region using an optimum region (OR) search
algorithm developed by Scibetta and Schuurmans [13]. This algorithm performs
linear regressions on every possible contiguous subset of the truncated data and
locates the region with the lowest standard error. This is done by starting with the
bottom 20 % of the data, performing the regression to obtain the normalized resid-
ual, then adding one point to the upper end and repeating the process until the full
range is fit. The process continues by starting with the point second from the bot-
tom and fitting 20 %, then adding one point to the upper end and repeating the
process as before. The last regression data set consists of the 20 % at the top end.
This typically results in thousands of regressions but can be done pretty quickly us-
ing modern computers. The OR is the subset of the data with the lowest normalized
residual.
The range of fit, which is the first metric on quality of fit, is determined by tak-
ing the difference between the normalized y-values at the top and bottom of the fit
range. The minimum allowable range is 0.4, which corresponds to 40 % of the tan-
gency point y-value. If the range is less than the allowable, the linear region may be
unacceptably small. The minimum allowable of 40 % is based on examination of
lots of data sets. As the linear range drops below 40 % of the tangency point y-value,
there is some question about whether a clear linear range exists.
The quality of the fit can be evaluated by examining the variation in the
y-residuals as a function of x. Next the y residuals are calculated for each point
in the OR. The residual array is then separated into the following three regions
based on the range of x, 0 % to 25 % of x-range, 25 % to 75 % of x-range, and
75 % to 100 % of x-range. The slopes of the residuals versus x in the outer quar-
tiles are then calculated. The allowable residual slope is 0.05 times the slope from
the linear regression. This limits the deviation in slope in the outer quartiles
to 6 5 % of the fitted slope. If either residual slope exceeds the allowable, then
the data exhibit excessive curvature in the outer quartiles. This is the second met-
ric on quality of fit.

Sensitivity Study
The sensitivity of the SDAR algorithm to characteristics of the data set, such as
noise, digital resolution, and sampling rate, was investigated by creating synthetic
data sets where the slope in the linear region was known and the characteristics
could be controlled. This was done by creating data sets with three functional
regions, as shown in Fig. 8. Synthetic data sets with a relatively large linear region
(yb – ya) were examined, and it was found that noise and digital resolution over the
ranges considered had very little effect on the accuracy of the OR slope, so this
study focused on data sets with a small linear region.
The parameters used to create the data sets are provided in Table 1. Equations
for the derived parameters were developed such that the slopes of the functions
144 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

TABLE 1 Parameters used to generate synthetic data sets.

Data Set Parameters Set-1 Set-2

Selected Parameters
xa ¼ 0.001 0.0007
xb ¼ 0.002 0.002
xf ¼ 0.05 0.05
ya ¼ 50 100
yb ¼ 150 200
C1 ¼ 3000 3000
C2 ¼ 0.1 0.1
Test duration (seconds) ¼ 100 100
Derived Parameters
yf ¼ 798 804
M¼ 1E5 7.6923E4
a0 ¼ 0 0
a1 ¼ 0 2.088E5
a2 ¼ 5E7 9.419E7
b0 ¼ 50 46.154
x0 ¼ 4.267E-4 1.057E-4
c0 ¼ 1.423E3 1.42E3
Min mx 25 25
Min my 5 4

were the same at the boundaries between the regions. Data Set-1 had an initial non-
linear region with increasing slope, as shown in Fig. 9, and Data Set-2 had an initial
nonlinear region with decreasing slope.
The number of points generated was determined by the sampling rate times the
test duration and the sampling rates were 10 Hz and 40 Hz. Normally distributed
random noise with a mean of zero and a standard deviation proportional to the x
or y value at the upper end of the linear range (xb or yb) was added to the x and y
data independently. The proportionality factors were varied to investigate the effect
of noise level on the resulting slope determination (Eq 1):

rNx ¼ Fx xb (1)

where:
rNx ¼ standard deviation of normally distributed noise,
xb ¼ value at top of linear range, and
Fx ¼ noise factor.
The same form was used for the y-channel. Independent parameters Fx and Fy were
used to simulate noise for the x and y channels, respectively.
The data were then discretized independently in x and y to simulate the effect
of digital resolution. The digital resolution for a channel was simulated by dividing
GRAHAM, DOI 10.1520/STP159820160080 145

FIG. 9 Functional description of synthetic data sets.

REGION 3 Yf
Power Law

yb

REGION 2 c0
Linear

ya
REGION 1
Quadrac
xa x0 xb xf

the full-scale transducer value by two raised to the power of the number of bits in
the data acquisition system, as shown in Eq 2:
FSx
DxR ¼ (2)
2n

where:
DxR ¼ digital resolution, and
n ¼ number of bits for data acquisition system.
The full-scale value was represented by a factor m times the value at the top of the
linear range (Eq 3):
FSx ¼ mx xb (3)

where:
xb ¼ value of x at the top of the linear range, and
mx ¼ ratio of the transducer full-scale to final value.
Digital resolution for the y-channel was simulated using the corresponding my
value. Both x and y channels used the same number of bits, n ¼ 12. The minimum
m values were xf/xb and yf/yb for the x and y channels, respectively. For the two data
sets described earlier, the minimum mx is 25 and the minimum my are 5.3 and 4 for
the two data sets, respectively. The values investigated were:
mx ¼ 0, 25, 100
my ¼ 0, 6, 24
Fx ¼ 0, 0.01, 0.05
Fy ¼ 0, 0.01, 0.05
146 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

The zero values correspond to cases where the data set was not digitized, or
there was no noise added to the dataset. Zero values are used to isolate the effect of
noise or resolution for a specific channel. As they are not realistic because all data
acquisition systems have some noise and limited resolution, the results from those
cases will not be presented. The mx values were higher than the my values because,
in a typical test of a stiff specimen (high slope), the x-value at the upper end of the
linear region will be a much smaller portion of the x-transducer full-scale range
than the y-value is to the y-transducer full-scale range. If the noise is less than the
digital resolution, it will not be fully captured in the digitized signal and, conse-
quently, the algorithm to extract the noise will not be reliable. The minimum Fx val-
ues corresponding to mx values of 25 and 100 are 0.002 and 0.008, respectively, so a
minimum Fx ¼ 0.010 was used in this study. The minimum Fy values corresponding
to my of 6 and 24 are 0.0005 and 0.002, respectively, which are significantly smaller
than the Fx values. These minimum Fy values are very small and do not test the
algorithm, so the same F values are used for both the x and y channels.
Results for a subset of all the combinations of noise, resolution, and sampling
rate that were investigated are provided in Table 2.

Discussion
Referring to Table 2, Case 1 for each data set serves as the benchmark where digital
resolutions are based on full-scale values that are just larger than the final values,
xf and yf, and noise is relatively low (1 % of xb and yb).
Case 2 has five times the noise in x compared with the benchmark. For Data Set-1,
the slope error increased significantly from þ1.3 % to 18.1 %. This is because the
noise extends the OR significantly (0.009 on the left and 0.0018 on the right), and the
points added to each end tend to drive the slope down (Fig. 10). For Data Set-2, the in-
crease in error is much smaller, 2.29 % to 3.07 %, because the points added to the
left of the OR drive the slope up, while those added to the right drive it down (Fig. 11).
Case 3 has five times the noise in y compared with the benchmark. For Data
Set-1, the slope error decreases from 18.1 % for Case 2 to 12.0 %. For the shape
of this data set, y-noise has less effect on slope than x-noise. For Data Set-2, the
effect is the opposite; the slope error increases from 3.07 % for Case 2 to 5.76 %.
From this comparison, it is apparent that error in slope is a complex interaction
between the magnitude of noise on the two channels and the shape of the curve.
Case 4 has a digital resolution in x that is one-quarter of the benchmark. Comparing
Cases 1, 2, and 4 shows that digital resolution has less effect on slope error than noise for
the x-channel. This is true for both data sets; however, for Data Set-2, the Case 4 error is
about one-quarter of the benchmark (0.61 % versus 2.29 %). This may be because
having a low digital resolution partially filters out some of the noise. It is not clear why
Data Set-1 does not show this partial filtering effect (6.75 % versus þ1.3 %).
Case 5 has a digital resolution in y that is one-quarter the benchmark. For Data
Set-1, this reduces the slope error from 6.75 % for Case 4 to 3.72 %. For Data
TABLE 2 Results of sensitivity study.

Sample Relative Predicted Slope Prediction Fit Quality


Case # Rate (Hz) mx Fx my Fy x @ BORa x @ EORb Fit Range Slope for OR Error Checks

Data set-1: Increasing initial slope: Actual values ! 0.001 0.002 1.00000Eþ05
1 10 25 0.01 6 0.01 0.0008 0.0023 1.19 1.01304Eþ05 1.30 %
2 10 25 0.05 6 0.01 0.0001 0.0038 2.08 8.18960Eþ04 18.10 % Fails curvature
3 10 25 0.01 6 0.05 0.0002 0.0034 1.96 8.79645Eþ04 12.04 % Fails curvature
4 10 100 0.01 6 0.01 0.0005 0.0028 1.58 9.32476Eþ04 6.75 % Fails curvature
5 10 25 0.01 24 0.01 0.0006 0.0026 1.48 9.62774Eþ04 3.72 % Fails curvature
6 40 100 0.01 6 0.01 0.00054 0.0027 1.55 9.50815Eþ04 4.92 % Fails curvature
Data set-2: Decreasing initial slope: Actual values ! 0.0007 0.002 7.69231Eþ05
1 10 25 0.01 6 0.01 0.0006 0.0026 1.16 7.51637Eþ04 2.29 % Fails curvature
2 10 25 0.05 6 0.01 0.0001 0.0036 1.68 7.45638Eþ04 3.07 % Fails curvature
3 10 25 0.01 6 0.05 0.0003 0.0036 1.85 7.24944Eþ04 5.76 % Fails curvature
4 10 100 0.01 6 0.01 0.0004 0.0027 1.22 7.64564Eþ04 0.61 % Fails curvature

GRAHAM, DOI 10.1520/STP159820160080


5 10 25 0.01 24 0.01 0.0005 0.0027 1.22 7.61913Eþ04 0.95 % Fails curvature
6 40 100 0.01 6 0.01 0.00044 0.00274 1.40 7.63678Eþ04 0.56 % Fails curvature

Note: aBOR ¼ beginning of optimum region; bEOR ¼ end of optimum region.

147
148 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 10 Data Set-1, Case 2: Fit region and residuals plots.

Set-2, the slope error increases from 0.61 % for Case 4 to 0.95 %. Again, we see
that slope error is a complex interaction between the digital resolution of the two
channels and the shape of the curve.
Case 6 has a sampling rate that is four times Case 4, with the same noise and digital
resolutions. For Data Set-1, the extension of the OR is less on both ends and, conse-
quently, the slope error decreases from 6.75 % for Case 4 to 4.92 %. For Data Set-2,
the extension of the OR is less on the left and more on the right, which accounts for
the relatively small change in error from 0.61 % for Case 4 to 0.56 %. Now we see
that the effect of sampling rate on slope error is also related to the shape of the curve.
The two data sets studied here had an intentionally small linear region; the rela-
tive fit ranges were only 19 % and 16 % larger than the minimum allowable. In all but

FIG. 11 Data Set-2, Case 2: Fit region and residuals plots.


GRAHAM, DOI 10.1520/STP159820160080 149

one case (benchmark for Data Set-1, Case 1), the final fit failed the quality check on
curvature. This is because noise and digital resolution are extending the OR beyond
the actual linear region. This check serves as a warning to the analyst to examine the
fit. The slopes of the residuals in the outer quartiles can be seen in Fig. 10 and Fig. 11.
The current draft of ASTM WK55299, New Practice for Determination of the
Slope in the Linear Region of a Test Record, has a maximum allowable noise level of
0.005 times the tangency point value, which is generally a little higher than the
point xb, yb. For the narrow linear range data sets considered in this study, the tan-
gency point is approximately twice xb and yb. Consequently, the benchmark noise
level of 0.01 corresponds pretty closely with the maximum allowable and, for these
data sets, results in slope errors of 1.3 % to 2.3 %.
The current draft standard has a minimum allowable digital resolution of three
times the tangency point value divided by 212. For these narrow linear range data
sets, this corresponds to m values of about 6. This is practical for the y-channel (typ-
ically, force) but does not provide a high enough full-scale value for the x-channel
(typically, displacement). The results for these data sets show that m for the
x-channel can be increased to 25 without incurring more than 2.3 % slope error.

Conclusions
1. Determination of slope is an important part of many ASTM standards and
can have a significant effect on the precision and bias of the standards.
2. Fixed fit ranges that are often used for linear regression can lead to slope bias
for data sets that exhibit curvature within the fixed range. This is particularly
true for data sets that have an initial curvature or that have a small linear
region.
3. Relying on the judgment of an analyst to determine fit range is time consum-
ing and subjective. Two or more analysts are not likely to get the same slopes
when analyzing the same data set.
4. There is a need for an ASTM standard practice for determining slope that
does not rely on the judgment of the analyst. Having such a standard would
reduce the uncertainty in calculated values that rely on slope determination.
The standard should include quality checks to alert the analyst if the data are
noisy or have low digital resolution, if the data set exhibits a small linear range,
or if the regression analysis is picking up data in a nonlinear part of the curve.
5. In general, when a data set has an initial nonlinear region with increasing
slope and decreasing slope after the linear region, noise causes an underpredic-
tion of slope because the data outside the linear region that are included in the
regression pulls the slope down.
6. When a data set has an initial nonlinear region with decreasing slope, the
effect of noise is less pronounced because the data picked up at the lower end
are compensated for by data picked up at the higher end. The two effects offset
one another and the slope error is smaller.
7. Noise has much more effect on slope prediction error than digital resolution
over the ranges considered in this sensitivity study.
150 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

References

[1] ASTM E8/E8M-16a, Standard Test Methods for Tension Testing of Metallic Materials,
ASTM International, West Conshohocken, PA, 2016, www.astm.org

[2] ASTM E9-09, Standard Test Methods of Compression Testing of Metallic Materials at
Room Temperature, ASTM International, West Conshohocken, PA, 2009, www.astm.org

[3] ASTM E111-04, Standard Test Method for Young’s Modulus, Tangent Modulus, and Chord
Modulus, ASTM International, West Conshohocken, PA, 2010, www.astm.org

[4] ASTM E399-12e3, Standard Test Method for Linear-Elastic Plane-Strain Fracture Tough-
ness KIc of Metallic Materials, ASTM International, West Conshohocken, PA, 2012,
www.astm.org

[5] ASTM E1290-08, Standard Test Method for Crack-Tip Opening Displacement (CTOD)
Fracture Toughness Measurement, ASTM International, West Conshohocken, PA, 2008,
www.astm.org

[6] ASTM E1820-08a, Standard Test Method for Measurement of Fracture Toughness, ASTM
International, West Conshohocken, PA, 2015, www.astm.org

[7] ASTM B557-06, Standard Test Methods for Tension Testing Wrought and Cast Alumi-
num- and Magnesium-Alloy Products, ASTM International, West Conshohocken, PA,
2015, www.astm.org

[8] Graham, S. M. and Stanley, R., “Test Results from Round Robin on Precracking and CTOD Test-
ing of Welds,” J. ASTM Int., Volume 5, No. 9, 2008, pp. 1–26, https://doi.org/10.1520/JAI101540

[9] McKeighan, P. C. and James, M. A., “Analysis Round Robin Results on the Linearity of
Fracture Toughness Test Data,” Application of Automation Technology in Fatigue and
Fracture Testing and Analysis, ASTM STP1571, P. McKeighan and B. Arthur, Eds., ASTM In-
ternational, West Conshohocken, PA, 2014, pp. 116–133, http://dx.doi.org/10.1520/
STP157120130117

[10] McKeighan, P. C., “ASTM Task Group Activity Report: E399 Data Analysis Round Robin
Results,” ASTM International, West Conshohocken, PA, 2011, unpublished.

[11] Graham, S. M. and Adler, M. A., “Determining the Slope and Quality of Fit for the Linear
Part of a Test Record,” J. Test. Eval., Vol. 39, No. 2, 2011, pp. 1–9, https://doi.org/10.1520/
JTE103038

[12] Luecke, W., Ma, L., Graham, S., and Adler, M., “Repeatability and Reproducibility of Com-
pression Strength Measurements Conducted According to ASTM E9,” NIST Technical
Note 1679, National Institute of Standards and Technology, Gaithersburg, MD, 2009.

[13] Scibetta, M. and Schuurmans, J., “Development and Qualification of an Algorithm for
the Determination of the Initial Linear Portion of a Force Versus Displacement Record,”
J. Test. Eval., Vol. 32, No. 6, 2004, pp. 500–503. https://doi.org/10.1520/JAI101540.
FATIGUE AND FRACTURE TEST PLANNING, TEST DATA ACQUISITIONS AND ANALYSIS 151

STP 1598, 2017 / available online at www.astm.org / doi: 10.1520/STP159820160050

Michael Wächter1 and Alfons Esderts2

Contribution to the Evaluation of


Stress-Strain and Strain-Life Curves
Citation
Wächter, M. and Esderts, A., “Contribution to the Evaluation of Stress-Strain and Strain-Life
Curves,” Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis, ASTM STP1598,
Z. Wei, K. Nikbin, P. C. McKeighan, and D. G. Harlow, Eds., ASTM International, West Conshohocken,
PA, 2017, pp. 151–185, http://dx.doi.org/10.1520/STP1598201600503

ABSTRACT
This paper presents an overview of different methods of performing a regression
on the results of strain-controlled tests to obtain the cyclic properties that are
used to describe the strain-life and stress-strain curves. For the strain-life curve,
the approach of Coffin and Manson is used, whereas the cyclic stress-strain
curve is described using the approach of Ramberg and Osgood. Three different
regression approaches are explained: a simple linear regression and a Deming
regression, which are applied separately to the elastic part of the strain
amplitude and the number of cycles and to the plastic part of the strain
amplitude and the number of cycles, as well as a spatial regression applied to
the stress amplitude, plastic strain amplitude, and number of cycles. Because
the populations of the parameters that are to be approximated by the
regression are usually unknown, Monte Carlo simulations are performed to
simulate artificial strain-controlled tests as a basis for judging the accuracy of
the approximation of the mean curves and the scatter around these curves. It is
shown that all three regressions are almost equally accurate. By using the
simulation results, factors for shifting the estimated properties to a confidence
level of 10 % or 90 % are derived.

Manuscript received February 29, 2016; accepted for publication August 5, 2016.
1
Clausthal University of Technology, Institute for Plant Engineering and Fatigue Analysis, Leibnizstrasse 32, 38678
Clausthal-Zellerfeld, Germany http://orcid.org/0000-0001-9568-8407
2
Clausthal University of Technology, Institute for Plant Engineering and Fatigue Analysis, Leibnizstrasse 32, 38678
Clausthal-Zellerfeld, Germany http://orcid.org/0000-0002-1269-5677
3
ASTM Symposium on Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis on May 4–5,
2016 in Grand Hyatt, San Antonio, TX.

Copyright V
C 2017 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.
152 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

Keywords
stress-strain curve, strain-life curve, cyclic properties, regression, Monte Carlo
method, fatigue test, test evaluation, scatter, confidence level

Nomenclature
symbol meaning
a slope of the regression line
b fatigue strength exponent
b intercept of the regression line
c fatigue ductility exponent
E elastic modulus
i continuous index
j factor for shifting an estimate to a confidence level of 10 % or 90 %
K0 cyclic strength coefficient
mN logarithmic mean of all Nexp,m/Npre
mr logarithmic mean of all rexp;m =rpre
n sample size/number of specimens
N number of cycles
n0 cyclic strain hardening exponent
NE number of cycles for the endurance limit
Nm mean of the number of cycles
Pf probability of failure
rep number of repetitions of a simulation
slog standard deviation of the log-normal distribution
slog;corr bias corrected standard deviation of the log-normal distribution
slog;N slog for the number of cycles within the strain-life curve
slog;r slog for the stress amplitude within the stress-strain curve
xi i-th value of the regression variable x
yi i-th value of the regression variable y
De strain range
De
2 strain amplitude
Deel elastic strain range
DeE
2 strain amplitude of the fatigue limit
Depl plastic strain range
eel;D residual in elastic strain
epl;D residual in plastic strain
ra stress amplitude
ra;m mean of the stress amplitude
rD residual in stress
pre predefined value (population)
exp value from an (artificial) experiment
WÄCHTER AND ESDERTS, DOI 10.1520/STP159820160050 153

Introduction
For a given material, the stress-strain curve (ra  De
2 curve) and the strain-life curve
De
( 2  N curve), along with the cyclic properties that describe them, are usually de-
termined by conducting multiple strain-controlled fatigue tests of unnotched speci-
mens using constant amplitude loading with a constant strain ratio of
Re ¼ eemax
min
¼ 1 at various strain-range values. Throughout the literature, various
methods of evaluating such series of tests have been proposed [1–8]. They vary
both in the procedures for separating the amplitude of the total strain into elastic
and plastic components and in the manner in which the regression of the results of
the individual tests is performed to determine the cyclic properties. Some such
methods are standardized in national or international standards [1–6]; others are
not but nonetheless seem promising [7,8]. The various possible approaches to per-
forming a regression to determine the cyclic properties of a material are the focus
of this paper.
Typically, only the mean curves of the stress-strain relation and the strain-life
relation are addressed in an evaluation. To perform fatigue-life estimations for
components with probabilities of failure other than Pf ¼ 50 %, the scatter around
the mean curves also must be accurately estimated. Different methods of evaluating
the cyclic properties can also be expected to have different influences on the esti-
mated scatter.
Because the population of the cyclic properties to be estimated through
fatigue tests is usually unknown, it is difficult for researchers and test engineers to
determine which of the available techniques is preferable for obtaining accurate
estimates. In such cases, when the population is unknown, the so-called Monte
Carlo method can be helpful. Using this tool, it is possible to predefine a popula-
tion on which to perform repeated random sampling. This means that artificial
tests can be conducted in which the population is known. These artificial tests can
be evaluated in different ways, and the results can be compared with the known
population. In this manner, different evaluation techniques can be judged with
regard to the accuracy of their results, and a recommendation for the user can be
obtained.
This paper is structured as follows. First, the approaches to describing the
strain-life and stress-strain curves and the cyclic properties that describe them are
introduced. Subsequently, various evaluation techniques reported in the literature
are considered. Three specific methods of performing a regression on the results of
individual tests are selected for further investigation. The two following sections
demonstrate how the residuals of a strain-controlled test can be modeled and how
artificial tests can be simulated. Simulated tests are then evaluated in the section
thereafter, enabling an assessment of the performance of the different regression
methods and the derivation of factors for shifting the estimated parameters to a
confidence level of 10 % or 90 %. Both are summarized in the concluding section of
the paper.
154 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

Cyclic Properties
The strain-life curve and the cyclic stress-strain curve are basic functions in the field
of fatigue analysis. The strain-life curve, De
2  N, of a metallic material is typically
described using the approach of Coffin and Manson [9,10]:

De Deel Depl r0 f
¼ þ ¼ ð2N Þb þ e0 f ð2N Þc (1)
2 2 2 E

where:
De
2 ¼ amplitude of the total strain,
Deel
2 ¼ elastic part of the strain amplitude,
Depl
2 ¼ plastic part of the strain amplitude,
r0 f ¼ fatigue strength coefficient,
e0 f ¼ fatigue ductility coefficient,
b ¼ fatigue strength exponent,
c ¼ fatigue ductility exponent, and
E ¼ elastic modulus.

This equation describes the fatigue behavior of a material as a relation between the
number of cycles N that can be applied to the material before crack initiation and
the amplitude of the total strain De
2 , as shown in Fig. 1. In the Coffin-Manson approach,
the elastic and plastic parts of the total strain amplitude are considered separately.
De
Both relations, the De2 el  N and 2pl  N components, respectively, appear as straight
lines when plotted on a logarithmic scale. The relationship expressed in Eq 1 is valid
only above the endurance limit De2 E of the material. Below De2 E ; the strain-life curve is a
horizontal line for most steel materials (see Fig. 1). Certain materials may also show a
further decrease of strength for high numbers of cycles (N > NE), e.g., aluminum [11].
By multiplying the part of the strain-life curve associated with the elastic strain
amplitude in Eq 1 by the elastic modulus, the stress-life curve is obtained (Eq 2):

ra ¼ r0 f ð2N Þb (2)

where ra is the stress amplitude.


The r-e curve is often described using the approach of Ramberg and Osgood
[12], as shown in Eq 3 and Fig. 2:

De Deel Depl ra ra n10


¼ þ ¼ þ (3)
2 2 2 E K0

where:
K0 ¼ cyclic strength coefficient,
n0 ¼ cyclic strain hardening exponent, and
E ¼ elastic modulus.

This equation describes the deformation behavior of a material and can also be divided
into separate terms related to the elastic and plastic parts of the total strain amplitude.
WÄCHTER AND ESDERTS, DOI 10.1520/STP159820160050 155

FIG. 1 Strain-life curve described using the approach of Coffin and Manson.

Both the strain-life curve and the stress-strain curve are typically determined
through a series of strain-controlled tests with constant strain amplitudes and no
mean strain.
The properties r0 f , e0 f , b, c, K0 , and n0 are referred to as cyclic properties. If
both the Coffin-Manson approach and the Ramberg-Osgood approach are accepted

FIG. 2 Stress-strain curve described using the approach of Ramberg and Osgood.
156 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

to be valid for a material in question, these properties are related by the compatibili-
ty equations (Eqs 4 and 5):

b
n0 ¼ (4)
c
0 r0 f
K ¼ 0 : (5)
ðe0 f Þn

Different authors show that this assumption may not be valid for all metallic
materials [8,13]. If the approaches of Ramberg/Osgood and Coffin/Manson
are used anyway to describe the material behavior, the compatibility may not be
valid [8].

Evaluation of Strain-Controlled Tests


Recommended practices regarding the execution and evaluation of strain-
controlled tests for the determination of cyclic properties are provided in
several national and international standards and guidelines [1–6].
The result of a strain-controlled test is a data triplet of the strain amplitude,
stress amplitude, and the number of cycles endured. Although the determined
number of cycles is usually valid for crack initiation, the stress and strain amplitude
 
is determined at half-life N2 of the specimen. By this time, the material behavior
has already stabilized after the so-called initial cyclic hardening or softening that
occurs for many metallic materials.
If the Coffin-Manson or Ramberg-Osgood approach is used to describe the strain-
life or stress-strain curve, then the amplitude of the total strain must be split into the
elastic and plastic components for evaluation. Multiple opinions exist regarding how
this splitting of the strain components should be performed, and they will be consid-
ered in greater detail in the first part of this section. Another aspect of evaluation for
which different approaches exist is the type of regression that is used to determine the
cyclic properties. This aspect will be addressed in the second part of this section.

SEPARATION OF THE ELASTIC AND PLASTIC STRAINS


The most common method of dividing the elastic part from the plastic part of the
strain amplitude is as follows. The elastic strain amplitude is calculated from the
strain amplitude by means of an elastic modulus E:

Deel ra
¼ : (6)
2 E

The plastic strain amplitude can then be easily calculated by subtracting the
elastic strain amplitude from the total strain amplitude:

Depl De Deel
¼  : (7)
2 2 2
WÄCHTER AND ESDERTS, DOI 10.1520/STP159820160050 157

Differing opinions exist regarding which value should be used for the elastic
modulus in Eq 6. One possibility is to use the monotonic elastic modulus that can
be determined in a static tensile test or from the first quarter-cycle of a test (Possi-
bility 1). ASTM E606, Standard Practice for Strain-Controlled Fatigue Testing [1],
suggests the use of an elastic modulus that can be determined in cyclic tests with
loads below the elastic limit (Possibility 2). The so-called mid-life modulus of elas-
ticity, which is calculated from the slopes measured just after the maximum and
minimum of a hysteresis loop, is used in ASD-STAN prEN 3988-1998 [5] (Possibil-
ity 3). The two slopes are averaged, and the hysteresis loop to be examined is select-
ed from a range around half of the specimen’s fatigue life. This approach suffers
from the problem that each individual test may yield a different value for the elastic
modulus, and it is not clear which of them should be used in Eqs 1 and 3. A rather
pragmatic approach is presented in a set of German guidelines for the testing of
steel sheets, SEP 1240, Testing and Documentation Guideline for the Experimental
Determination of Mechanical Properties of Steel Sheets for CAE-Calculations [6], in
which a material-group-specific value of E ¼ 206 GPa for steel is used (Possibility 4).
This approach is based on the work of Masendorf [14], who showed that the elastic
modulus determined from a hysteresis loop (similar to Possibility 3) is highly depen-
dent on the strain amplitude and strain rate. Furthermore, Masendorf found that indi-
vidual tests are described by a uniform value of E almost as well as by the separate
values determined from the first-quarter hysteresis loop in each test.
An alternative approach to determining the plastic strain without requiring a
value for E is specified in BS 7270 [3] and also in ISO 12106 [4]. Here, the plastic
part of the strain amplitude is determined by measuring the hysteresis at the level
of the mean stress (Possibility 5); see Fig. 3.
Kandil compared Possibility 5 with Possibilities 2 and 3 and concluded that
Possibility 5 should be used because the moduli of elasticity as determined using
Possibilities 2 and 3 vary too widely [15]. However, in the case of Possibility 5, the
question still remains regarding which value of E is to be used in the Coffin-Manson
and Ramberg-Osgood approaches.

TYPES OF REGRESSION
After the strain amplitudes from the single tests have been divided into their elastic
and plastic components, regression can be performed to determine the cyclic prop-
erties of the tested material. Various methods of performing this regression have
been suggested in the literature and are presented here.
When determining the plastic part of the strain amplitude using the stress
amplitude (Eqs 6 and 7), values may arise that are very small compared with the
total strain or, depending on the value of the elastic modulus used, may even
become negative. To clarify this problem, let us assume a total strain amplitude of
De Deel
2 ffi 2 ¼ 0:1 % ¼ 0:001 and an elastic modulus of the material of Emat ¼ 207
GPa. The stress amplitude for this load level would be ra ¼ De2el  Emat ¼ 207 MPa.
If the strain amplitude and the stress amplitude would be measured during a test
158 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 3 Determining the plastic strain range following BS 7270:2006 [3] and ISO
12106:2003 [4].

and a value for the elastic modulus of Eeval ¼ 206 GPa is used for the evaluation of
the test results instead of the unknown Emat, the calculated value would be
Deel 207MPa
2 ¼ Eeval ¼ 0:001005, and the plastic part of the strain amplitude would
De Deel
become negative: 2pl ¼ De 2  2 ¼ 0:000005. In any case, the elastic modulus
used for the evaluation is just an estimate for the real value of the material and
therefore may have a rather strong influence on the determined value of the plastic
part of the strain amplitude. This is especially true for strain amplitudes with small
plastic parts.
Because regression is performed using the logarithm of the plastic strain com-
ponent, the uncertainty in the determination of the plastic component of the strain
amplitude would exert a rather large influence on the regression if such small values
were to be considered. Therefore, plastic strain amplitudes smaller than 0.01 % typi-
cally are excluded from the regression [6–8], and negative values cannot be consid-
ered anyway.

Simple Linear Regression in Accordance with ASTM E739 and SEP 1240
In both ASTM E739-80, Standard Practice for Statistical Analysis of Linear or Line-
arized Stress-Life (S-N) and Strain-Life (e-N) Fatigue Data [2], and in SEP 1240 [6],
separate regressions are performed for the elastic and plastic parts of the strain-life
curve. It is assumed that the number of cycles is the dependent (random) variable,
whereas the elastic and plastic strain amplitudes, respectively, are the independent
variables. In ASTM E739-80 [2], it is noted that one could question whether the
elastic and plastic strain amplitudes are really independent variables because the to-
tal strain amplitude is the controlled quantity. Nonetheless, there is no doubt that
the fatigue life is a dependent variable. Therefore, a simple
 linear regression is per-
  De
formed either on log De2el and logðN Þ or on log 2pl and logðN Þ, respectively,
while minimizing the squared residuals of logðN Þ; see Fig. 4. Test results in which
the plastic strain amplitudes are smaller than 0.01 % are excluded from the
WÄCHTER AND ESDERTS, DOI 10.1520/STP159820160050 159

FIG. 4 Simple linear regression.

regression of the plastic part of the strain-life curve [6]. All test results are used for
the regression of the elastic part when the elastic part of the strain-life curve is as-
sumed to be equal to the total strain amplitude [6].
The equations
 used for the regression are described as follows [2]. First,
Deel  De
log 2 or log 2pl is replaced with x and logðN Þ is replaced with y to simplify the
equations. Eqs 8 and 9 are valid for the elastic part of the strain-life curve:
 
Deel;i
xi ¼ log (8)
2
yi ¼ logðNi Þ: (9)

Eqs 10 and 11 are valid for the plastic part:


 
Depl;i
xi ¼ log (10)
2
yi ¼ logðN i Þ: (11)

The regression lines take the form given in Eq 12:

y ¼ a x þ b : (12)

The slope and the intercept with the y axis are calculated using Eqs 13 and 14:
P P  P 
 xi  n1 xi yi  n1 yi
a ¼ P P 2 (13)
xi  n1 xi
1X 1X
b ¼ y i  a xi : (14)
n n
160 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

where n is the number of specimens in the given data set; a* and b* are used to deter-
mine the cyclic properties that describe the elastic and plastic parts of the strain-life
curve:

1
b¼ (15)
a
b
r0 f ¼ E  0:5b  10a (16)
1
c¼ (17)
a
b
e0 f ¼ 0:5c  10a : (18)

Whereas ASTM E739-80 [2] provides no information concerning how to ob-


tain the cyclic properties of the cyclic stress-strain curve, SEP 1240 [6] states them
to be defined by the compatibility conditions, Eqs 4 and 5.

Orthogonal Regression Following Bäumel and Seeger


Because it may be doubted whether the elastic and plastic parts of the strain ampli-
tude are truly independent variables, as mentioned before, Bäumel and Seeger [7]
suggest applying an orthogonal regression, which is a special case of Deming regres-
sion [16] and also a type of total least-squares (TLS) regression.
As in the simple linear regression approach, the components associated with
the elastic and plastic strains are evaluated separately. The difference is that, in this
case, the perpendicular residuals between the test results and the regression line are
minimized (see Fig. 5), which suggests that the elastic or plastic strain component
and the number of cycles are both dependent variables.

FIG. 5 Deming regression.


N
y

x
WÄCHTER AND ESDERTS, DOI 10.1520/STP159820160050 161

The orthogonal regression is performed in a manner quite similar to the simple


linear regression (see Eqs 8–18), but the equation for the slope of the regression line
becomes more complex [17] (Eq 19):

rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
P 
P P 2 P  P 2 P 2 P  P 2 2 P  P  P 2
yi  n1 yi  xi  n1 xi þ yi  n1 yi  xi  n1 xi þ4 xi  n1 xi yi  n1 yi
a ¼ P P   P  :
2 xi  n1 xi yi  n1 yi
(19)

As in the case of the simple linear regression, plastic strain amplitudes of less
than 0.01 % are excluded from the regression.

Spatial Regression Following Nieslony et al.


A third alternative method of performing the regression has been suggested by
Nieslony et al. [8]. These authors perform a spatial regression using three quanti-
ties: the stress amplitude, the plastic strain amplitude, and the number of cycles.
This is possible because the logarithms of all three quantities appear as a straight
line in a three-dimensional coordinate system. Nieslony et al., therefore, minimize
the spatial perpendicular distances between the individual test results and the re-
gression line, as shown in Fig. 6. Thus, the compatibility requirements between the
cyclic stress-strain curve and the strain-life curve are satisfied automatically [8],
which may not be the case for the other two types of regression. The equations

FIG. 6 Spatial regression.


z

x
y
162 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

required for this spatial regression approach are rather complex and may be found
in Jacquelin [18].
For the simple linear regression and the Deming regression, appearance of
test results with small plastic strain amplitudes is not a problem. These test
results are just excluded from the regression of the plastic part of the strain-life
curve, as described earlier, but are still considered for the regression of the elastic
part. For the spatial regression, the exclusion of test results with small plastic
strain amplitudes would result in them not being considered at all in the evalua-
tion of the cyclic properties. Therefore, these values are excluded only in an ini-
tial regression. Using the cyclic properties that are obtained from this first
regression, the plastic strain amplitudes smaller than 0.01 % are then recalcu-
lated, and a second regression using all test results is performed to determine the
final set of cyclic properties [8].

Model of the Residuals in Strain-Controlled Tests


In this section, a model of the residuals of individual test results around the Coffin-
Mason and Ramberg-Osgood curves is developed. To this end, the strain-life and
cyclic stress-strain curves are considered more closely. The two functions contain
dependencies among five quantities:
• The number of cycles, N
• The total strain amplitude, De
2
• The elastic part of the strain amplitude, Deel
2
De
• The plastic part of the strain amplitude, pl
2
• The stress amplitude, ra

For each of the aforementioned quantities, a residual around at least one of the
functions can be observed. It is shown hereafter that this rather large number of
residuals is not necessary to characterize the deviation of the individual tests. In
fact, it can be reduced to the deviation of only two quantities. Therefore, the follow-
ing approach is taken.
• First, the cyclic stress-strain curve and the strain-life curve are considered
more closely.
• Afterward, one individual test result is examined with respect to the curves de-
termined in a regression, and it is assumed that the distance of this test result
from the mean curve can be regarded as representative of the scatter.
• It is shown which of the different deviations are appropriate for characterizing
the scatter around the cyclic stress-strain curve and strain-life curve.
The considered model depends on the relation among the elastic, plastic, and
total strain amplitudes. This is assumed to be sufficiently described by Eqs 6 and 7.
The individual tests for determining the cyclic properties of a material typically
are performed under total strain control. Because the total strain is the controlled
quantity, the stress amplitude and the number of cycles are observed quantities.
Therefore, it is assumed that the total strain amplitude is not subject to errors and
therefore is an independent variable, whereas the other two quantities are
WÄCHTER AND ESDERTS, DOI 10.1520/STP159820160050 163

dependent variables. The number of cycles and the stress amplitude are random
variables that depend on the amplitude of the total strain. The remaining
quantities—namely, the elastic and plastic strain amplitudes—depend on the other
quantities and are influenced by their scatter.
For the following analysis, it is essential to consider the cyclic stress-strain
curve more closely. This curve describes the relation between the amplitudes of the
total strain and the stress and can be separated into two relations: the relation be-
tween the stress amplitude and the elastic strain amplitude and the relation between
the stress amplitude and the plastic strain amplitude (see Figs. 7a–c and Eq 3).
Because the elastic part of the total strain can be calculated using the stress am-
plitude and the modulus of elasticity (Eq 6) but cannot be measured directly like
the total strain, the relation between the stress and the elastic strain amplitude does
not contain any information regarding the behavior of the material except the mod-
ulus of elasticity that is used to calculate it. By contrast, the relation between the
stress amplitude and the plastic strain amplitude contains the two cyclic properties

FIG. 7 Different ways to illustrate the deformation behavior of a material: As stress-


strain curve (a) relation between stress amplitude and amplitude of elastic
strain, (b) relation between stress amplitude and amplitude of plastic strain, and
(c) relation between the amplitudes of elastic and plastic strain.

n n
E K E K

n
K
164 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

K0 and n0 of the Ramberg-Osgood equation and describes the plastic deformation


behavior of the material; see Fig. 7c.
Another diagram can be drawn by substituting the amplitude of the elastic
strain for the stress amplitude, as shown in Fig. 7d. As mentioned earlier, the stress
amplitude and the amplitude of the elastic strain provide identical information as
long as the elastic modulus is known. These two quantities are proportional to each
other. Thus, the curves shown in Fig. 7c and d are equivalent with regard to the
information they contain.
To demonstrate the behavior observed in an individual test under the influence of
residuals with regard to the number of cycles and the stress amplitude, the following fictive
exemplary scenario is considered: Several strain-controlled tests are conducted at different
strain amplitudes for a certain material. The results from these tests are used to determine
the strain-life and stress-strain curves via regression. For the following explanation, the
reader shall bring to mind a single test result and its position with respect to the deter-
mined curves. It is assumed that the test is conducted at a total strain amplitude of
De
2 ¼ 0:007. Furthermore, it is assumed that the observed stress amplitude is larger than
that indicated by the cyclic stress-strain curve at the same strain amplitude. This residual
is represented by the distance rD in Fig. 8a. Again, this combination of total strain ampli-
tude and stress amplitude is simply an example. The stress amplitude in a particular test
could also be smaller than the stress amplitude indicated by the cyclic stress-strain curve.
Figs. 8b and 8c shows the elastic strain amplitude plotted alongside the plastic strain ampli-
tude, on both linear and logarithmic scales. Because the elastic stress amplitude is calculat-
ed from the stress amplitude using Eq 6, the distance rD found from Fig. 8a can also be
determined as eel,D in Fig. 8c because the expression in Eq 20 holds:
rD
eel;D ¼ : (20)
E

Because the sum of the elastic and plastic strain amplitudes is always equal to
the total strain amplitude, a test result in which the elastic strain amplitude is larger
than that indicated by the mean curve must correspond to a plastic strain amplitude
that is smaller than that indicated by the mean curve by the same amount, as shown
in Fig. 8c (Eq 21):

epl;D ¼ eel;D : (21)

To illustrate the influence of the scatter around the strain-life curve on a single
test, a single test with a number of cycles larger than that of the strain-life curve is
imagined. The distance is indicated by the range ND in Fig. 9a. This residual can
also be determined from the elastic and plastic parts of the strain amplitude. By also
considering the already known distances eel,D and epl,D from Fig. 8c, the deviation of
the test result from the mean curve can be described; see Fig. 9b and c.
It can be concluded that, to describe the scatter around these curves and all
related quantities, it is sufficient to know the scatter of the number of cycles with
WÄCHTER AND ESDERTS, DOI 10.1520/STP159820160050 165

FIG. 8 Different ways to illustrate the residual of a single test to the stress-strain curve
(a) in the stress-strain curve, (b) for the elastic and plastic parts of the strain
amplitude (logarithmic scale), (c) for the elastic and plastic parts of the strain
amplitude (linear scale).

respect to the strain-life curve and the scatter of the stress amplitude with respect to
the cyclic stress-strain curve. However, this statement is valid only if the relations
between the stress amplitude and the elastic, plastic, and total strain amplitudes
(Eqs 6 and 7) are valid—at least to a reasonable approximation. This assumption is
adopted in the discussion that follows.
For convenience, the following assumptions are sometimes made in the field of
fatigue analysis although different behaviors may be observed in reality: first, that
the log-normal distribution is applicable, and second, that the variance remains
constant throughout the entire range of fatigue life (the data set is assumed to be
homoscedastic); see, for example, ASTM E739-80 [2]. Under these assumptions,
166 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 9 Different ways to illustrate the residual of a single test around the strain-life
curve.

N N

the standard deviation of the log-normal distribution can be used to describe the
scatter (Eqs 22 and 23):
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2ffi
n  
1 X Ni
slog;N ¼ log (22)
n  2 i¼1 N m;i
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2ffi
n  
1 X ra;i
slog;r ¼ log (23)
n  2 i¼1 ra;m;i

where:
n ¼ sample size (number of failed specimens) in a test series,
Ni ¼ number of cycles in the i-th individual test,
Nm,i ¼ mean number of cycles at the strain amplitude of the i-th individual test, Fig. 10a,
WÄCHTER AND ESDERTS, DOI 10.1520/STP159820160050 167

FIG. 10 Determination of the mean values of the number of cycles (a) and the stress
amplitude (b) for a single test.

N N
N

ra,i ¼ stress amplitude of the i-th individual test, and


ra,m,i ¼ mean stress amplitude at the strain amplitude of the i-th individual
test, Fig. 10b.

While the variance s2log is known to lead to unbiased estimates of the variance of the
population, Martin, Hinkelmann, and Esderts [19] showed that the standard devia-
tion slog (Eqs 22 and 23) tends to underestimate the standard deviation of a popula-
tion for small sample sizes, which is caused by the square root function that is used
to calculate the standard deviation from the variance [19]. To obtain unbiased esti-
mates, Martin, Hinkelmann, and Esderts derived Eq 24 empirically to correct the
results of Eqs 22 and 23.
n  1:74
slog;corr ¼ slog  (24)
n2

Simulation of Strain-Controlled Tests


This section examines which of the discussed methods of regression yields the most
accurate estimates of the cyclic properties and of the standard deviation in the
stress-strain curve and in the strain-life curve. For this purpose, Monte Carlo simu-
lations are performed. This type of simulation offers the possibility of performing
random experiments with a known population. Monte Carlo simulations are well-
known in the field of fatigue analysis. Hück, for example, uses this tool to improve
the evaluation of staircase tests [20]. It has also been used in more recent papers to
obtain information regarding the accuracy of various testing and evaluation techni-
ques used in fatigue analysis: [19,21–27].
By drawing so-called pseudorandom numbers from a predefined population,
an artificial series of strain-controlled tests can be generated. The results of these
artificial tests consist of the three values of the strain amplitude, stress amplitude,
and number of cycles, just as in real tests. In this manner, it is possible to evaluate
168 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

the performance of regression methods using Monte Carlo simulations by com-


paring the cyclic properties as evaluated via regression with those of the true
population.
In the following subsections, the simulation model used to generate the artifi-
cial test series is first presented. Later, the quality criteria used in the evaluation are
introduced, and the simulation results are shown.

SIMULATION MODEL
To define populations for the simulation of artificial strain-controlled tests, the
cyclic properties determined in the evaluations of a number of 369 different steel
materials from [7,14,28–45].
For the simulation of single tests, it is important to know whether dependencies
exist among the cyclic properties. If so, these dependencies should be included in
the simulation model. From Fig. 11, there appear to be no dependencies between the
coefficients r0 f and e0 f or between the exponents b and c. A rather strong depen-
dency can be observed, however, between the properties of the plastic part of the
strain amplitude (e0 f ) and c. A much less-pronounced but still noticeable dependency
can be seen between the properties of the elastic part of the strain amplitude (r0 f )

FIG. 11 Relations between the cyclic properties of steel.


WÄCHTER AND ESDERTS, DOI 10.1520/STP159820160050 169

and b. It seems that the elastic part of the strain-life curve becomes steeper with
increasing r0 f .
Dependencies as discussed here may result from one or from both of the fol-
lowing sources:
1. There is a real dependency between the intercepts and the slopes. One could
assume that high-strength steels have low intercepts and flat strain-life curves,
while low-strength steels show the opposite behavior (or vice versa).
2. The shown dependencies result from statistical variations that lead to errors in
the estimation of the regression lines. The number of cycles N ¼ 0.5 for which
r0 f and e0 f are calculated lies outside the area of N for which experimental
results exist. This means that the regression lines must be extrapolated over a
rather large region to determine r0 f and e0 f . Small errors in the estimation of
the regression lines lead to large errors in r0 f and e0 f and thus to combinations
of underestimated coefficients and slopes that are too small as well as overesti-
mated coefficients and slopes that are too steep.
It is very probable that Source 2 has an effect on the results. The question is whether
Source 1 needs to be considered, too. Therefore, another property, the tensile strength,
can be examined. Using the same data sources as in this study, it is shown by Wächter,
Esderts, and Masendorf [46] that steel materials with high tensile strength show higher
values of r0 f and lower values for e0 f when compared to steels with low tensile strength;
there is no relation between the tensile strength and the exponents b and c. This supports
the findings of Bäumel and Seeger [7], who use constant values for b and c to predict
strain-life curves from the tensile strength. Therefore, it is reasoned that there are no dis-
trict dependencies between r0 f and b or e0 f and c that result from Source 1.
The shown dependencies between r0 f and b as well as between e0 f and c disap-
pear if the stress amplitude and plastic strain amplitude at a higher number of
cycles, rather than the intercepts r0 f and e0 f at N ¼ 0.5, are used to define the two
components of the strain-life curve. If, for example, the stress amplitude at a num-
ber of cycles equal to N ¼ 3000 and the amplitude of the plastic strain at N ¼ 600
are used, then the dependencies between the slopes and intercepts disappear, as
seen in Fig. 12. For values of N > > 3000 or N > > 600, respectively, the aforemen-
tioned dependencies reverse. These two numbers of cycles can be replaced with
others in the same range (N  3000 and N  600), in which case a behavior similar
to that observed in Fig. 12 will appear. The set values (N ¼ 3000 and N ¼ 600) were
chosen on the one hand because the dependencies disappear, as described earlier,
and on the other hand because these values also lead to the best relations between
the support points in the strain-life curve and the tensile strength in Wächter,
Esderts, and Masendorf [46]; N ¼ 3000 and N ¼ 600 are closer to the center of the
area of fatigue life for which experimental results exist. Therefore, the effects of the
errors in the regression lines described earlier from Source 2 are reduced.
Although the typical ranges of the cyclic properties can be identified from these test
results, there is not enough information to determine the type of distribution functions.
The log-normal distribution is assumed to describe both the cyclic properties and the
170 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 12 Relations between the cyclic properties of steel as determined using the stress
amplitude at N ¼ 3000 and the plastic strain amplitude at N ¼ 600 to define the
strain-life curve.
N

N N

standard deviations slog,N and slog,r. From the data considered earlier, there is no evi-
dence that indicates that another type of distribution function would be more appropri-
ate for describing the cyclic properties. The idea of considering the scatter of the
deviation with respect to the fatigue-life curve within a simulation model is attributed
to Müller et al. [24].
To simulate random populations for determining the strain-life curves, the log-
arithmic means and standard deviations of the log-normal distributions for the fol-
lowing quantities were determined:
• The stress amplitude at 3000 cycles
• The fatigue strength exponent b
• The elastic modulus
• The plastic strain amplitude at 600 cycles
• The fatigue ductility exponent
• The logarithmic standard deviation in the direction of N in the strain-life
curve, slog,N
• The logarithmic standard deviation in the direction of ra in the stress-strain
curve, slog,r
WÄCHTER AND ESDERTS, DOI 10.1520/STP159820160050 171

The means m and standard deviations slog of the distributions that are used for
these quantities are listed in Table 1.
A negative sign is added to the exponents b and c because a log-normal distri-
bution cannot be defined otherwise.
The simulation of the test series is performed in the following way: First, ran-
De
dom values of the properties ra at N ¼ 3000, b, E, 2pl at N ¼ 600, and c are generat-
ed using the distribution parameters given in Table 1.
Based on these values, a mean strain-life curve can be defined by calculating the
coefficients of the Coffin-Mason approach (Eqs 25 and 26).
 
0 0:5 b
rf ¼  ra (25)
3000
 c
0:5 Depl
e0 f ¼  : (26)
600 2

Using the compatibility equations, Eqs 4 and 5, the properties K’ and n’ of the
Ramberg-Osgood approach (Eq 3) are calculated and the strain-life curve thereby is
defined automatically. The scatter around the two curves is defined by generating
random values of slog,N and slog,r. The corresponding standard deviations are distin-
guished by the index pre, indicating “predefined,” for later reference. From the scat-
ter model introduced previously, it is clear that with the scatter predefined in the
direction of N and ra, the deviations of the elastic and plastic amplitudes are con-
sidered automatically.
To obtain individual test results, artificial tests are conducted at multiple levels
of strain, just as in real tests. Within the simulations, the sample size—namely, the
number of specimens in a test series—is varied between 8 and 20. Within each test
series, each test is conducted at a unique strain amplitude. The strain levels are
equally logarithmically spaced between 0.1 % and 1 %, meaning that each pair of
neighboring strain levels is always related by the same multiplicative factor. The
number of cycles of the mean strain-life curve, Npre, and the strain amplitude of the

TABLE 1 Means and standard deviations of the log-normal distributions for the parameters of the
population used in the simulation.

Quantity X m(X) slog(X)

ra ðN ¼ 3000Þ in MPa 418 0.165


Depl
2 ðN ¼ 600Þ 0.0074 0.199
b 0.0853 0.142
c 0.574 0.109
E in MPa 205,810 0.0224
slog,N 0.120 0.221
slog,r 0.013 0.236
172 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

stress-strain curve, rpre, are calculated for each strain level. Then, using the values
Npre and rpre, which serve as the means of the distributions, and the standard devi-
ations slog,N,pre and slog,r,pre, random values of N and r are generated. These values
are labeled Nexp and rexp.
The equation of Coffin and Manson for the strain-life curve is just valid above
the fatigue limit or below a number of cycles smaller NE, respectively. Depending
on the concrete material, the number of cycles of the transition into the fatigue limit
regime may vary between, for example, N ¼ 5  105 as reported by Bäumel and See-
ger [7], and higher values up to 5108 [11]. For the simulation performed in this
study, in the case of an Nexp value greater than 5  105 cycles, the corresponding test
is designated as stopped before failure (run-out) and is excluded from further evalu-
ation. Because strain-controlled tests are usually conducted at low-test frequencies,
there are just a few test results that show numbers of cycles near the fatigue limit,
and the values used to predefine the populations in Table 1 are not statistically firm
in this region of fatigue life. Therefore, the rather low value of N ¼ 5  105 is chosen
to stop the artificial tests and to define them as run-outs.
The procedure used to simulate the strain-controlled tests is summarized in Fig. 13.

SIMULATIONS AND EVALUATION OF THE RESULTS


As described earlier, for the simulation of the strain-controlled tests, random values are
drawn from predefined populations. For this purpose, random numbers (pseudorandom
numbers) are generated by a computer using the Mersenne Twister [47], which is distin-
guished by a very long period. Therefore, it is able to generate pseudorandom numbers
of high quality that are close to real random numbers. Initially, this random number gen-
erator generates random numbers equally distributed between 0 and 1. Using predefined
values of the mean and standard deviation, these equally distributed random numbers
are then converted into random numbers from a specified log-normal distribution.
To determine the relative performance of the regression methods of interest,
the test series that are simulated as described earlier are evaluated using each regres-
sion method. To split the strain amplitudes into their elastic and plastic compo-
nents, the material-group-specific elastic modulus of E ¼ 206 MPa is used, in
accordance with SEP 1240 [6]. The predefined quantities Npre, rpre, slog,N,pre, and
slog,r,pre are compared with those determined from each regression. For the strain-
life curve, the experimentally determined numbers of cycles Nexp,m, as shown in
Fig. 14a, are compared with the predefined values Npre by calculating the ratios
Nexp,m/Npre. The logarithmic mean mN and the standard deviation slog,N of the log-
normal distribution were calculated for the n ratios obtained from one simulated
test series. For an ideal evaluation technique, the average over all repetitions (rep)
of the simulation should be 1 for both mN and slog,N/slog,N,pre. These are the criteria
of an unbiased estimator. The bias for quantity X can be written as in Eq 27:

1
P
rep

rep logðXi Þ
biasðX Þ ¼ 10 i¼1 : (27)
WÄCHTER AND ESDERTS, DOI 10.1520/STP159820160050 173

FIG. 13 Procedure for the simulation of strain-controlled tests (example for a test series
comprising five individual tests).

N
p p

N N

For the stress-strain curve, an analogous procedure is performed. The mean mr


and standard deviation slog,r are calculated and compared against the values for an
unbiased estimator, for which mr and slog,r/slog,r,pre should be equal to 1 on aver-
age. It should be noted that, in practice, estimators that have a slight bias are often
considered acceptable. This bias metric is the first of two criteria used to assess the
quality of the regression methods.
The second quality criterion used is the goodness of estimation. It is defined
following Müller [27], who uses a similar quantity. It is a criterion related to the ex-
tent to which a single estimate may differ from the population (Eq 28):
10 %  quantileðXÞ
goodness of estimationðX Þ ¼ GoEðX Þ ¼ : (28)
90 %  quantileðXÞ
174 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 14 Determination of the mean values Nexp,m (a) and rexp,m (b) from the curves
determined by performing a regression on simulated individual tests.

N N
N

This quantity describes the range between the 10 % and 90 % quantiles, which
contains 80 % of all simulated values. It does not depend on any particular type of
distribution. A good estimator should possess a GoE that is as large as possible. The
GoE can take a maximum value of 1, which indicates that the estimation will always
yield the same value.
Fig. 15 schematically illustrates the performance of a good and a bad estimator
in terms of the quality criteria of bias and goodness of estimation.
To determine a reasonable number of repetitions of the simulation after which
stable values can be obtained for the bias and goodness of estimation, a sample size
of n ¼ 8 is initially considered. Many repetitions are performed, and for each of
these repetitions, the bias and goodness of estimation are calculated for all previous

FIG. 15 Schematic illustration of the characteristics of (a) a good estimator and (b) a
bad estimator in terms of bias and goodness of estimation.

GoE large GoE

x x
WÄCHTER AND ESDERTS, DOI 10.1520/STP159820160050 175

FIG. 16 GoE(slog,N/slog,N,pre) for different numbers of repetitions of the simulation, n ¼ 8.

0.35
GoE

repetitions. As an example, the GoE(slog,N/slog,N,pre) results for the three considered


methods of regression are shown in Fig. 16. As seen in this figure, after a number of
repetitions (between approximately 104 and 105), no further appreciable variations
in the considered criterion are apparent. Similar results are found for the other cri-
teria. Therefore, 5  104 is selected as a reasonable number of repetitions.
The quality criteria are determined for various sample sizes n ¼ 8, …, 20, where
the sample size refers to the number of specimens at the beginning of one test se-
ries. However, the actual sample size considered in the evaluation may be smaller.
On the one hand, it is common for run-outs to occur at low strain levels; on the
other hand, it is also possible (although less probable) that tests performed at high
strain levels may lead to numbers of cycles smaller than 1. In both cases, these tests
must be excluded from the regression, causing the real sample size to be smaller
than the number of simulated tests. In the following discussions, the sample size is
always assumed to be equal to the number of simulated tests. Furthermore, it may
occur that an entire test series cannot be evaluated at all, either because too few sin-
gle tests are available for the regression or because implausible values are deter-
mined for the cyclic properties (e.g., positive values for the exponents b and c). In
such cases, these repetitions are not used to determine the bias and the goodness of
estimation. The number of repetitions may therefore not be equal for all methods
of regression. However, in this evaluation, the quantities of such unevaluable test se-
ries are small compared with the absolute number of repetitions (less than 1 %; see
following section). Therefore, these discrepancies are not considered a problem.
176 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

SIMULATION RESULTS
The following figures show the quality criteria of bias and goodness of estimation
for the various parameters in question plotted versus the sample size. Figs. 17
through 20 show these two quality criteria for the following quantities:
• For the strain-life curve:
*
Fig. 17: Bias and GoE for the logarithmic mean of the cycles to failure for
the strain-life curve, Eq 29:
P
n
1
n logðNexp;m =Npre Þ
mN ¼ 10 i¼1 (29)

*
Fig. 18: Bias and GoE for the standard deviation slog,N/slog,N,pre of the num-
ber of cycles for the strain-life curve.
• For the stress-strain curve:
*
Fig. 19: Bias and GoE for the logarithmic mean of the stress amplitude for
the stress-strain curve, Eq 30:
P
n
1
n logðra;exp;m =ra;pre Þ
mr ¼ 10 i¼1 (30)

*
Fig. 20: Bias and GoE for the standard deviation slog,r/slog, r, pre of the stress
amplitude from the stress-strain curve.

FIG. 17 Bias and goodness of estimation for the mean of Nexp,m/Npre.


m
GoE m

n
WÄCHTER AND ESDERTS, DOI 10.1520/STP159820160050 177

FIG. 18 Bias and goodness of estimation for slog,N/slog,N,pre, uncorrected values.

GoE

FIG. 19 Bias and goodness of estimation for the mean of rexp,m/rpre.


m
GoE m

n
178 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 20 Bias and goodness of estimation for slog,r/slog,r,pre, uncorrected values.

GoE

The strain-life curve is considered first, as shown in Fig. 17. The mean number of
cycles is estimated with adequate accuracy. The deviations from the population are
less than 2 %. The bias of the mean shows no notable dependence on the sample size.
This is not the case for the goodness of estimation, however, which improves with an
increasing number of samples. The GoEs for all three regressions do not differ notably.
The uncorrected standard deviation of the strain-life curve is slightly underesti-
mated by all three methods (e.g., 7 % at n ¼ 8). Again, the GoEs for the three meth-
ods increase with the sample size but do not differ noticeable for the three
regressions, Fig. 18.
For the cyclic stress-strain curve, the following observations can be made: The
mean stress amplitude is estimated in an unbiased manner by all three regression
methods, and the GoEs of the three methods are virtually equal and on a very high
level compared to the mean in the strain-life curve, as seen in Fig. 19. Regarding the
uncorrected standard deviation, an underestimation of the same magnitude as that
for the strain-life curve is observed at small sample sizes, Fig. 20. No noticeable dif-
ferences in the GoE are evident among the three methods. As already mentioned
earlier, it is known that the estimator for the standard deviation, Eqs 22 and 23, is
biased for small sample sizes. This can be corrected by applying Eq 24. When this
correction is applied to the results of the simple linear regression and the Deming
regression, the bias can be restored to an acceptable value, Fig. 21.
From Fig. 22, it is evident that the number of test series that cannot be evaluated
is very small for all three methods. On the one hand, a test series may become
WÄCHTER AND ESDERTS, DOI 10.1520/STP159820160050 179

FIG. 21 Bias of the standard deviations after being corrected using Eq 24.

unevaluable if too many single test results must be excluded from the evaluation be-
cause of excessively small plastic strain components or run-outs. These phenomena
affect all regression methods equally. On the other hand, a test series may also be
deemed unevaluable because the regression yields implausible cyclic properties,
such as negative values of the exponents b and c. This may be caused by the

FIG. 22 Percentage of unevaluable test series.

n
180 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

regression itself or by the simulated test series. In all cases, the overall numbers of
such unevaluable test series are below 1 %.

DERIVING A CONFIDENCE LEVEL OF 10 % OR 90 % FOR THE ESTIMATION


The plots for the goodness of estimation in Figs. 17 through 20 can be used to derive
the confidence level of 10 % and 90 % for the estimated parameters. The goodness
of estimation is the ratio of the 10 % and the 90 % quantile of the simulation results.
Therefore, 80 % of all values will appear inside the range of the inverse of the good-
1
ness of estimation GoE . For the log-normal distribution, which is symmetric in loga-
rithmic scale, the distance between the 10 % quantile and the mean or the 90 %
1 1
quantile and the mean is the square root of GoE . Therefore, the square root of GoE can
be used as a factor j to shift the estimate to a confidence level of 10 % or 90 % (Eq 31):
rffiffiffiffiffiffiffiffiffi
1
j¼ (31)
GoE

If this factor is applied to the estimates to determine the values for a confidence
level of 10 % for the means and for a confidence level of 90 % for the standard devi-
ations, just 10 % of the possible estimates will be unsafe.
To clarify the application of this factor, two examples are provided:
1. Example 1: The mean number of cycles at a certain strain amplitude of a steel mate-
rial is evaluated by using the results of n ¼ 8 specimens. It shall be valid for a proba-
bility of occurrence of 10 % (no overestimation in 90 % of all cases). The GoE of the
mean of N in the strain-life curve can be determined from Fig. 17: GoE(n ¼ 8) ¼ 0.7.
Therefore, the factor for the confidence level of 10 % is 1.20 (Eq 32):
rffiffiffiffiffiffi
1 pffiffiffiffiffiffiffiffiffi
j¼ ¼ 1:43¼ 1:20: (32)
0:7

This can be applied to an estimated mean fatigue life Nm from the strain-life
curve as shown in Eq 33:

Nm
N m;90% ¼ (33)
j

2. Example 2: The cyclic properties of a steel material are evaluated by using the
results of n ¼ 13 specimens. The standard deviation of the stress amplitude in
the cyclic stress-strain curve is estimated to be slog,r ¼ 0.0140 by Eq 23. First,
this value must be corrected by Eq 24 (see Eq 34):

n  1:74 13  1:74
slog;r;corr ¼ slog;r  ¼ 0:0140  ¼ 0:0143: (34)
n2 13  2

The GoE of the standard deviation of ra in the stress-strain curve can be deter-
mined from Fig. 20: GoE(n ¼ 13) ¼ 0.5. Therefore, the factor for a confidence level
of 90 % is 1.41 (Eq 35).
WÄCHTER AND ESDERTS, DOI 10.1520/STP159820160050 181

TABLE 2 Comparison of the GoE and j factors for the estimation of the examined parameters.

n¼8 n ¼ 20

Estimated Quantity X GoE(X) j GoE(X) j

mN 0.70 1.20 0.80 1.12


slog,N 0.37 1.64 0.59 1.30
mr 0.96 1.02 0.97 1.02
slog,r 0.37 1.64 0.59 1.30

rffiffiffiffiffiffi
1 pffiffiffi
j¼ ¼ 2 ¼ 1:41: (35)
0:5

This can be applied to the standard deviation of the stress-strain curve, as


shown in Eq 36.

slog;r;90% ¼ slog;r;corr  j ¼ 0:0143  1:41 ¼ 0:0201 (36)

Conclusion
It is shown how strain-controlled tests that are used to estimate strain-life and
stress-strain curves and also the standard deviations for these curves can be simulat-
ed for steel materials. With the results of these simulations, conclusions can be de-
rived for sample sizes between 8 and 20 specimens.
No significant differences can be found for the strain-life curve or the cyclic
stress-strain curve for values of the quality criteria of the three examined regression
methods. The estimated mean of the number of cycles in the strain-life curve as
well as the estimated mean of the stress amplitude in the stress-strain curve are un-
biased. For the standard deviations of both curves, this is not the case. The standard
deviations are slightly underestimated for small sample sizes, which can be cor-
rected by applying a known correction formula (Eq 24). Furthermore, the mean
number of cycles in the strain-life curve is harder to estimate compared to the
mean of the stress amplitude in the stress-strain curve, which shows an excellent
GoE. The GoE for the standard deviations in the strain-life curve and the stress-
strain curve are almost the same for each sample size and are far worse compared
to the ones for the means of both curves.
Based on these findings, none of the examined regressions can be preferred due
to their performance in estimating the cyclic properties. Therefore, it is justified to
use the simple linear regression because its formulae are rather simple compared to
the alternatives. However, these findings are just valid if the Coffin/Manson and
Ramberg/Osgood approaches are valid for the material in question. In other cases,
the Deming or spatial regression could lead to more accurate results.
182 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

Furthermore, it is shown how to derive factors j to shift the estimated parame-


ters to a confidence level of 10 % or 90 % to get safe estimates. Table 2 compares the
GoE and factors j calculated therefrom for the sample sizes n ¼ 8 and n ¼ 20. It can
be seen that the necessary factors become lower for bigger sample sizes—except for
the mean of the stress amplitude in the cyclic stress-strain curve, which shows an al-
most constant GoE over the examined range on n.
It can also be seen that, even with 20 specimens, the GoE of the estimation of
the standard deviations is still worse than that of the means at n ¼ 8 and that the
estimation of the standard deviations is a rather challenging problem.

ACKNOWLEDGMENTS
Parts of this work were funded by the German Federation of Industrial Research Asso-
ciations (AiF) through Project 17612 of the Research Association for Mechanical
Engineering (FKM). The authors would like to thank AiF and FKM for their support.
The authors also wish to express their gratitude to Dr.-Ing. Christian Müller and
Dr.-Ing. Karsten Hinkelmann for several intense discussions on the evaluation of the
standard deviation with respect to the strain-life curve.

References

[1] ASTM E606-04, Standard Practice for Strain-Controlled Fatigue Testing, ASTM Interna-
tional, West Conshohocken, PA, 2010, www.astm.org

[2] ASTM E739-80, Standard Practice for Statistical Analysis of Linear or Linearized Stress-
Life (S-N) and Strain-Life (e-N) Fatigue Data, ASTM International, West Conshohocken,
PA, 2010, www.astm.org

[3] BS 7270:2006, Metallic Materials. Constant Amplitude Strain Controlled Axial Fatigue.
Method of Test, British Standards Institution, London, 2006.

[4] ISO 12106:2003, Metallic Materials—Fatigue Testing—Axial-Strain-Controlled Method,


International Organization for Standardization, Geneva, Switzerland, 2003.

[5] ASD-STAN prEN 3988-1998, Aerospace Series Test Methods for Metallic Materials Constant
Amplitude Strain-Controlled Low Cycle Fatigue Testing, CEN, Brussels, Belgium, 1998.

[6] SEP 1240, Testing and Documentation Guideline for the Experimental Determination of
Mechanical Properties of Steel Sheets for CAE-Calculations, VDEh, Düsseldorf, Germany,
2006.

[7] Bäumel, A. and Seeger, T., Materials Data for Cyclic Loading, Supplement 1, Elsevier
Science, Amsterdam, 1990.

[8] Nieslony, A., el Dsoki, C., Kaufmann, H., and Krug, P., “New Method for Evaluation of the
Manson–Coffin–Basquin and Ramberg–Osgood Equations with Respect to
Compatibility,” Int. J. Fatigue, Vol. 30, No. 10, 2008, pp. 1967–1977, http://dx.doi.org/
10.1016/j.ijfatigue.2008.01.012
WÄCHTER AND ESDERTS, DOI 10.1520/STP159820160050 183

[9] Coffin, L. F., Jr., Ed., “A Study of the Effects of Cyclic Thermal Stresses on a Ductile
Metal,” Trans. ASME, Vol. 76, 1954, pp. 931–950.

[10] Manson, S. S., “Fatigue: A Complex Subject—Some Simple Approximations,” Exp. Mech.,
Vol. 5, No. 7, 1965, pp. 193–226.

[11] Campbell, F. C., Ed., Elements of Metallurgy and Engineering Alloys, ASTM International,
West Conshohocken, PA, 2008, www.astm.org

[12] Ramberg, W. and Osgood, W. R., Description of Stress-Strain Curves by Three Parameters,
Technical Note No. 902, National Advisory Committee for Aeronautics, Washington, DC, 1943.

[13] Fatemi, A., Plaseied, A., Khosrovaneh, A. K., and Tanner, D., “Application of Bi-linear
Log–Log S–N Model to Strain-Controlled Fatigue Data of Aluminum Alloys and Its Effect
on Life Predictions,” Int. J. Fatigue, Vol. 27, No. 9, 2005, pp. 1040–1050.

[14] Masendorf, R., Einfluss der Umformung auf die zyklischen Werkstoffkennwerte von
Feinblech, PhD dissertation, Clausthal University of Technology, Clausthal-Zellerfeld,
Germany, 2000.

[15] Kandil, F. A., “Potential Ambiguity in the Determination of the Plastic Strain Range Com-
ponent in LCF Testing,” Int. J. Fatigue, Vol. 21, No. 10, 1999, pp. 1013–1018.

[16] Deming, W. E., Statistical Adjustment of Data, Wiley, New York, 1943 (Dover Publications
Edition, Mineola, NY, 1985).

[17] Glaister, P., “Least Squares Revisited,” Math. Gaz., Vol 85, No. 502, 2001, pp. 104–107.

[18] Jacquelin, J., “3-D Linear Regression,” https://de.scribd.com/doc/31477970/Regres-


sions-et-trajectoires-3D (accessed February 17, 2016).

[19] Martin, A., Hinkelmann, K., and Esderts, A., “Zur Auswertung von Schwingfestigkeitsver-
suchen im Zeitfestigkeitsbereich—Teil 2: Wie Zuverlässig Kann die Standardabweichung
aus Experimentellen Daten Geschätzt Werden?” Mater. Test., Vol. 53, No. 9, 2011,
pp. 502–512, http://dx.doi.org/10.3139/120.110256

[20] Hück, M., “Ein Verbessertes Verfahren für die Auswertung von Treppenstufenversuchen,”
Materialwissenschaft und Werkstofftechnik, Vol. 14, No. 12, 1983, pp. 406–417, http://
dx.doi.org/10.1002/mawe.19830141207

[21] Pollak, R., Palazotto, A., and Nicholas, T., “A Simulation-Based Investigation of the
Staircase Method for Fatigue Strength Testing,” Mech. Mater., Vol. 38, No. 12, 2006,
pp. 1170–1181, http://dx.doi.org/10.1016/j.mechmat.2005.12.005

[22] Martin, A., Hinkelmann, K., and Esderts, A., “Zur Auswertung von Schwingfestigkeitsver-
suchen im Zeitfestigkeitsbereich—Teil 1: Wie Zuverlässig Können 50 %-Wöhlerlinien
aus Experimentellen Daten Geschätzt Werden?” Mater. Test., Vol. 53, No. 9, 2011,
pp. 502–512, http://dx.doi.org/10.3139/120.110256

[23] Müller, C., Hinkelmann, K., Wächter, M., Masendorf, R., and Esderts, A., “Zur Wiederver-
wendung von Durchläufern im Treppenstufenversuch,” Mater. Test., Vol. 54, Nos. 11–12,
2012, pp. 786–792, http://dx.doi.org/10.3139/120.110390

[24] Müller, C., Hinkelmann, K., Masendorf, R., and Esderts, A., Zur Treffsicherheit der experi-
mentellen Dauerfestigkeitsschätzung, Technical Report Series Fac3-14-02, Clausthal Uni-
versity of Technology, Clausthal-Zellerfeld, Germany, 2014.
184 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

[25] Ellmer, F. and Eulitz, K.-G., “Vergleich verschiedener Verfahren zur Bestimmung von
Werkstoff- und Bauteildauerfestigkeiten hinsichtlich der zu erwartenden statistischen
Sicherheit der ermittelten Parameter,” 42, Tagung des DVM-Arbeitskreises Betriebsfes-
tigkeit. Betriebsfestigkeit—Bauteile und Systeme unter komplexer Belastung, Dresden,
Germany, October 7–8, 2015.

[26] Störzel, K., “Zuverlässigkeit bei der statistischen Auswertung von Schwingfestigkeitsversuchen,”
DVM-Workshop Zuverlässigkeit und Probabilistik, Munich, Germany, November 26–27, 2015.

[27] Müller, C., Zur statistischen Auswertung experimenteller Wöhlerlinien, PhD dissertation,
Clausthal University of Technology, Clausthal-Zellerfeld, Germany, 2015.

[28] Ahmadi-Dahaj, A., Lebensdauerabschätzung schwingend beanspruchter metallischer


Werkstoffe mittels Simulation des Mikrorisswachstums, PhD dissertation, Clausthal Uni-
versity of Technology, Clausthal-Zellerfeld, Germany, 2003.

[29] Boller, C. and Seeger, T., Materials Data for Cyclic Loading. Part A: Unalloyed Steels,
Elsevier, Amsterdam, 1987.

[30] Boller, C. and Seeger, T., Materials Data for Cyclic Loading. Part B: Low-Alloy Steels,
Elsevier, Amsterdam, 1987.

[31] Boller, C. and Seeger, T., Materials Data for Cyclic Loading. Part C: High-Alloy Steels,
Elsevier, Amsterdam, 1987.

[32] Boller, C. and Seeger, T., Materials Data for Cyclic Loading. Part E: Cast and Welded
Metals, Elsevier, Amsterdam, 1987.

[33] Bork, C., Huenecke, J., and Golmann, M., Neue Stähle mit hoher statischer, dynamischer
und Dauerfestigkeit für den Automobilbau, Report on Research Project 03 M 3021,
German Ministry of Education and Research, Bremerhaven, 1995.

[34] Traupe, M., Meinen, H., and Zenner, H., Sichere und Wirtschaftliche Auslegung von Eisen-
bahnfahrwerken, Report on Research Project 19 P 0061 A to F, German Ministry of
Education and Research, Clausthal-Zellerfeld, 2004.

[35] Lütkepohl, K., Esderts, A., Luke, M., and Varfolomeev, I., Sicherer und Wirtschaftlicher Betrieb
von Eisenbahnfahrwerken, Report on Research Project 19 P 4021 A to F, German Ministry of
Economy, Clausthal-Zellerfeld, 2009, http://dx.doi.org/10.2314/GBV:614268583

[36] Zenner, H. and Thorms, V., Betriebsfestigkeit von umgeformten Karosseriestählen,


FAT-Schriftenreihe No. 166, FAT, Frankfurt, 2001.

[37] Wagener, R. and Schatz, M., Leichtbau mit Hilfe von zyklischen Werkstoffkennwerten für
Strukturen aus umgeformtem höherfesten Feinblech. FAT-Schriftenreihe No. 191, FAT,
Frankfurt, 2005.

[38] Medhurst, T. and Süße, D., Nutzung des Leichtbaupotentials von höherfesten Stahlfein-
blechen durch die Berücksichtigung von Fertigungseinflüssen auf die Festigkeitseigen-
schaften, FAT Schriftenreihe No. 242, FAT, Frankfurt, 2012.

[39] Harste, D., Untersuchung zur Auswirkung von Überlasten auf die Dauerfestigkeit, PhD
dissertation, Clausthal University of Technology, Clausthal-Zellerfeld, Germany, 1996.

[40] Hollmann, C., Übertragbarkeit von Schwingfestigkeits-Eigenschaften im Örtlichen Kon-


zept, PhD dissertation, Universität Dresden, Dresden, Germany, 2004.
WÄCHTER AND ESDERTS, DOI 10.1520/STP159820160050 185

[41] Medhurst, T., Zyklisches Verhalten Metastabiler Austenitischer Feinbleche in Abhängig-


keit des Umformgrades, PhD dissertation, Clausthal University of Technology, Clausthal-
Zellerfeld, Germany, 2014.

[42] Society of Automotive Engineers (SAE): Fatigue Design and Evaluation Committee,
Experimental HTML Fatigue Database, http://fde.uwaterloo.ca/Fde/Materials/dindex.
html (accessed September 4, 2013).

[43] Engl, B., Steinbeck, G., and Nicklas, D., Erarbeitung werkstoff- und verarbeitungsger-
echter Kennwerte für Feinblech aus normal- und höherfesten sowie nichtrostenden Stäh-
len, Report, VDEh, Düsseldorf, 2001.

[44] Bäumer, A., May, U., Steinbeck, G., and Bork, C.-P., Ermittlung des Werkstoffverhaltens
und des Beschichtungseinflusses durch rechnerische Methoden zur Verkürzung der
Entwicklungszeiten im Fahrzeugbau mit Stahl, Report, VDEh, Düsseldorf, 2006.

[45] Bork, C.-P., Reichert, B., Wagener, R., Dahmen, K., Masendorf, R., Geffert, A., Gerlach, J.,
Menne, M., and Till, E., Validierung und Erweiterung von Berechnungsmethoden für die
Blechumformung, Betriebsfestigkeits- und Crashauslegung im Fahrzeugbau mit Stahl,
Report Project No. P 744, VDEh, Düsseldorf, 2013.

[46] Wächter, M., Esderts, A., and Masendorf, R., “Methoden zur Abschätzung zyklischer
Werkstoffkennwerte,” Tagungsband, 1. Niedersächsisches Symposium Materialtechnik,
Clausthal-Zellerfeld, February 12–13, 2015.

[47] Matsumoto, M. and Nishimura, T., “Mersenne Twister: A 623-Dimensionally Equidistrib-


uted Uniform Pseudo-Random Number Generator,” ACM Trans. Model. Comput. Simul.,
Vol. 8 No. 1, 1998, pp. 3–30.
186 FATIGUE AND FRACTURE TEST PLANNING, TEST DATA ACQUISITIONS AND ANALYSIS

STP 1598, 2017 / available online at www.astm.org / doi: 10.1520/STP159820160046

Bruce A. Young,1 Richard C. Rice,1 Steven R. Thompson,2


and Doug Hall3

Methods Development for


Nonlinear Analysis of Fatigue
Data
Citation
Young, B. A., Rice, R. C., Thompson, S. R., and Hall, D., “Methods Development for Nonlinear
Analysis of Fatigue Data,” Fatigue and Fracture Test Planning, Test Data Acquisitions and
Analysis, ASTM STP1598, Z. Wei, K. Nikbin, P. C. McKeighan, and D. G. Harlow, Eds., ASTM
International, West Conshohocken, PA, 2017, pp. 186–197, http://dx.doi.org/10.1520/
STP1598201600464

Abstract
Currently, ASTM E739-10, Standard Practice for Statistical Analysis of Linear or
Linearized Stress-Life (S-N) and Strain-Life (e-N) Fatigue Data, provides analysis
methods to determine confidence limits on the mean of linearized fatigue life
data. Although these techniques are sufficient for the casual user, rigorous data
analysis for structural applications typically involves the calculation of tolerance
bounds for nonlinear data sets to ensure design criteria. The effect of
nonlinearity becomes more significant as life predictions become significantly
longer approaching the material fatigue limit of certain combinations of
materials and test conditions (maximum stress, R-ratio, Kt, etc.). While analyzing
numerous fatigue data sets, several issues have been seen that do not currently
fit within the limitations associated with ASTM E739-10. These limitations
include: (1) The treatment of data sets containing nonuniform variance over a
range of stress levels needs to be considered. (2) Nonsymmetric distributions in
non-transformed space (e.g., two-parameter Weibull or log-normal) may
describe the failure distributions better for given stress levels and materials than
a linear failure model in non-transformed space. (3) Data taken over a whole

Manuscript received March 21, 2016; accepted for publication September 15, 2016.
1
Battelle Memorial Institute, 505 King Ave., Columbus, OH 43201
2
Air Force Research Laboratory, Wright Patterson AFB, OH
3
Honeywell Aerospace, South Bend, IN
4
ASTM Symposium on Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis on May 4–5,
2016 in Grand Hyatt, San Antonio, TX.

Copyright V
C 2017 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.
YOUNG ET AL., DOI 10.1520/STP159820160046 187

range of stress levels may provide unrealistic tolerance bounds for low stresses
when compared to the development of tolerance bounds on a stress-level by
stress-level basis. This study addresses some of these limitations using numerical
simulation-based statistical distributions obtained from several relatively large
data sets. These numerical simulations are analyzed using the linear methods
shown in ASTM E739-10 as well as other techniques such as those outlined in the
Metallic Materials Properties Development and Standardization (MMPDS)
Handbook to develop confidence and tolerance limits.

Keywords
fatigue, statistical, nonlinear, analysis, data, ASTM E739

Nomenclature
b Intercept value (log cycles)
COV Coefficient of variation (standard deviation divided by mean)
Nf Cycles to failure
NRO Predefined runout cycles for testing
m Slope value
Kt Notch-concentration value
R Stress ratio (minimum stress divided by the maximum stress)
r Stress (ksi or MPa)
rf Best estimate of fatigue strength at runout cycles (ksi or MPa)
T99 Statistically computed, one-sided lower tolerance limit, representing a 95 %
confidence lower limit on the first percentile of the distribution
T90 Statistically computed, one-sided lower tolerance limit, representing a 95 %
lower confidence limit on the tenth percentile of the distribution

Introduction
The subcommittee responsible for ASTM E739-10, Standard Practice for Statistical
Analysis of Linear or Linearized Stress-Life (S-N) and Strain-Life (e-N) Fatigue Data
[1], is currently undertaking an effort to revise the standard to update the methods of
analyses. As revisions to ASTM E739-10 [1] proceed, nonlinear analysis methods
should be considered as part of the suite of analysis options to determine both confi-
dence limits and tolerance limits of fatigue life data. As shown by Rice, Young, and
Thompson [2], the effect of nonlinearity becomes more significant as life predictions
become significantly longer approaching the material fatigue limit of certain combi-
nations of materials and test conditions. In addition, the accuracy of the prediction of
the fatigue strength at the runout cycles is also dependent on the number of stress
ranges tested, the number of specimens tested per stress range, and the variation in
the cycles to failure near the fatigue threshold. Several models have been put forth in
the literature to model high cycle fatigue stress-life behavior. Two linear or linearized
models currently put forth in ASTM E739-10 are provided in Eqs 1 and 2.
188 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

 
log Nf ¼ b þ mr (1)
 
log Nf ¼ b þ m½log ðrÞ (2)

The model shown in Metallic Materials Properties Development and Standardization


(MMPDS)-09 [3] provides guidance for a nonlinear model for stress-life fatigue.
 
log Nf ¼ b þ m½log ðr þ AÞ (3)

Based on work completed [2], the best representation for a majority of the data
sets considered was a distribution of cycles to failure at a given stress as a log-
normal distribution. Based on that assumption, and a nonlinear model such as
Eq 3, Eq 4 is the baseline nonlinear model for the scope of this paper. Eq 4 takes the
nonlinearity of Eq 3 and develops meaningful relationships between the runout
cycle count and the fatigue strength at the runout cycles. It should be noted that
although Eq 3 has three degrees of freedom, by declaring the dependencies between
the runout cycles and the fatigue strength at the runout cycles, the number of
degrees of freedom are reduced to two for Eq 4. As shown in the nomenclature sec-
tion, rf is the fatigue strength at the runout cycles (ksi or MPa), which is the stress
at the predefined runout cycle level. Typically, the runout cycle level is set to the
typical endurance limit cycle levels for various alloys (i.e., 107 for steels, 3  107 for
aluminum, and 109 for titanium).
    
log Nf ¼ log ðNRO Þ  m log r  rf þ 1 (4)

As shown in Fig. 1, linear and nonlinear representations of using the input in


Table 1 provide very different representations of the data. This representation leads to
a very different conclusion about the underlying behavior such as a material fatigue
limit (or not) for a given material. Results in Fig. 1 are plotted using Eqs 2 and 4.
Data sets were chosen to provide a wide range of characteristics. Of those char-
acteristics, one in particular was chosen to elicit the variation in data sets between
homoscedastic and non-homoscedastic. That is, for homoscedastic data sets, the
coefficient of variation (COV) for the statistical distribution of cycles to failure in
log-normal space is the same for all stress values.
Other data sets were chosen to compare the results of the fatigue strength at
the runout cycles for variation in number of stress levels per data set, the number of
tests specimens per stress level, and a variation in the slope.

Assumptions
Several assumptions were made during the course of the analyses presented within
this paper. The guiding assumptions are provided in the following list:
• Analyses were completed in stress space such that the stress is considered to
be the maximum stress during testing. The authors believe that similar analy-
ses could be completed in strain, pseudo-strain, or plastic-strain range spaces
with controlling parameters of mean value or alternating value (i.e., mean
stress or alternating stress).
YOUNG ET AL., DOI 10.1520/STP159820160046 189

FIG. 1 Comparison of linear and nonlinear models in log-log space (for visualization
only, not for a specific data set).

1.9

1.8
Linear (Eq. 2)

1.7 Non-Linear (Eq. 4)


Log10(Stress)

1.6

1.5

1.4

1.3
4 4.5 5 5.5 6 6.5 7 7.5 8
Log10(Cycles to Failure)

• The data sets contain only a single stress ratio (R)


• The data sets contain only a single-notch concentration value (i.e., Kt ¼ 1.0 for
smooth-bars or Kt ¼ 3.0…etc. for some notch configurations, but not mixed in
the same data set).
• The runout cycle count is considered to be 10,000,000 cycles.
• The fatigue strength at the runout cycles corresponds to the mean value of
stress at the predefined runout stress during testing.
• For these analyses, the failures occur at the values of the cycles determined
from the random draw of a normal distribution in log space. No consideration
of runouts as a maximum likelihood methodology is considered for this paper
but should be considered for laboratory fatigue data whereby the tests are dis-
continued at the runout cycles.

TABLE 1 Parameters used to show differences in models.

Variable Values

Stress Level Number 1 67.5


Mean Cycles at Stress Level Number 1 100,000
Stress Level Number 2 22.5
Mean Cycles at Stress Level Number 2 10,000,000
Runout Cycles during Testing 10,000,000
Fatigue Strength at Runout Cycles 22.5
190 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

STATISTICAL DATA SETS


As discussed earlier, the basis of this study is to address some of the following issues
through numerical simulation. The issues include:
• The treatment of data sets containing nonuniform variance over a range of
stress levels.
• Nonsymmetric distributions (e.g., log-normal).
• Data taken over a whole range of stress levels may provide unrealistic toler-
ance bounds for low stresses when compared to the development of tolerance
bounds on a stress-level by stress-level basis.
The parameters and the values of those parameters used for the basis of the
numerical simulation work are provided in Table 2. The COV for these cases is
defined in log space as the standard deviation in cycles to failure divided by the
mean cycles to failure at a given stress level expressed as a percent.

Analysis Results
Although the assumption is that the underlying behavior of the cycles to failure at a
given stress level has a log-normal distribution, the numerical simulations were an-
alyzed using both the linear methods shown in ASTM E739-10 as well as the non-
linear log-normal method shown in Eq 1. A general form of the log-normal method
is provided in Chapter 9 of the MMPDS Handbook [3].

REVIEW OF THE SENSITIVITY ANALYSIS


Because fatigue strength at runout stress is one of the focus areas of this paper,
Fig. 2 was developed to compare the variability in fatigue strength at the run-
out stress as a function of variability in COV of the cycles to failure within the
data set using the parameters shown in Table 2. Based on the assumptions of the
log-normal distribution of cycles to failure and an infinite data set, a COV of

TABLE 2 Parameters used for the sensitivity study.

Variable Values

Stress Levels per Data Set 5


Specimens per Stress Level 3 or 5
Cycles to Failure Distribution Log-Normal
Yield Stress 75
Highest Stress: Mean Cycles to Failure 100,000
Highest Stress: Percent yield strength (YS) (%) 90
Highest Stress: COV Cycles to Failure (%) 2 or 5
Lowest Stress: Mean Cycles to Failure 10,000,000
Lowest Stress: Percent YS (%) 30
Lowest Stress: COV Cycles to Failure (%) 2, 5, 10, or 13.7**
Simulations per data set 1,000

Note: **COV at the lowest stress was always greater than or equal to the COV at the highest stress.
YOUNG ET AL., DOI 10.1520/STP159820160046 191

FIG. 2 Comparison of COV for fatigue strength at the runout cycles (in normal space)
as a function of COV of cycles to failure (in log space).

12.0

Specimens per Level = 3 // COV(Nf=1E5) = 2


10.0
COV(Fague Stength at Runout Cycles)

Specimens per Level = 3 // COV(Nf=1E5) = 5

Specimens per Level = 5 // COV(Nf=1E5) = 2


8.0
Specimens per Level = 5 // COV(Nf=1E5) = 5
6.0

4.0

2.0

0.0
0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0 16.0
COV(Nf at Runout)

2 (in log space) at a mean cycle count of 100,000 provides the same lower toler-
ance limit (T99) value for a mean of 10,000,000 with a COV of 13.7.

CONFIDENCE AND TOLERANCE LIMITS


Both confidence limits (i.e., limits on how confident the mean curve fits the data)
and tolerance limits (i.e., with how much certainty the next data point will fall with-
in some bounds of the data set) are well-characterized for linear regression. How-
ever, for nonlinear regression (which is not polynomial in nature), confidence and
tolerance limits are much less characterized in analytical forms. Thus, to continue
the discussion for confidence and tolerance limits, Eq 4 will be recast using Eqs 5, 6,
and 7, into Eq 8. Thus, the equation is linearized after finding the estimated fatigue
strength at the runout cycles, rf :

Y ¼ mX þ B (5)
where
  
X ¼  log r  rf þ 1 (6)
 
Y ¼ log Nf (7)
B ¼ log ðNRO Þ: (8)

Information in the way of statistical characterizations of a given data set is


required to complete both confidence and tolerance limit calculations. Eq 9 pro-
vides the information to calculate the standard deviation, sY, using the information
192 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

obtained from Eq 8. To obtain the lower confidence limit, Eq 10 should be used


and can be found in Section 9.8.2 of the MMPDS Handbook [3]. The t-statistic for
95 % confidence and varying degrees of freedom can be found in Table 9.10.4 of the
handbook [3]. Given the data in Table 3, Eq 4 through 8 were used to analyze the
data and create the confidence limit shown in Figs. 3 and 4. Similar to the confi-
dence limit, Eq 11 can be used to calculate the lower tolerance bound. Section
9.5.6.1 in the MMPDS Handbook [3] provides the guidance for calculating the tol-
erance bound including reference to the k- and t-tables.
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
PðYi Y i Þ2
sY ¼ (9)
n2
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u  P 2
u
u1 X  Xi=n
YLC ¼ mX þ B  t1a;df sY u
tn þ P P 2. (10)
ð Xi2 Þ  ð Xi Þ n

YLT ¼ mX þ B  k0 sY (11)
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
0  ð1 þ D Þ
k ¼t (12)
n
0  P 2 1
X  Xi=
B n C
D¼B C (13)
B 
BP P 2, C C
@ Xi  Xi n A
n

TABLE 3 Data used for confidence and tolerance limit analysis.

Stress Cycles rf X Y B M

22.50 32,981,179 22.88 0.21 7.518 7.00 1.245


22.50 12,034,853 22.88 0.21 7.080 7.00 1.245
22.50 15,575,249 22.88 0.21 7.192 7.00 1.245
29.61 357,686 22.88 0.89 5.554 7.00 1.245
29.61 376,983 22.88 0.89 5.576 7.00 1.245
29.61 2,436,226 22.88 0.89 6.387 7.00 1.245
38.97 35,996 22.88 1.23 4.556 7.00 1.245
38.97 648,604 22.88 1.23 5.812 7.00 1.245
38.97 780,439 22.88 1.23 5.892 7.00 1.245
51.29 233,728 22.88 1.47 5.369 7.00 1.245
51.29 61,871 22.88 1.47 4.791 7.00 1.245
51.29 148,275 22.88 1.47 5.171 7.00 1.245
67.50 128,593 22.88 1.66 5.109 7.00 1.245
67.50 120,451 22.88 1.66 5.081 7.00 1.245
67.50 92,843 22.88 1.66 4.968 7.00 1.245
YOUNG ET AL., DOI 10.1520/STP159820160046 193

where:
Yi ¼ individual log(cycle counts) at stress level i,
Y i ¼ average log(cycle counts) at stress level i,
n ¼ total number of data point,
YLC ¼ lower confidence limit,
Xi ¼ individual X values as defined in Eq 5,
t1a;df ¼ t-statistic with a being the level of confidence and df being the degrees
of freedom, and
t* ¼ the non-central t-distribution with non-central t-parameter and n-2
degrees of freedom.
The non-central t-parameter is the k value for an infinite population divided by
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ð1 þ DÞ=n. For the example provided herein, the degrees of freedom will be 13,
and the infinite population k will be 2.326.
Figs. 3, 4, and 5 were developed using a random selection of data with a nonuni-
form theoretical COV, such that the theoretical COV for the cycles at the runout
cycles was 13.7 and the theoretical COV at the highest stress level was 2. Actual
COV values for the data set at any given stress level ranged from 3 to 14. Using the
aforementioned mean values of the data set in Table 3 and requiring the data at each
stress level to average to near the mean value along with the requirement that the
data set have a uniform COV of approximately 5, Figs. 6 and 7 were constructed to
demonstrate the improvements in the confidence and tolerance bounds that can be
realized when a reasonable, uniform COV value is obtained. Typical fatigue data

FIG. 3 Confidence limit and tolerance bounds example in transformed coordinates.

8.00

7.50

7.00

6.50

6.00

5.50
Y

Simulaon Data
5.00
Mean Curve
4.50
95 % Confidence
4.00
T95 Tolerance Bound
3.50
T99 Tolerance Bound
3.00
-2.00 -1.50 -1.00 -0.50 0.00 0.50
X
194 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 4 Confidence limit and tolerance bounds example in stress versus cycles space.

70.00
Simulaon Data

60.00 Mean Curve

95 % Confidence Limit

50.00 T95 Tolerance Bound


Stress

T99 Tolerance Bound


40.00

30.00

20.00
1.0E+03 1.0E+04 1.0E+05 1.0E+06 1.0E+07 1.0E+08
Cycles to Failure

sets do not have a uniform COV, thus penalties on tolerance bounds are taken in
the areas of lower COV (on a per stress-level basis). This typically occurs for lower
cycle counts as can be seen in Fig. 4. Additional work needs to be considered, such
as weighting function based on residual analysis of standard deviation. This is
beyond the topic of this paper but is discussed in the “Future Work” section.

FIG. 5 Upper and lower confidence limits and tolerance bounds near the runout cycles
in stress versus cycles space.

30.00

29.00 Simulaon Data


Mean Curve
28.00
95 % Confidence Limit
27.00 T95 Tolerance Bound

26.00
Stress

25.00

24.00

23.00

22.00

21.00
1.0E+05 1.0E+06 1.0E+07 1.0E+08 1.0E+09
Cycles to Failure
YOUNG ET AL., DOI 10.1520/STP159820160046 195

FIG. 6 Example using a constant COV (in log space) of 5.0.

70.00

Simulaon Data
60.00
Mean Curve

95 % Confidence
50.00
T95 Tolerance Bound
Stress

T99 Tolerance Bound


40.00

30.00

20.00
1.0E+04 1.0E+05 1.0E+06 1.0E+07 1.0E+08
Cycles to Failure

Because Eqs 1 and 2 can easily be put into the form of Eq 8, similar analysis
procedures can be used to calculate their confidence and tolerance bounds. As
shown in Figs. 5 and 7, the confidence and tolerance bounds are based on the vari-
ability in cycles (the dependent variable in the analyses) not on the stress, thus the
confidence and tolerance limits are stopped at the mean stress associated with the
runout cycles (i.e., the fatigue strength at the runout cycles). Although the fatigue

FIG. 7 Upper and lower confidence limits and tolerance bounds near the runout cycles
in stress versus cycle space using a constant COV (in log space) of 5.0.

30.00

29.00 Simulaon Data


Mean Curve
28.00 95 % Confidence
T95 Tolerance Bound
27.00
Stress

26.00

25.00

24.00

23.00

22.00
1.0E+05 1.0E+06 1.0E+07 1.0E+08 1.0E+09
Cycles to Failure
196 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

strength at runout cycles can provide an estimate of the fatigue limit because of the
uncertainty associated with this value, a more systematic way of finding the material
fatigue limit is to use either the probit method or the staircase method [4].

Discussion and Conclusions


Currently, ASTM E739-10 has no method to complete analysis of fatigue data with
a fatigue strength at runout cycles component, nor does it contain a method of
determining tolerance limits for a given data set. Based on the information provided
in this paper, Eq 4 is capable of fitting data that has an apparent fatigue strength
component. Based on the results of the data analysis, and the assumption that the
variability comes from the variation in cycles to failure, the nonlinear equations can
be linearized and confidence and tolerance bounds on cycles can be established.
In addition to the methods developed in this paper for conducting analysis of
data with a fatigue strength component, the following conclusions should be con-
sidered when updates to ASTM E739-10 are investigated.
• The methods of analysis outlined in this paper do not require the COV to be
uniform over the range of data. However, as shown in Figs. 4 and 5, COV val-
ues of a data set provide insight into the magnitude of the confidence and
tolerance limits calculated for a given data set.
• COV values should be calculated at test stress levels to determine if the data
set has near-uniform variance. This requires multiple test results for a given
stress level. Although not a focus of this paper, the level of confidence in the
COV value increases with the number of tests conducted at a given stress level.
Therefore, a minimum number of tests per stress level should be investigated.
*
When using a nonweighted, nonuniform variance data set, large COV val-
ues may provide unrealistic confidence limits and tolerance bounds for
high stress levels.
*
Large COV values may provide insight into whether the tests are being
conducted under controlled conditions or whether random variables are
controlling the failure lives (i.e., surface roughness, inclusions near the sur-
face, misalignment of test fixtures, etc.).
*
Data sets examined by Rice, Young, and Thompson [2] show the tendency
for COV values to increase at longer lives (lower stresses).
• A determination based on the data set available is required to use either
linear or nonlinear analyses. A quantity such as the goodness-of-fit parameter
(r-squared value) of the fit using Eqs 1, 2, or 4 should be required.
• A value of 10 should be used as the upper-bound COV value for a valid data
set (on a stress level by stress level basis).
*
Confidence limit analyses require that the cycles-to-failure data near the
runout cycles have a COV on cycles of 13.7 % or less in log space.
*
To obtain a reasonable COV on the fatigue strength at the runout cycles
(rf), it has been shown (Fig. 2) that, for five stress levels with three tests at
each level, the COV (in log space) of the cycles to failure near the runout
cycles should be less than 10.
YOUNG ET AL., DOI 10.1520/STP159820160046 197

Future Work
Future work should consist of investigating the following items:
• Data requirements (i.e., number of specimens, number of stress levels, speci-
mens per stress level)
• Methods of maximum likelihood for the stress levels that contain runouts
• Methods of weighting nonuniform variance so as not to penalize low-variance
stress levels with any high-variance stress levels
• Methods of developing data sets using an equivalent stress model (i.e., Walker
model) for data sets with multiple r-ratios.

ACKNOWLEDGMENTS
The authors would like to thank the MMPDS membership for their support in the
development of this paper and their support in ASTM development efforts.

References

[1] ASTM E739-10, Standard Practice of Linear or Linearized Stress-Life and Strain-Life
Fatigue Data, ASTM International, West Conshocken, PA, 2015, www.astm.org

[2] Rice, R. C., Young, B., and Thompson, S., “Fatigue Analysis for Off-Nominal Conditions,”
presented at the Symposium on Fatigue and Fracture Test Planning, Test Data
Acquisition and Analysis, San Antonio, TX, May 4–5, 2016, ASTM International, West
Conshohocken, PA—unpublished.

[3] Metallic Materials Properties Development and Standardization (MMPDS), Version 10,
Battelle Memorial Institute, Columbus, OH, 2015.

[4] Little, R. E., Manual of Statistical Planning and Analysis ASTM STP588, ASTM Interna-
tional, West Conshocken, PA, 1975, http://dx.doi.org/10.1520/STP588-EB
198 FATIGUE AND FRACTURE TEST PLANNING, TEST DATA ACQUISITIONS AND ANALYSIS

STP 1598, 2017 / available online at www.astm.org / doi: 10.1520/STP159820160054

Hong-Tae Kang,1 Xiao Wu,1 Abolhassan K. Khosrovaneh,2


and Zhen Li 1

Data Processing Procedure for


Fatigue Life Prediction of Spot-
Welded Joints Using a Structural
Stress Method
Citation
Kang, H.-T., Wu, X., Khosrovaneh, A. K., and Li, Z., “Data Processing Procedure for Fatigue Life
Prediction of Spot-Welded Joints Using a Structural Stress Method,” Fatigue and Fracture Test
Planning, Test Data Acquisitions and Analysis, ASTM STP1598, Z. Wei, K. Nikbin, P. C.
McKeighan, and D. G. Harlow, Eds., ASTM International, West Conshohocken, PA, 2017, pp. 198–211,
http://dx.doi.org/10.1520/STP1598201600543

ABSTRACT
The structural stress method is widely used in fatigue-life prediction of various
welded joints in vehicle structures because this method is less dependent on mesh
sizes in finite-element models. However, there are several parameters that define the
structural stress, and extensive analysis is required to obtain an optimum structural-
stress versus fatigue-life curve of the joints. This paper establishes a data-processing
procedure for fatigue-life prediction of spot-welded joints using a structural stress
method. The structural stress method used in this study is influenced by sheet
thicknesses, joint dimensions, and stress types. These influencing parameters are
optimized by a nonlinear, generalized, reduced-gradient solver. Previous fatigue test
data from the Auto/Steel Partnership is reanalyzed using this proposed optimization
method, and the degree of scatter in the reproduced master curve is found to be
reduced compared to past results with default parameters. The estimated specimen
life using this procedure correlates well with the actual life observed in the
laboratory.

Manuscript received March 1, 2016; accepted for publication June 12, 2016.
1
The University of Michigan-Dearborn, Dept. of Mechanical Engineering, 4901 Evergreen Rd., Dearborn,
MI 48128
2
General Motors LLC, 30001 VanDyke Rd., Warren, MI 48090
3
ASTM Symposium on Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis on May 4–5,
2016 in Grand Hyatt, San Antonio, TX.

Copyright V
C 2017 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.
KANG ET AL., DOI 10.1520/STP159820160054 199

Keywords
spot weld, structural stress, fatigue-life prediction

Introduction
Electrical resistance spot welding has been widely used to join sheet metals in the au-
tomotive industry. A typical automobile may contain more than 3,000 to 5,000 spot
welds joining various body and structural components [1]. Fatigue failure, however,
is the most common failure type occurred in spot-welded joints. Therefore, it is im-
portant to investigate the fatigue-life prediction method for spot-welded joints. Many
researchers have proposed various fatigue-damage methods to predict the fatigue life
of spot-welded joints [1–8]. A commonly used fatigue-damage methodology for spot-
welded joints in automotive industry is the structural stress approach [1,4–8].
Compared to local stress methods, the structural stress method is less dependent
on mesh sizes of finite-element models. However, empirical factors are usually
included in the equations to consider the effect of different sheet thicknesses, joint
diameters, and stress types from different loading geometries, such as lap shear, coach
peel, or cross tension. These empirical factors are very dependent on the data includ-
ed. If the data are expanded or changed, the empirical factors must be determined
again according to the new database. Thus, it is necessary to have a scientific method
to find the optimized empirical factors when the databases expand or change.
In this study, a data processing procedure is established for fatigue-life prediction
of spot-welded joints using Rupp, Störzel, and Grubisic’s structural stress method [1].
The empirical factors in the structural stress equations are optimized by a nonlinear,
generalized, reduced-gradient solver. Finally, fatigue test data [9] from the Auto/Steel
Partnership (A/SP) is reanalyzed using this proposed optimization method, and the
results are compared with those obtained without optimized empirical factors.

Rupp and Coworkers’ Structural Stress Equation


The structural stress definition for spot-welded joints proposed by Rupp, Störzel,
and Grubisic [1] is used in this study. The structural stress is calculated from forces
and moments acting on a rigid kernel, as shown in Fig. 1. It is assumed that the size
of the rigid kernel is ten times smaller than the circular plate. The radial stresses
due to the forces and moments can be calculated as described by Rupp, Störzel, and
Grubisic [1].
This equation [1] has limitations when the database is expanded beyond the
initial database used in the study. Thus, a procedure has been developed to over-
come the shortcomings of the original equation [1]. The structural stress is calculat-
ed at a spot-welded joint using Eqs 1–6. The angle h in Eq 1 can be measured
counterclockwise from Fx.
   
rstru ¼ rmax ðFx Þ cos hrmax Fy sin h þ rðFz Þ þ rmax ðMx Þ sin hrmax My cos h
(1)
200 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 1 Circular plate model for a spot-welded specimen [1].

where
FX
rmax ðFX Þ ¼  SFFXY  d DEFXY  t TEFXY (2)
pdt
FY
rmax ðFY Þ ¼  SFFXY  d DEFXY  t TEFXY (3)
pdt
1:744FZ
rðFZ Þ ¼  SFFZ  dDEFX  t TEFZ if FZ > 0 or if FZ < 0 (4)
t2
1:872FX
rmax ðMX Þ ¼  SFMXY  dDEMXY  t TEMXY (5)
dt 2
1:872FY
rmax ðMY Þ ¼  SFMXY  dDEMXY  t TEMXY (6)
dt 2
and where t is the sheet thickness and d is the diameter of the weld nugget. SFFXY,
SFFZ, and SFMXY are scale factors used to adjust the contributions of the shear
forces, axial forces, and bending moments due to different loading geometries.
DEFXY, DEFZ, DEMXY, TEFXY, TEFZ, and TEMXY are the factors that allow a
size effect in terms of nugget diameter and sheet thickness to be taken into ac-
count. Therefore, there are nine discrete parameters affecting the final structural
stress results. The initially suggested parameter values [10] are shown in Table 1.

Finite-Element Models for Spot-Welded Joints


As mentioned previously, a typical vehicle structure contains 3,000 to 5,000 spot-
welded joints. Thus, it is almost impossible to represent those joints in a finite-element

TABLE 1 The initial empirical factor values for spot-welded joints made of steel.

Component Factor Diameter Exponent Thickness Exponent

FX, FY SFFXY ¼ 1.0 DEFXY ¼ 0.0 TEFXY ¼ 0.0


MX, MY SFMXY ¼ 0.6 DEMXY ¼ 0.0 TEMXY ¼ 0.5
FZ SFFZ ¼ 0.6 DEFZ ¼ 0.0 TEFZ ¼ 0.5
KANG ET AL., DOI 10.1520/STP159820160054 201

model with fine meshes. Instead of detailed finite-element models to evaluate the stress
states near the spot-welded joints, coarse finite-element models, as shown in Fig. 2, are
constructed to reduce computation time. The stress information from the coarse finite-
element model is not useful to evaluate fatigue characteristics of the joints because the
stress is dependent on the mesh density in the finite-element model. Thus, the nodal
forces and moments, which are less mesh-dependent, are utilized to calculate the struc-
tural stress to evaluate the fatigue characteristics of spot-welded joints.
Two-dimensional shell elements are used to represent both sheets, and the
spot-welded joint is modeled as a bar element with the same diameter of the weld
nugget. The dimensions of the finite element analysis (FEA) model are decided by
the average value of all the measured dimensions of each set. Boundary conditions
in the FEA models are replicated as the real test conditions. Both the sheets and the
nugget are assumed to have the same material property of steel.

Nonlinear, Generalized, Reduced-Gradient


Algorithm
To optimize the calculated structural stress for both tensile shear and coach peel load-
ing cases, the nine empirical parameters should be determined by an optimization
method under certain ranges. The generalized reduced-gradient (GRG) method, a di-
rect method for handling nonlinear optimization problems, is used in this study.
GRG is an extension of the reduced-gradient method that was proposed for
solving problems with linear constraints [11]. Its basic idea is to “reduce” the

FIG. 2 Finite-element models of coach peel and tensile shear specimens.

Tensile Shear

Coach Peel
202 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

optimization problem with constraints to an unconstrained problem by solving


equality constraints for the state (dependent) variables in terms of the design (inde-
pendent) variables. Then, the approximated unconstrained nonlinear programming
(NLP) problems can be solved using proven efficient approaches (e.g., the steepest
descent method [12], conjugate gradient method [13,14], and quasi-Newton meth-
od [15], etc.). The basic idea of a GRG algorithm is shown below.
Consider a general NLP problem [16]:
Minimize f ðXÞ (7)
subject to

hj ðXÞ  0; j ¼ 1; 2; …; m (8)
lk ðXÞ ¼ 0; k ¼ 1; 2; …; l (9)
ðl Þ ðu Þ
xi  xi  xi ; i ¼ 1; 2; …; n (10)
ðl Þ ðu Þ
wherein, xi and xi are the lower and upper boundary of the ith variable, respectively.
By adding m nonnegative slack variables to inequality constraints in Eqs 7–10, the prob-
lem is modified to

Minimize f ðXÞ (11)


subject to

hj ðXÞ þ xnþj ¼ 0; j ¼ 1; 2; …; m (12)


lk ðXÞ ¼ 0; k ¼ 1; 2; …; l (13)
ðl Þ ðu Þ
xi  xi  xi ; i ¼ 1; 2; …; n (14)
xnþj  0; j ¼ 1; 2; …; m (15)

Clearly, in Eqs 11–15, the total number of variables is n þ m, and the total number
of equality constraints is m þ l. So, the aforementioned constraints can be further
rewritten in a general form:
gj ðXÞ ¼ 0; j ¼ 1; 2; …; m þ l (16)
ðl Þ ðuÞ
xi  xi  xi ; i ¼ 1; 2; …; n þ m (17)

where the boundaries of the slack variables in Eq 17 are 0 and infinity (i ¼ n þ 1,


n þ 2, …, n þ m).
It is easy to know that, from n þ m variables and m þ l equality constraints, the
original variables [X] can be partitioned into an independent set [Y] and a depen-
dent set [Z] as shown in Eq 18:
 
Y
½X  ¼ ¼ ½ y1 y2    ynl z1 z2    zmþl T (18)
Z
where y1 ; y2 ; …; ynl are design variables, and z1 ; z2 ; …; zmþl are state variables and
dependent on the design variables according to Eq 16.
KANG ET AL., DOI 10.1520/STP159820160054 203

Consider the first variations of the objective function and constraints:


X
nl
@f X
mþl
@f
df ðXÞ ¼ dyi þ dzi (19)
i¼1
@yi i¼1
@zi
X
nl
@gi X
mþl
@gi
dgi ðXÞ ¼ dyj þ dzj (20)
j¼1
@yj j¼1
@zi
Eqs 19 and 20 can be expressed in matrix notation:
df ðXÞ ¼ rTY fdY þ rTZ fdZ (21)
dg ¼ ½CdY þ ½DdZ: (22)
Assuming the GRG method starts at a feasible point leads to gðXÞ ¼ 0, and assum-
ing the maintained feasibility at X þ dX leads to dg ¼ 0. Then, Eq 22 can be
expressed as
dZ ¼ ½D1 ½CdY: (23)
Substituting Eq 23 into Eq 21 results in Eq 24:
 
df ðXÞ ¼ rTY f  rTZ ½D1 ½C dY ¼ GTR dY: (24)
GR is called the generalized reduced gradient and can be used to generate the search
direction S.
Up till now, the constrained optimization problem with n variables is trans-
formed to an NLP problem without constraints. As introduced, many efficient
methods can be utilized to solve this NLP problem, and the optimal solution will be
achieved after iterations. This study used the Broyden–Fletcher–Goldfarb–Shanno
approach [17–21], an updated version of quasi-Newton methods, to find the opti-
mal solution under prescribed constraints.

Data Processing Procedure


The fatigue data used in this study for spot-welded joints comes from the A/SP [9].
Fig. 3 shows the specimen types and geometries, from which there are eight different
kinds of steel with different sheet thicknesses and nugget diameters.
Fig. 4 plots the load versus fatigue-life curve for the case of load ratio R ¼ 0.1.
Apparently, there is a large gap between tensile shear and coach peel results. And it
shows that the load-life curve is also influenced by sheet thickness and nugget
diameter. To construct and optimize the S-N curve that can consider these effects,
the following procedures are introduced:
1. Create FEA models, see Fig. 2, for specimens with different geometries.
2. Extract the nodal forces and moments for unit load from the FEA results as
shown in Table 2 and Table 3.
3. Substitute the nodal forces and moments into the structural stress equation
(Eq 1) to calculate the maximum structural stress (rstru ) for unit load based on
the initial empirical factors as shown in Table 4 and Table 5.
4. Multiply the maximum structural stress (rstru ) for unit load by the applied fa-
tigue load range to obtain the maximum structural stress range corresponding
to each test load range.
204 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 3 Specimen geometry and material combination.

FIG. 4 Load versus fatigue-life curves for tensile shear and coach peel specimens.
KANG ET AL., DOI 10.1520/STP159820160054 205

TABLE 2 Nodal forces and moments for unit load of the tensile shear FEA models.

LS Thickness Load N D mm s1, mm s2, mm Fx1 N Fy1 N Fz1, N Mx1, My1, Mz1,
Combination N-mm N-mm N-mm

HLSA 1.78 mm 1 7.4 1.78 1.78 1 0 0.023734 0 0.89541 0


DQSK 1.62 mm 1 7.6 1.62 1.62 1 0 0.021607 0 0.826416 0
IF 1.63 mm 1 7.6 1.63 1.63 1 0 0.021607 0 0.826416 0
RA830 1.38 mm 1 7.26 1.38 1.38 1 0 0.018411 0 0.706131 0
DP600 1.50 mm 1 7 1.5 1.5 1 0 0.020007 0 0.755438 0
DP800 1.60 mm 1 7 1.6 1.6 1 0 0.02134 0 0.805428 0
MS1300 1.62 mm 1 7.2 1.62 1.62 1 0 0.021607 0 0.826416 0
TRIP 1.60 mm 1 7.23 1.6 1.6 1 0 0.021339 0 0.805428 0

TABLE 3 Nodal forces and moments for unit load of the coach peel FEA models.

CP Thickness Load N D, mm s1, mm s2, mm Fx1 N Fy1 N Fz1 N Mx1, My1, Mz1,
Combination N-mm N-mm N-mm

HLSA 1.78 mm 1 7.4 1.78 1.78 0 0 1 10.78 1.04E-6 0


DQSK 1.62 mm 1 7.6 1.62 1.62 0 0 1 10.80 1.13E-6 0
IF 1.63 mm 1 7.6 1.63 1.63 0 0 1 10.80 1.13E-6 0
RA830 1.38 mm 1 7.26 1.38 1.38 0 0 1 10.83 1.31E-6 0
DP600 1.50 mm 1 7 1.5 1.5 0 0 1 10.81 1.22E-6 0
DP800 1.60 mm 1 7 1.6 1.6 0 0 1 10.80 1.07E-6 0
MS1300 1.62 mm 1 7.2 1.62 1.62 0 0 1 10.80 1.01E-6 0
TRIP 1.60 mm 1 7.23 1.6 1.6 0 0 1 10.81 1.19E-6 0

TABLE 4 Calculation sheet for tensile shear structural stress of unit load.

LS Combination r (Fx1) r (Fy1) r (Fz1) r (Mx1) r (My1) h r Sheet1 for


Unit Load

HLSA 1.78 mm 0.0242 0 0.0105 0 0.0572 180 0.0919


DQSK 1.62 mm 0.0259 0 0.0110 0 0.0592 180 0.0961
IF 1.63 mm 0.0257 0 0.0109 0 0.0587 180 0.0952
RA830 1.38 mm 0.0318 0 0.0119 0 0.0674 180 0.1110
DP600 1.50 mm 0.0303 0 0.0114 0 0.0660 180 0.1077
DP800 1.60 mm 0.0284 0 0.0110 0 0.0639 180 0.1033
MS1300 1.62 mm 0.0273 0 0.0110 0 0.0625 180 0.1008
TRIP 1.60 mm 0.0275 0 0.0110 0 0.0618 180 0.1004
206 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

TABLE 5 Calculation sheet for coach peel structural stress of unit load.

CP Combination r (Fx1) r (Fy1) r (Fz1) r (Mx1) r (My1) h r Sheet1 for


Unit Load

HLSA 1.78 mm 0 0 0.4406 0.6891 6.65E-08 90 1.1297


DQSK 1.62 mm 0 0 0.5075 0.7743 8.10E-08 90 1.2818
IF 1.63 mm 0 0 0.5028 0.7672 8.02E-08 90 1.2700
RA830 1.38 mm 0 0 0.6455 1.0340 1.25E-07 90 1.6795
DP600 1.50 mm 0 0 0.5696 0.9450 1.07E-07 90 1.5145
DP800 1.60 mm 0 0 0.5170 0.8567 8.48E-08 90 1.3738
MS1300 1.62 mm 0 0 0.5075 0.8174 7.64E-08 90 1.3248
TRIP 1.60 mm 0 0 0.5170 0.8305 9.13E-08 90 1.3475

5. Plot the structural stress range versus test fatigue life (S-N curve) as shown in
Fig. 5, then use this S-N curve to calculate the predicted fatigue life for each
structural stress range (black line).
6. The purpose is to reduce the scatter, so R2 is the target to be maximized. See
Eqs 25 and 26 as follows:
R2 ¼ 1  SSE =SST (25)
X  2 X
SSE ¼ i
yi  ^yi ; SST ¼ ðy  yÞ2 :
i i
(26)

FIG. 5 S-N curve with the initial empirical factors.


KANG ET AL., DOI 10.1520/STP159820160054 207

FIG. 6 Nonlinear GRG solver settings in Excel.

SST is a constant, so SSE should be minimized; yi and ^yi represent the test life
and predicted life, respectively. Thus, calculate the differences between each pre-
dicted life and test life, which is called error, then minimize the sum of error
squares (SSE ) of each difference using the nonlinear GRG solver shown in Fig. 6.
The sum of the square of the error of all the specimens is the target for the GRG
solver, and the nine empirical factors are the variables with constraints that the
values are changing from 0 to 1. After completing this, nine new empirical fac-
tors are obtained, which includes the effect of the differences in thickness com-
binations, loading geometries, and nugget diameters on fatigue life of the joints.
7. Calculate the optimized structural stress range based on the nine newly
obtained empirical factors as shown in Table 6 from Step 6 and reconstruct the
S-N curve after optimization as in Fig. 7.

TABLE 6 Optimized empirical factors based on GRG solver.

Scale Factor Diameter Exponent Thickness Exponent

Material SFFXY SFMXY SFFZ DEFXY DEMXY DEFFZ TEFXY TEMXY TEFZ

Steel 0.5810 0.7069 0 0 0.0208 0 1 1 0.7072


208 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 7 (a) Structural stress-life curve with initial empirical factors. (b) Structural stress-
life curve with optimized empirical factors.

(a)

(b)
KANG ET AL., DOI 10.1520/STP159820160054 209

FIG. 8 Predicted versus experimental fatigue life with the optimized empirical factors.

Discussion
The optimized structural stress range versus fatigue-life plot for the test data is pre-
sented in Fig. 7b. The scatter level is obviously smaller after the optimization when
comparing Fig. 7b with Fig. 7a. As a result, R-square is also improved from 0.7283 to
0.8342. Note that in most fatigue data, factor bands around the perfect prediction
line (e.g., factor of five lines placed at the mean 5 and mean /5) provide the reader
with a visual means of assessing the quality of the scatter. Thus, “5” multiplier is
chosen to provide the best visual benchmark for the comparison and to assess the
scatter in this case.
Fig. 8 shows the predicted versus experimental fatigue-life plot, and it also
shows that the estimated specimen life using this procedure correlates well with the
actual fatigue life observed during fatigue testing for different kinds of steel, differ-
ent specimens, different sheet thicknesses, and different nugget diameters. Most of
the data points are located in the “5” scale band. Less scatters in both the S-N
curve and the predicted versus experimental fatigue-life plot are observed after the
parameter optimization.

Conclusions
The structural stress method is widely used in fatigue-life prediction of various
welded joints in vehicle structures because this method is less dependent on mesh
210 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

sizes in finite-element models. However, it requires data-processing procedures to


obtain an optimum structural stress versus fatigue-life curve of the joints. The
structural stress method used in this study is influenced by sheet thicknesses, joint
dimensions, and stress types. By utilizing a nonlinear GRG solver, this paper
proposes a systematic data-processing procedure to optimize the empirical factors
for the structural stress calculation when the database changes or expands. The
optimized results show a lower degree of scatter than previous calculations with
the initial empirical factors. The fatigue life predicted using this procedure is
well-correlated with the test results. Therefore, the proposed optimization
procedure is well-validated.

References

[1] Rupp, A., Störzel, K., and Grubisic, V., “Computer Aided Dimensioning of Spot-Welded
Automotive Structures,” SAE Technical Paper 950711, Society of Automotive Engineers
(SAE), Warrendale, PA, 1995, http://dx.doi.org/10.4271/950711

[2] Wang, P.-C., Corten, H., and Lawrence, F., “A Fatigue Life Prediction Method for Tensile-
Shear Spot Welds,” SAE Technical Paper 850370, SAE, Warrendale, PA, 1985, http://
dx.doi.org/10.4271/850370

[3] Swellam, M., Aś, G. B., and Lawrence, F., “A Fatigue Design Parameter for Spot Welds,”
Fatigue Fract Eng M, 1994, Vol. 17, No. 10, pp. 1197–1204.

[4] Sheppard, S. D., “Further Refinement of a Methodology for Fatigue Life Estimation in
Resistance Spot Weld Connections,” Advances in Fatigue Lifetime Predictive Techniques,
ASTM STP1292, M. R. Mitchell, and R. W. Landgraf, Eds., ASTM International, West Con-
shohocken, PA, 1996, pp. 265–282, http://dx.doi.org/10.1520/STP16142S

[5] Fermér, M., Andréasson, M., and Frodin, B., “Fatigue Life Prediction of MAG-Welded
Thin-Sheet Structures,” SAE Technical Paper 982311, SAE, Warrendale, PA, 1998, http://
dx.doi.org/10.4271/982311

[6] Dong, P., “A Structural Stress Definition and Numerical Implementation for Fatigue Anal-
ysis of Welded Joints,” Int J Fatigue, Vol. 23, No. 10, 2001, pp. 865–876.

[7] Kang, H. T., “Fatigue Prediction of Spot Welded Joints Using Equivalent Structural
Stress,” Mater Design, Vol. 28, No. 3, 2007, pp. 837–843.

[8] Kang, H. T., Dong, P., and Hong, J., “Fatigue Analysis of Spot Welds Using a Mesh-
Insensitive Structural Stress Approach, Int J Fatigue, Vol. 29, No. 8, 2007, pp. 1546–1553.

[9] Bonnen, J. F., Agrawal, H., Amaya, M. A., Iyengar, R. M., Kang, H.-T., Khosrovaneh, A. K.,
Link, T. M., Shih, H.-C., Walp, M., and Yan, B., “Fatigue of Advanced High Strength
Steel Spot-Welds,” SAE Technical Paper 2006-01-0978, SAE, Warrendale, PA, 2006,
http://dx.doi.org/10.4271/2006-01-0978

[10] nCode 12.1, DesignLife Theory Guide, HBM United Kingdom Limited, 2016,
www.ncode.com
KANG ET AL., DOI 10.1520/STP159820160054 211

[11] Gabriele, G. and Ragsdell, K., “The Generalized Reduced Gradient Method: A Reliable
Tool for Optimal Design,” J Eng Ind, 1977, Vol. 99, No. 2, 1977, pp. 394–400.

[12] Cauchy, A., “Méthode Générale pour la Résolution des Systemes d’Equations Simulta-
nées,” Comp. Rend. Sci. Paris, Vol. A, No. 25, 1847, pp. 536–538.

[13] Fletcher, R. and Reeves, C. M., “Function Minimization by Conjugate Gradients,” Compt J,
Vol. 7, No. 2, 1964, pp. 149–154.

[14] Hestenes, M. R. and Stiefel, E., “Methods of Conjugate Gradients for Solving Linear Sys-
tems,” J Res Nat Bur Stand, Vol. 49, No. 6, 1952, pp. 409–436.

[15] Broyden, C. G., “Quasi-Newton Methods and Their Application to Function Mini-
mization,” Math Comput, Vol. 21, No. 99, 1967, pp. 368–381.

[16] Rao, S. S. and Rao, S., Engineering Optimization: Theory and Practice, Wiley, New York,
2009.

[17] Broyden, C. G., “The Convergence of a Class of Double-Rank Minimization Algorithms, 1.


General Considerations,” IMA J Appl Math, Vol. 6, No. 1, 1970, pp. 76–90.

[18] Broyden, C. G., “The Convergence of a Class of Double-Rank Minimization Algorithms, 2.


The New Algorithm,” IMA J Appl Math, Vol. 6, No. 3, 1970, pp. 222–231.

[19] Fletcher, R., “A New Approach to Variable Metric Algorithms,” Compt J, Vol. 13, No. 3,
1970, pp. 317–322.

[20] Goldfarb, D., “A Family of Variable-Metric Methods Derived by Variational Means,” Math
Comput, Vol. 24, No. 109, 1970, pp. 23–26.

[21] Shanno, D. F., “Conditioning of Quasi-Newton Methods for Function Minimization,” Math
Comput, Vol. 24, No. 111, 1970, pp. 647–656.
212 FATIGUE AND FRACTURE TEST PLANNING, TEST DATA ACQUISITIONS AND ANALYSIS

STP 1598, 2017 / available online at www.astm.org / doi: 10.1520/STP159820160057

Xian-Kui Zhu1

More Accurate Elastic


Compliance Equation and Its
Inverse Solution for Compact
Specimens
Citation
Zhu, X.-K., “More Accurate Elastic Compliance Equation and Its Inverse Solution for Compact
Specimens,” Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis, ASTM
STP1598, Z. Wei, K. Nikbin, P. C. McKeighan, and D. G. Harlow, Eds., ASTM International, West
Conshohocken, PA, 2017, pp. 212–226, http://dx.doi.org/10.1520/STP1598201600572

ABSTRACT
This paper reinvestigates the compliance equation used in ASTM E1820,
Standard Test Method for Measurement of Fracture Toughness, for compact
tension (C(T)) specimens. It is found that the elastic compliance equation is
inconsistent with that determined from the stress intensity factor (K) solution
that is also used in ASTM E1820 for the same compact specimens. Because the K
solution in ASTM E1820 is very accurate, as verified by different finite element
analysis (FEA) results, the compliance equation in ASTM E1820 can be less
accurate, making it necessary to develop a more accurate solution. Using the
shooting method and an accurate FEA result of elastic compliance at a/W ¼ 0.5,
the more accurate compliance values are obtained for ASTM E1820 C(T)
specimens over a wide range of a/W by integration of the accurate K solution.
The inverse solution of the new compliance equation is then obtained by curve
fitting. The comparison shows that the proposed compliance inverse solution can
determine very accurate crack length with negligible errors.

Keywords
fracture toughness, fracture testing, stress intensity factor, elastic compliance,
compact specimen, unloading compliance method

Manuscript received March 1, 2016; accepted for publication April 4, 2016.


1
EWI, 1250 Arthur E. Adams Dr., Columbus, OH 43221 http://orcid.org/0000-0002-1554-8998
2
ASTM Symposium on Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis on May 4–5,
2016 in Grand Hyatt, San Antonio, TX.

Copyright V
C 2017 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.
ZHU, DOI 10.1520/STP159820160057 213

Introduction
Fracture toughness testing is essential to crack assessments using fracture mechan-
ics methods. ASTM E1820, Standard Test Method for Measurement of Fracture
Toughness [1], is a well-developed fracture toughness test standard, and it has been
widely utilized in practical engineering for testing fracture initiation toughness and
fracture resistance curves (R-curves) against crack growth for ductile materials in
terms of J-integral or crack tip opening displacement (CTOD). This standard rec-
ommends the elastic unloading compliance technique for measuring instantaneous
crack lengths during crack growth in a single specimen test. In an unloading com-
pliance test, specimen compliance is measured at each unloading-reloading cycle,
and crack length is predicted using the measured compliance and a compliance in-
verse equation for the test specimen. In this case, the accuracy of predicted crack
length depends on the accuracy of the specimen compliance equation. Compact
tension (C(T)) specimens are among the standard specimens used most often in
ASTM E1820 for R-curve testing [2–4].
In ASTM E1820, the standard compliance equation was developed by Saxena
and Hudak [5] for C(T) specimens based on the numerical results obtained by
Newman [6,7], using the boundary collocation method (BCM), and on the analyti-
cal solution obtained by Wilson [8] for very deep cracks. This compliance equation
for C(T) specimens was considered accurate and was used by Joyce and Gudas [2]
in their improved unloading compliance method. This became the basis of the first
ASTM J-R curve test standard (ASTM E1152-87, Standard Test Method for Deter-
mining J-R Curves) and of ASTM E1820 for measuring crack length from a single
specimen test.
For an elastic crack, however, the stress intensity factor (K) and the specimen
unloading compliance (C) at the load-line displacement (LLD) have a simple rela-
tionship, that is, the K factor is a function of force, specimen width, specimen
thickness, crack length, Young’s modulus, and a derivative of the LLD elastic
compliance (dC/da). Thus, the LLD compliance can be determined for a given
K solution for any fracture specimen in the elastic conditions. In fact, Newman [6]
published the BCM numerical results of the K factor for the standard C(T) speci-
mens in ASTM E399, Standard Test Method for Linear-Elastic Plane-Strain Frac-
ture Toughness Klc of Metallic Materials [9]. Based on Newman’s numerical K
results and an asymptotical solution obtained by Bentham and Koiter [10],
Srawley [11] proposed a wide-range stress intensity factor solution for ASTM
E399 C(T) specimens. The Srawley’s K solution was considered very accurate and
was adopted first by ASTM E399 [9] in 1983, then by ASTM E1152 in 1987, and
by ASTM E1820 [1] in 1996.
Using the K solution adopted in ASTM E1820 for the standard C(T) speci-
mens, the corresponding LLD compliance is determined here. It is found that the
K-determined compliance disagrees with the compliance equation used in ASTM
E1820 with a larger error. As verified by finite element analysis (FEA) results, the
214 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

K solution is very accurate for ASTM E1820 compact specimens, and thus, the
ASTM E1820 compliance equation is less accurate. A new compliance equation is
then determined by integration from the accurate K solution using the shooting
method, and its inverse solution is curve fitted. A comparison shows that the
ASTM E1820 inverse equation underestimates actual crack sizes, and that the pro-
posed compliance inverse solution can predict accurate crack sizes.

Problems of the ASTM E1820 Compliance


Equation for C(T) Specimens
ASTM E1820 STRESS INTENSITY FACTOR SOLUTION FOR C(T) SPECIMENS
Appendix A2 of ASTM E1820-15 [1] details the special requirements for C(T)
specimen testing, including the stress intensity factor and the elastic compliance
equations. At a given applied force, the stress intensity factor K for a C(T) specimen
is calculated (Eq 1) using the ASTM E1820 equations (A2.2) and (A2.3) that were
obtained by Srawley [11]:

P a
K¼ f (1)
ðBBN W Þ1=2 W
 h    2  3  4 i
a 2 þ Wa 0:886 þ 4:64 Wa  13:32 Wa þ 14:72 Wa  5:6 Wa
f ¼  3=2 (2)
W 1  Wa

where:
P ¼ the applied force,
W ¼ the specimen width,
a ¼ the crack length,
B ¼ the specimen thickness,
BN ¼ the net thickness for side-grooved specimens,
and BN ¼ B if no side grooves are considered. Eq 2 is accurate within 0.5 % in the
full range of 0.2  a/W  1 for C(T) specimens with a half-height to width ratio,
H/W ¼ 0.6. As detailed in Srawley [11], Eq 2 is a curve-fit function based on the
BCM numerical results of K obtained by Newman [6,7] and on an asymptotic solu-
tion obtained by Bentham and Koiter [10] for C(T) specimens at a/W ¼ 1. Note that
ASTM E399 [9] adopted the K solution in Eqs 1 and 2 for a K-test C(T) specimen
that has no big notch between the load line and the crack mouth in comparison to
the J-test C(T) specimen with a big notch in order to measure LLD.

ASTM E1820 COMPLIANCE EQUATION FOR C(T) SPECIMENS


The specimen elastic compliance for a C(T) specimen at the load line is calculat-
ed using the ASTM E1820 equation (A2.10) that was obtained by Saxena and
Hudak [5]:
ZHU, DOI 10.1520/STP159820160057 215

  a  a 2
1 Wþa 2
CLL ¼ 2:163 þ 12:219  20:065
EBe W  a W W
 a 3  a 4  a 5 (3)
 0:9925 þ 20:609  9:9314
W W W

where CLL ¼ V/P is the load-line compliance, with V being the LLD (for simplicity,
C ¼ CLL is used hereafter), E is the Young’s modulus of materials, Be ¼ B  (B  BN)2/B
is an equivalent thickness to consider the three-dimensional (3D) effect, and Be ¼ B if
no side grooves are considered. Eq 3 is valid over the range of 0.2  a/W  0.975. More
accurate numerical results of C(T) specimen compliance were obtained by Newman
[12] using the BCM and by Tobler and Carpenter [13] using the FEA. Comparisons
[6,7,12,13] indicate that the numerical results of specimen compliance in the plane stress
conditions are better matched with the experimental test data in the 3D conditions.
Thus, the compliance equation (Eq 3) used in ASTM E1820 is given for the plane stress
conditions rather than for the plane strain conditions. Moreover, Tobler and Carpenter
[13] showed that the big notch in ASTM E1820 CT specimens has an ignorable effect
on the LLD compliance of ASTM E399 C(T).
For an elastic unloading compliance test on a single C(T) specimen, the speci-
men load-line compliance is measured during the test, and then the instantaneous
crack size is calculated using the ASTM E1820 equation (A2.12) that was also
obtained by Saxena and Hudak [5]:
a
¼ 1:000196  4:06319u þ 11:242u2  106:043u3 þ 464:335u4  650:677u5 (4)
W

where u ¼ 1/[(BeEC)1/2þ1] is a normalized compliance parameter, and C is the speci-


men load-line crack opening compliance (DV/DP) measured on an unloading/reloading
sequence in a single specimen test. Note that Eq 4 is the curve fit of the inverse function
of Eq 3. As a result, the accuracy of Eq 4 depends on the accuracy of Eq 3 and deter-
mines the accuracy of crack sizes in an unloading compliance test.

RELATIONSHIP BETWEEN STRESS INTENSITY FACTOR AND ELASTIC


COMPLIANCE
Based on the energy balance principle and the relationship between the energy
release rate, G, and the stress intensity factor, K, for an elastic crack, Tada, Paris,
and Irwin [14], in their Appendix A (Compliance Calibration Method), used Eq 5
to relate the K to compliance C:
 
1 P 2 d ðEBC Þ
K2 ¼ W : (5)
2 BW d ða=W Þ

From the aforementioned differential equation, the following integral equation


is obtained:
ð  pffiffiffiffiffi2
a=W
KB W
EBC ¼ EBC0:2 þ 2 da (6)
P
0:2
216 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

where C0.2 is the initial specimen load-line compliance at a/W ¼ 0.2, and the para-
meter a ¼ a/W. To avoid the pinhole effect, usually the crack size a/W  0.2, and thus
a/W ¼ 0.2 is the starting boundary point of the integral Eq 6.

INCONSISTENCY BETWEEN STRESS INTENSITY FACTOR AND ELASTIC


COMPLIANCE
Newman [12] obtained improved numerical results of C(T) specimen compliance
using the BCM, and EBC0.2 ¼ 8.554. With this initial value, substituting the K
expression in Eq 1 into Eq 6 and using the trapezoidal rule of the numerical integra-
tion technique, one obtains the compliance results over the range of 0.2  a/W < 1
for C(T) specimens, as shown in Fig. 1, where a logarithmic scale is used for the
y-axis (i.e., EBC). This figure plots the ASTM E1820 compliance solution from Eq 3
and the present K-determined compliance results as well as their errors relative to
the K-based compliance solution. The comparison shows that (1) these two compli-
ance solutions look similar but are not identical to each other—with a big difference
for shallow cracks; (2) the ASTM E1820 compliance solutions are larger than
K-based compliance results for 0.2  a/W < 0.88; (3) the relative errors are larger
than 1 % for 0.2  a/W < 0.65 and less than 1 % for very deep cracks of a/W > 0.65;
and (4) the maximum error is 6.3 % at a/W ¼ 0.325, and the error is 4.3 % at
a/W ¼ 0.45 allowed by ASTM E1820. Those errors may lead to inaccurate crack

FIG. 1 Comparison of ASTM E1820 compliance and K-determined compliance for C(T)
specimens.

10,000 8 %
K-determined compliance
ASTM E1820 compliance (A2.10) 7%

Error-ASTM E1820 compliance 6%


1,000
5%

4%
Error
EBC

100 3%

2%

1%
10
0%

–1 %

1 –2 %
0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
a/W
ZHU, DOI 10.1520/STP159820160057 217

length estimates. In particular, an overestimated compliance equation predicts


underestimated crack sizes.
In Saxena and Hudak [5], Eq 3 was curve-fitted based on the numerical compli-
ance data obtained by Newman [6,7] for 0.2  a/W  0.8 and on the K-determined
results obtained using the analytical K solution of Wilson [15] for 0.8 < a/W  0.975.
Accordingly, the relative errors are small for those deep cracks. This may imply that
the numerical results of C(T) compliance obtained by Newman [6,7] are not very
accurate for shallow cracks or are sensitive to the big notch of ASTM E1820 C(T)
specimens. Therefore, a more accurate compliance equation that is consistent with the
K solution is developed next for ASTM E1820 C(T) specimens.

Evaluation of Stress Intensity Factor of C(T)


Specimens
Neale and Priddle [16] discussed the stress intensity factor solutions used in different
versions of ASTM E399 for C(T) specimens. The early version of ASTM E399-74
adopted the K solution obtained by Wessel [17], and the geometric function was
expressed as:
1 3 5 7 9
f ðaÞ ¼ 29:6a2  185:5a2 þ 655:7a2  1017a2 þ 638:9a2 (7)

where a ¼ a/W, the valid range is 0.3  a/W  0.7, and the accuracy is within 0.8 %.
In 1983, ASTM E399 (and, in 1987, ASTM E1152, and now ASTM E1820) adopted
the K solution (Eq 1) that was obtained by Srawley [11], and the geometric function is
given in Eq 2. As mentioned previously, Srawley’s K solution was curve-fitted using the
numerical results of Newman [6,7] and an asymptotic solution at a/W ¼ 1. In fact, New-
man [6,7] also obtained another curve-fit K solution with the geometric function as:

f ðaÞ ¼ 4:55  40:32a þ 414:7a2  1698a3 þ 3781a4  4287a5 þ 2017a6 : (8)

This function is accurate within 0.7 % of their numerical results over the range
of 0.2  a/W  0.8.
For C(T) specimens with very deep cracks, Wilson [15] obtained an analytical
K solution, and this geometric function can be written as:

5 þ 3a
f ðaÞ ¼ : (9)
2ð1  aÞ3=2

Fig. 2 compares the geometric function of the ASTM E1820 K solution in Eq 2


with the other three existing solutions (Eqs 7–9) over the range of 0.2  a/W  0.975.
These K solutions agree over their valid ranges and look like one merged curve.
However, details of this figure show that (1) Wessel’s solution in Eq 2 matches
the ASTM E1820 K solution with an error less than 1 % over the range of
0.38  a/W < 0.70; (2) Newman’s solution in Eq 8 matches the ASTM E1820
218 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 2 Comparison of geometric functions of different stress intensity factors for C(T)
specimens.

140
ASTM E1820 (Srawley, 1976)
ASTM E399-74 (Wessel, 1968)
120
Newman (1974)
Wilson (1970)
Newman BCM (1974)
100
Newman BEM (2010)
The present FEA result
80
f(a/W)

60

40

20

0
0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
a/W

K solution with an error less than 1 % over the range of 0.2  a/W < 0.72; and (3)
Wilson’s solution in Eq 9 has excellent agreement with the ASTM E1820 K solution
with an error less than 0.5 % for deep cracks of a/W  0.65 and an overall reasonable
agreement with an error of 10 % or less over the range of 0.2  a/W < 0.65. As a
result, the ASTM E1820 (actually ASTM E399) K solution can represent the other
three K solutions with great accuracy.
Also included in Fig. 2 are the numerical results of K for ASTM E399 C(T) speci-
mens that were obtained by Newman [7] using the BCM and by Newman, Yamada,
and James [18] using the boundary element method (BEM). The comparison shows
that the Srawley’s K solution agrees well with the numerical K results, with an error
less than 0.9 %. In addition, Riddell and Piascik [19] showed that their FEA results
for K for ASTM E399 C(T) specimens matched well with Srawley’s results over the
range of 0.2  a/W  0.9, with an error less than 0.3 %. In order to verify the Sraw-
ley’s K solution in Eq 1 as accurate for ASTM E1820 C(T) specimens having a large
notch, FEA calculations are performed for two typical ASTM E1820 1T C(T) speci-
mens, with a/W ¼ 0.454 and 0.5 in the plane stress conditions as detailed here in
Appendix A. The FEA results of K are included in Fig. 2, and the comparison shows
that the differences between the Srawley’s K solution and the FEA results are very
ZHU, DOI 10.1520/STP159820160057 219

small and less than 0.5 %. Therefore, the Srawley’s K solution obtained for the
ASTM E399 C(T) specimen is also very accurate for ASTM E1820 C(T) specimens.

More Accurate Elastic Unloading Compliance


Equation of C(T) Specimens
K-DETERMINED COMPLIANCE EQUATION
With the accurate K solution in Eq 1, more accurate load-line compliance can be
obtained by integration of the integral equation (Eq 6), where an accurate initial
compliance C0.2 is needed for the integration. Recall that the ASTM E1820 compli-
ance equation (Eq 3) was obtained in reference to the numerical results of Newman
[6,7], where EBC0.2 ¼ 8.6. Newman [12] obtained improved compliance values and
EBC0.2 ¼ 8.554. With this initial value, the K-based compliance was determined and
compared with the ASTM E1820 compliance solution (Eq 3) as shown in Fig. 1,
where the ASTM E1820 compliance solutions are slightly higher than the K-determined
solutions, particularly for shallower cracks, and the maximum error is approximate-
ly 6.3 % at a/W ¼ 0.325. In addition, accurate FEA results of the load-line compli-
ance were obtained by Shih and deLorenzi [20] for an ASTM E399 C(T) specimen
with a/W ¼ 0.6 in the 3D conditions and by Tobler and Carpenter [13] for ASTM
E1820 C(T) specimens with different crack sizes in the plane stress conditions. The
present work also obtains the FEA results of the compliance for two typical ASTM
E1820 C(T) specimens, as detailed in Appendix A. Fig. 3 compares these FEA
results with the ASTM E1820 compliance solution and the K-determined compli-
ance results, where the logarithmic scale is used for the y-axis (EBC). Fig. 3 shows
that the FEA results better match the K-determined compliance results.
Because the big notch of ASTM E1820 C(T) specimens may affect the compli-
ance at a/W ¼ 0.2, and the crack length of ASTM E1820 C(T) specimens ranges
from 0.45 to 0.7, the FEA compliance EBC0.5 ¼ 35.539 at a/W ¼ 0.5 is selected as a
target value for integration of the integral equation (Eq 6) using the shooting meth-
od. The shooting accuracy is controlled by an error of 0.5 % of the target compli-
ance value. This target value is about 3.4 % lower than the BCM value obtained by
Newman [7], and so their difference is not large.
With the target compliance value and the accurate ASTM E1820 K solution in
Eq 1, the K-based compliance is obtained and shown in Fig. 4. Using the least squares
regression method, the best curve fit of the K-determined compliance is obtained as:
  a  a 2  a 3
1 Wþa 2
CLL ¼ 4:406  10:525 þ 54:428  111:15
EB W  a W W W
 a 4  a 5 (10)
þ96:672  29:753 :
W W

Fig. 4 compares the new curve-fit compliance (Eq 10) with the K-determined
compliance values and the ASTM E1820 compliance (Eq 3). The relative errors of
220 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 3 Comparison of ASTM E1820 compliance with K-determined and FEA results for
C(T) specimens.

500

K-determined compliance
ASTM E1820 compliance (A2.10)
The present FEA results
FEA - Tobler (1985)
FEA - Shih (1974)
EBC

50

5
0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
a/W

Eq 10 and Eq 3 when compared with the more accurate K-determined compliance


are also included in Fig. 4, which shows that the newly proposed compliance
(Eq 10) is very accurate, with relative errors less than 0.05 % over the range of
0.2  a/W  0.77 and 0.18 % for other deep cracks. In contrast to this, the ASTM
E1820 compliance equation (Eq 3) has errors larger than 1 % for a/W  0.66, an
error of approximately 5 % at a/W ¼ 0.45, and a maximum error of approximately
7.4 % at a/W ¼ 0.32. Those errors can be reduced if the new Eq 10 is used.

INVERSE EQUATION OF K-DETERMINED COMPLIANCE


From Eq 10, using the compliance parameter u ¼ 1/[(EBC)1/2 þ 1], the a/W values
can be determined as a function of u, and then the compliance inverse equation is
curve fitted as:
a
¼ 1:0  4:0279u þ 10:161u2  95:45u3 þ 461:33u4  740:65u5 : (11)
W

Fig. 5 directly compares the predicted a/W from ASTM E1820 Eq 4 and the
proposed Eq 11. From Fig. 5, it is observed that (1) the newly proposed inverse com-
pliance (Eq 11) is very accurate, with errors less than 0.05 % for deep cracks of
0.42  a/W  0.975 and 0.2 % for shallow cracks of 0.2 < a/W < 0.42; (2) the
ASTM E1820 compliance inverse equation (Eq 4) underestimates the crack sizes
ZHU, DOI 10.1520/STP159820160057 221

FIG. 4 K-determined compliance versus ASTM E1280 compliance and the relative
errors.

10,000 8%

K-determined compliance
ASTM E1280 Compliance (A2.10) 6%
1,000
New curve-fit compliance
Error-new compliance
Error-E1820 compliance 4%

Error
EBC

100

2%

10
0%

1 –2 %
0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
a/W

over 0.45  a/W  0.7 for R-curve evaluations; and (3) the ASTM E1820 predic-
tions are accurate only for deep cracks of a/W > 0.55 with an accuracy of 1 % and
are inaccurate for all other crack sizes, with a maximum error of approximately 5 %
at a/W ¼ 0.29. Because the proposed inverse compliance equation (Eq 11) has a
higher accuracy, it can be used to estimate more accurate crack sizes for C(T) speci-
mens in an unloading compliance test.
Based on the BCM results of load-line compliance obtained by Newman [12],
Kapp [21] curve fitted another compliance equation similar to Eq 3 in a lower-degree
polynomial and another compliance inverse equation similar to Eq 4 in the same de-
gree polynomial function for ASTM E399 C(T) specimens. A comparison shows that
both Kapp’s compliance and its inverse equations have errors comparable to the ASTM
E1820 compliance and its inverse equations; thus, Kapp’s results are not suggested for
practical use.

Applications of Proposed Inverse Compliance


Equation
Consider a standard 1T C(T) specimen with W ¼ 50.8 mm. Fig. 6 shows the differ-
ence between the predicted and actual crack sizes, where the predictions are
222 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 5 Comparison of predicted a/W by ASTM E1820 and proposed compliance


equations with actual a/W.

1.0 1%

0.9 0%

0.8 –1 %

0.7 –2 %
Predicted a/W

Error
0.6 –3 %

0.5 ASTM E1820 predicon –4 %


New predicon
0.4 –5 %
ASTM E1820 predicon error
0.3 New predicon error –6 %

0.2 –7 %
0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Actual a/W

FIG. 6 Difference between predicted and actual crack sizes for 1T C(T) specimen.

0.1
1T CT specimen
0.0

–0.1

–0.2
apredicted-aactual (mm)

–0.3

–0.4

–0.5
New predicon
ASTM E1820 predicon
–0.6

–0.7

–0.8

–0.9
0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
a/W
ZHU, DOI 10.1520/STP159820160057 223

obtained by the ASTM E1820 compliance inverse equation (Eq 4) and the pro-
posed compliance inverse equation (Eq 11), respectively. Over the entire range
of 0.2  a/W  0.95, differences between the crack sizes predicted by Eq 11 and
the actual crack sizes are negligibly small. In contrast, even in the ASTM E1820
allowed crack size range of 0.45  a/W  0.70, the differences between the crack
sizes predicted by the ASTM E1820 compliance inverse equation (Eq 4) and the
actual crack sizes are not small with apparent underestimation. The differences
are observed as 0.5 mm (or 2.2 %) at a/W ¼ 0.45, 0.234 mm (or 0.84 %) at
a/W ¼ 0.55, and 0.05 mm (or 0.14 %) at a/W ¼ 0.7. Accordingly, crack sizes
are underestimated largely by the ASTM E1820 compliance equation (Eq 4)
only for shorter cracks, and the underestimation decreases with increasing crack
sizes and approaches zero at an a/W of approximately 0.8. This shows the
ASTM E1820 compliance equation can be used to estimate reasonable crack
lengths within an accuracy of 1 % for compact specimens with deep cracks of
a/W  0.55.

Conclusions
This paper restudied the stress intensity factor solution and the compliance equa-
tion used in ASTM E1820 for C(T) specimens. It was determined that the ASTM
E1820 load-line compliance equation is inconsistent with that determined from the
stress intensity factor solution that is also used in ASTM E1820 for the same com-
pact specimens. The ASTM E1820 K solutions were verified by different numerical
results for K obtained using the FEA, BCM, and BEM for both ASTM E399 and
ASTM E1820 C(T) specimens and were shown to be very accurate. Thus, a more
accurate compliance equation was determined from this stress intensity factor solu-
tion by the integration and shooting methods. The results showed that the proposed
compliance equation and its inverse equation are very accurate, with negligible
errors over a wide range of 0.2  a/W  0.975 for compact specimens, and that the
ASTM E1820 compliance equation and its inverse equation are accurate only for
very deep crack sizes.
The proposed compliance equation and its inverse solution for ASTM E1820
could be used to determine accurate crack length by the normalization method
and the unloading compliance method, respectively. As a result, more accurate
R-curves could be developed using the new compliance solution in a single speci-
men test [22].

ACKNOWLEDGMENTS
This work was financially supported by an internal research and development (R&D)
project at EWI. The present author is grateful to Tom McGaughy, director of the EWI
R&D program, for his great support, and to Professor Jim Joyce at the U.S. Naval
Academy for his helpful discussions. Thanks also go to the anonymous reviewers for
their valuable comments and suggestions.
224 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

Appendix A: Finite element results of stress


intensity factor and compliance of ASTM E1820
C(T) specimen
The FEA calculations are performed using ABAQUS in the plane stress conditions to
determine the stress intensity factor and load-line elastic compliance for an ASTM
E1820 C(T) specimen with the specimen width as W ¼ 50.8 mm and two crack
lengths, a/W ¼ 0.454 and 0.5. Due to the symmetry, only one-half of the CT specimen is
modeled, as shown in Fig. A1. This FEA model contains 4693 nodes and 1501 plane stress
eight-node elements with reduced integration (ABAQUS element type is CPS8R).

FIG. A1 (a) FEA mesh and (b) deformed contour of one-half of the ASTM E1820 C(T)
specimen.

(a)

(b)
ZHU, DOI 10.1520/STP159820160057 225

The meshes are very fine near the crack tip, with the smallest element size of
2.94  104a, and become coarse elsewhere. The rigid pin contact with the specimen
at the load-pin holes is simulated by the rigid beam constraint built in ABAQUS.
Tensile force is applied at the pinhole center, where the horizontal displacement is
restrained. The symmetric boundary condition on the uncracked ligament is imposed
in the FEA model.
The material used in the FEA calculations is X60 pipeline steel with a yield stress
of 414 MPa, an ultimate yield stress of 552 MPa, the Young’s modulus of 207 GPa,
and the Poisson ratio of 0.3. The linear elastic FEA calculations were conducted for
the C(T) specimens. For a/W ¼ 0.5, at a force F ¼ 558.85 N, the FEA results are
K ¼ 792.8 N/mm3/2 and BEC ¼ 35.46. For a/W ¼ 0.454, at a force F ¼ 82.5 N, the FEA
results are K ¼ 138.4 N/mm3/2 and BEC ¼ 28.61.

References

[1] ASTM E1820-15, Standard Test Method for Measurement of Fracture Toughness, ASTM
International, West Conshohocken, PA, 2015, www.astm.org

[2] Joyce, J. A. and Gudas, J. P., “Computer Interactive JIc Testing of Navy Alloys,” Elastic-
Plastic Fracture, ASTM STP668, J. D. Landes, J. A. Begley, and G. A. Clarke, Eds., ASTM
International, West Conshohocken, PA, 1979, pp. 451–468, http://dx.doi.org/10.1520/
STP668-EB

[3] Bertolino, G., Meyer, G., and Ipina, J. P., “Mechanical Properties Degradation at Room Tem-
perature in ZRY-4 by Hydrogen Brittleness,” Mat. Res., Vol. 5, No. 2, 2002, pp. 125–129.

[4] Weiss, K. and Nyilas, A., “Specific Aspects on Crack Advance During J-Test Method for
Structural Materials at Cryogenic Temperatures,” Fatigue Fract. Eng. M., Vol. 29, No. 2,
2006, pp. 83–92.

[5] Saxena, A. and Hudak, S. J., “Review and Extension of Compliance Information for Com-
mon Crack Growth Specimens,” Int. J. Fracture, Vol. 14, No. 5, 1978, pp. 453–468.

[6] Newman, J. C., Jr., “Stress Analysis of the Compact Specimen Including the Effects of
Pin Loading,” Fracture Analysis: Proceedings of the 1973 National Symposium on Frac-
ture Mechanics, Part II, ASTM STP560, G. R., Irwin, Ed., ASTM International, West Con-
shohocken, PA, 1974, pp. 105–121, http://dx.doi.org/10.1520/STP33136S

[7] Newman, J. C., Jr., Crack Opening Displacements in Center-Crack, Compact and Crack-
Line Wedge-Loaded Specimens, NASA TN D-8268, NASA Langley Research Center,
Hampton, VA, 1976.

[8] Wilson, W. K., “Stress Intensity Factors for Deep Cracks in Bending and Compact Ten-
sion Specimens,” Eng. Fract. Mech., Vol. 2, No. 2, 1970, pp. 161–171.

[9] ASTM E399-12e3, Standard Test Method for Linear-Elastic Plane-Strain Fracture Tough-
ness KIc of Metallic Materials, ASTM International, West Conshohocken, PA, 2013,
www.astm.org
226 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

[10] Bentham, J. T. and Koiter, W. T., Mechanics of Fracture I—Methods of Analysis and Solu-
tion of Crack Problem, Noordhoff International, Leyden, Netherlands, 1973, pp. 159–162.

[11] Srawley, J. E., “Wide Range Stress Intensity Factor Expressions for ASTM E399 Standard
Fracture Toughness Specimens,” Int. J. Fracture, Vol. 12, No. 3, 1976, pp. 475–476.

[12] Newman, J. C., Jr., “Stress-Intensity Factors and Crack-Opening Displacements for
Round Compact Specimens,” Int. J. Fracture, Vol. 17, 1981, pp. 567–578.

[13] Tobler, R. L. and Carpenter, W. C., “A Numerical and Experimental Verification of Compli-
ance Functions for Compact Specimens,” Eng. Fract. Mech., Vol. 21, No. 3, 1985, pp.
547–556.

[14] Tada, H., Paris, P. C., and Irwin, G. R., The Stress Analysis of Cracks Handbook, 3rd ed.,
ASME, New York, 2000.

[15] Wilson, W. K., “Stress Intensity Factors for Deep Cracks in Bending and Compact Ten-
sion Specimens,” Eng. Fract. Mech., Vol. 2, No. 2, 1970, pp. 168–171.

[16] Neale, B. K. and Priddle, E. K., “J-Integral Functions of a Compact Tension Specimen for
Fracture and Fatigue Analysis,” Eng. Fract. Mech., Vol. 26, No. 3, 1987, pp. 323–331.

[17] Wessel, E. T., “State of the Art of the WOL Specimen for KIc Fracture Toughness Testing,”
Eng. Fract. Mech., Vol. 1, No. 1, 1968, pp. 77–103.

[18] Newman, J. C., Yamada, Y., and James, M. A., “Stress Intensity Factor Equations for Com-
pact Specimen Subjected to Concentrated Forces,” Eng. Fract. Mech., Vol. 77, No. 6,
2010, pp. 1025–1029.

[19] Riddell, W. T. and Piascik, R. S., A Back Face Strain Compliance Expression for the
Compact Tension Specimen, NASA/TM-1998-208453, NASA Langley Research Center,
Hampton, VA, 1998.

[20] Shih, C. F. and deLorenzi, H. G., “Elastic Compliance and Stress-Intensity Factors for
Side-Grooved Compact Specimens,” Int. J. Fracture, Vol. 13, No. 4, 1977, pp. 544–548.

[21] Kapp, J. A., “Improved Wide Range Expressions for Displacements and Inverse Displace-
ments for Standard Fracture Toughness Specimens,” Journal of Testing and Evaluation,
Vol. 19, No. 1, 1991, pp. 45–54, https://doi.org/10.1520/JTE12529J

[22] Zhu, X. K. and Joyce, J. A., “Review of Fracture Toughness (G, K, J, CTOD, CTOA) Testing
and Standardization,” Eng. Fract. Mech., Vol. 85, 2012, pp. 1–46.
FATIGUE AND FRACTURE TEST PLANNING, TEST DATA ACQUISITIONS AND ANALYSIS 227

STP 1598, 2017 / available online at www.astm.org / doi: 10.1520/STP159820160059

Hao Wu1 and Zheng Zhong1

A Novel Nonlinear Kinematic


Hardening Model for Uniaxial/
Multiaxial Ratcheting and Mean
Stress Relaxation
Citation
Wu, H. and Zhong, Z., “A Novel Nonlinear Kinematic Hardening Model for Uniaxial/Multiaxial
Ratcheting and Mean Stress Relaxation,” Fatigue and Fracture Test Planning, Test Data
Acquisitions and Analysis, ASTM STP1598, Z. Wei, K. Nikbin, P. C. McKeighan, and D. G. Harlow,
Eds., ASTM International, West Conshohocken, PA, 2017, pp. 227–245, http://dx.doi.org/
10.1520/STP1598201600592

ABSTRACT
Ratcheting behavior or cyclic mean stress relaxation can have a significant effect on
multiaxial fatigue lives when compared to proportional loads of similar range. The
most important feature for ratcheting simulation in cyclic plasticity constitutive
models is the nonlinear kinematic (NLK) hardening rule, which causes the
translation of the yield surface during a plastic strain increment. A general nonlinear
kinematic hardening model capable of simulating uniaxial ratcheting, multiaxial
ratcheting, and mean stress relaxation (MSR) response is presented in this paper to
capture these effects in incremental plasticity simulations. The uniaxial/multiaxial
loading paths, including cyclic MSR, are simulated accordingly. Verification of the
proposed model is achieved by comparing the predicted results with experimental
measurements with 316L tubular specimens and the results from other NLK models.

Keywords
multiaxial fatigue, nonlinear kinematic hardening, ratcheting, nonproportional
loading, incremental plasticity

Manuscript received March 7, 2016; accepted for publication November 14, 2016.
1
Tongji University, School of Aerospace Engineering and Applied Mechanics, 1239 Siping Rd., Shanghai,
200092 P.R. China
2
ASTM Symposium on Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis on May 4–5,
2016 in Grand Hyatt, San Antonio, TX.

Copyright V
C 2017 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.
228 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

Introduction
The Bauschinger effect, observed under cyclic elastoplastic loading (also called kine-
matic hardening), is a change in the absolute value of the opposite yield strength
after strain hardening that is due to the microscopic stress distribution induced
by the rearrangement of the dislocation structure inside the material associated
with the plastic strains. For a tension-torsion history, this effect can be repre-
sented as a translation of the yield surface, whose center translates to a position
h pffiffiffi iT
~¼ b b
b x xy 3 (called back stress), see Fig. 1. This back stress vector is the
quantity that stores plastic memory required for kinematic hardening calculations.
As mentioned by Meggiolaro, Castro, and Wu [1], dealing with multiaxial cal-
culations with reduced dimensions can save computational costs without modifying
the results. The transformation from six-dimensional (6D) stresses or strains to
deviatoric stresses ~
s and strains ~e can be represented by a 6 by 6 projection matrix
[A6D] through~ s ¼ A6D  ~ e ¼ A6D  ~
r and ~ e, given by Eq 1:
2 3
2=3 1=3 1=3 0 0 0
6 1=3 2=3 1=3 0 0 0 7
6 7
6 1=3 1=3 2=3 0 0 0 7
6
½A6D  ¼ 6 7: (1)
6 0 0 0 1 0 07 7
4 0 0 0 0 1 0 5
0 0 0 0 0 1

The 5D deviatoric stress vector~ s 0 has three important properties:


1. The Euclidean norm of the 5D vector ~ s 0 from the E5s deviatoric subspace is
equal to the Mises equivalent stress rMises, as shown in Eq 2:
pffiffiffiffiffiffiffi pffiffiffi
s 0 j ¼ j~
j~ sj= 2=3 ¼ rMises  sMises 3 (2)

2. The Euclidean distance in the E5s space between two stress states (points), A
and B, respectively, associated with the 6D deviatoric stresses, ~ sA and ~
sB , is
equal to the Mises range rMises between these stress states, giving:
pffiffiffiffiffiffiffi pffiffiffi
s 0B ~
j~ s 0A j ¼ j~
sB ~
sA j= 2=3 ¼ DrMises  DsMises 3: (3)

FIG. 1 Kinematic hardening in the x direction and associated yield surface translation in
pffiffiffi
the rx  sxy 3 von Mises diagram.
WU AND ZHONG, DOI 10.1520/STP159820160059 229

3. The locus of the points that have the same range rMises with respect to a stress
state ~ s 0 in E5s is the surface of a hypersphere with a center in ~ s 0 and radius
rMises, which is a corollary from the second property.
Kinematic hardening, for example, can be modeled in the 5D formulation (or in
its 3D, 2D, or 1D subspaces) by allowing the yield surface j~ s 0 j ¼ S to translate its cen-
ter from the origin of the E5s space to a 5D back stress position b ~0 , becoming repre-
sented by j~ 0 ~0
s  b j ¼ S, with no change in its radius S or shape. The proposed 5D
stress and strain formulation will be used in this paper to better describe nonlinear
incremental plasticity models, which are required for predicting ratcheting and mean
stress relaxation (MSR) effects. In some classical yield surface translation cases, the
uniaxial unbalanced histories are essentially cyclic histories with a significant mean
stress component. Such histories may present plastic strain accumulation in the
direction of the mean stress, called uniaxial ratcheting (UR). UR is a consequence of
anisotropy between the tension and compression behaviors with non-zero mean
stress, which forms an elastoplastic loop that does not close, causing the material to
accumulate a net strain during each cycle. The gradual plastic strain accumulation
along a certain direction under multiaxial load histories is called multiaxial ratcheting
(MR). However, MR happens for unbalanced histories even if the material follows
the Masing behavior [2], without any asymmetry in the hysteresis loops under ten-
sion and compression. On the other hand, strain-controlled deformation with an ini-
tial mean stress can cause MSR [3], a phenomenon closely related to ratcheting. The
mean stress gradually relaxes toward zero, both in uniaxial and multiaxial unbal-
anced histories. MSR caused by high plastic strain ranges is one of the reasons why
low-cycle fatigue lives are less influenced by the mean stress effect than high-cycle
fatigue lives. It is present as well in multiaxial elastoplastic paths with mixed stress
and strain control, as long as the relaxation direction is under strain control with an
initial mean stress. MSR can be quantitatively predicted from incremental plasticity
simulations [4] if nonlinear kinematic (NLK) models are used to describe the associated
asymmetrical behavior of the stress-strain curves as described in the next section.
The translation of the yield surface is associated with plastic straining, usually
assumed from the normality rule in the direction of the unit normal to the yield sur-
face, defined as ~ n for the 6D and ~ n0 for the adopted 5D formulation, evaluated at the
current stress point. The Prandtl-Reuss plastic flow rule assumes that the magnitude
of the plastic strain increment d~ epl (in 6D) or d~e pl0 (in 5D) depends on the applied stress
increment, being inversely proportional to the generalized plastic modulus P that
defines the slope between stress and plastic strain increments. The Prandtl-Reuss rule
is usually defined in tensor or 6D notation, but it is easy to show from the relations
pffiffiffiffiffiffiffi pffiffiffiffiffiffiffi
n0 ¼ A ~
~ n  2=3, ~ n0  2=3, and A  AT ¼ 1:5  I55 (where I5  5 is the 5 by 5
n ¼ AT ~
identity matrix) that it has an almost identical version in the proposed 5D spaces:

1 1 1
epl ¼
d~ rT  ~
 ðd~ nÞ  ~ sT  ~
n ¼  ðd~ nÞ  ~ e pl0 ¼  ðd~
n ) d~ s 0T  ~
n0 Þ  ~
n0 (4)
P P P

using the same P without even the need for a scaling factor.
230 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

There are several models to calculate the current value of the generalized plastic
modulus P as the yield surface translates, as well as the direction of such translation,
to obtain the associated plastic strain increments. Most of these strain-hardening
models can be divided into three classes: multi-surface [5,6], two-surface [7,8], and
nonlinear [9] kinematic hardening models.
Multi-surface kinematic hardening models assume that P is piecewise constant,
resulting in a multilinear description of the stress-strain curve (i.e., the nonlinear
shape of the stress-strain relation is approximated by several linear segments). NLK
hardening models, on the other hand, are more general because they adopt nonline-
ar equations to describe the surface translation direction and the value of P, leading
to a more precise nonlinear description of the stress-strain curve. A third class of ki-
nematic hardening models involves the so-called two-surface models, which use a
rather simplified formulation that combines elements of both nonlinear and multi-
surface kinematic models.
A major limitation of multi-surface models is they cannot predict any UR or MSR
caused by unbalanced loadings because their unrealistic perfectly symmetric hysteresis
loops always close. In addition, under several nonproportional (NP) loading conditions,
these models predict MR with a constant rate that never decays, severely overestimating
the ratcheting effect measured in practice [10]. As a result, multi-surface kinematic
hardening models should only be confidently applied to balanced proportional loading
histories. To correctly predict the stress-strain history associated with unbalanced load-
ings, it is necessary to use NLK hardening models. The original Armstrong and Freder-
ick formulation [11] was improved by Chaboche, Dang Van, and Cordier [12], who
indirectly introduced some multi-surface elements into the NLK models, in a better
nonlinear instead of the simplistic multilinear formulation, however.

Nonlinear Kinematic Hardening Models


NLK hardening models adopt nonlinear equations to describe P; therefore, their
formulation is the one adopted in this work. Two-surface models combine elements
of both NLK and Mróz multi-surface kinematic hardening models; however, their
description of plastic memory is limited and not recommended for complex vari-
able amplitude (VA) histories commonly found in fatigue. The original NLK model,
proposed by Armstrong and Frederick, uses nonlinear equations to calculate the
~ of the yield surface; however, its plastic memory capabilities are too
translation d b
limited to be applied to more complex VA loading. Chaboche, Dang Van, and
Cordier improved their model, indirectly introducing some multi-surface elements
into the NLK models, however in a better nonlinear formulation—instead of the
simplistic Mròz multilinear approach. Chaboche, Dang Van, and Cordier assumed
~ can be represented as a sum of M back stress compo-
that the yield surface center b
~
nents bi (i ¼ 1, 2, …, M) as seen in Eq 5 (see Fig. 2). Thus,

~ b
b ~ þ ::: þ b
~ þb ~ : (5)
1 2 M
WU AND ZHONG, DOI 10.1520/STP159820160059 231

pffiffiffi
FIG. 2 Representation of the yield surface center in the sx  txy 3 von Mises diagram
using M ¼ 3 back stress components.

τxy 3 s
S
yield
β1 surface

β
β2
σx
β3

Each back stress component b ~ is associated with a saturation value Dri that
i
limits the maximum value of its norm; thus, jb ~ j  Dri . A zero norm is related to an
i
unhardened material state with high generalized plastic modulus P (note that a
purely elastic state would have P ! 1), while a saturated norm jb ~ j ¼ Dri repre-
i
sents a hardened state with much lower P (conversely, a plastic collapse condition
with ~epl ! 1 would have P ! 0). Moreover, each b ~ is also associated with a gen-
i
eralized plastic modulus coefficient pi, used in the calculation of the current value of
P through Eq 6:

v1 T þ p2 ~
P ¼ ð2=3Þ  ðp1 ~ v2 T þ ::: þ pM ~
vM T Þ  ~
n (6)

where ~ vi are translation direction vectors for each surface that depend on the adopted
NLK model. Eqs 7 and 8 satisfies the consistency condition, which prevents the cur-
rent stress state~s from moving outside the yield surface; that is, it prevents the impos-
sible configuration j~ ~ > S, as long as the infinitesimal translations db
s  bj ~ of the back
i
~
stress components bi are such that

pi ~ ~ j < Dri ðunsaturated condition for b
vi  dp; ifjb ~Þ

db i i i ¼ 1; 2; …; M; (7)
i ~ ~
0; if jbi j ¼ Dri ðsaturated condition for bi Þ

 
where dp  ð2=3Þ  d~
epl  is the equivalent plastic strain increment, which, from Eq 4,
results in

sT  ~
dp ¼ ð2=3Þ  ðd~ nÞ=P: (8)

KINEMATIC HARDENING FORMULATION


Using the aforementioned notation, the only difference among the main NLK
models is the choice of the equation for the translation direction vectors ~
vi .
232 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

A generalized surface translation rule has been proposed by Meggiolaro, Castro,


and Wu [13]:

~ ~ þ ð1  di Þ  ðb
n  Dri vi   mi   ci  ½di  b
vi ¼ ~ ~T  ~ nÞ  ~ n (9)
|fflffl{zfflffl} |fflffl{zfflffl}i |fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl i
ffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl
ffl}
Prager dynamic radial
Ziegler recovery return

where the scalar functions vi  and mi  are defined in Eq 10:


!v 8h . imi
jb~j i < b ~T  ~
n jb ~T  ~
~ j ; if b n 0
 i  i i i
vi  and mi  (10)
Dri : ~ ~T
0; if b i n<0

where vi, mi, ci, and di are adjustable ratcheting parameters for each back stress com-
ponent. These versatile equations reproduce Chaboche’s model [12] if vi ¼ ci ¼ di ¼ 1
and mi ¼ 0; Jiang-Sehitoglu’s model [14,15] for ci ¼ di ¼ 1, mi ¼ 0 and adjustable
0 < vI < 1; Ohno-Wang II [16] for ci ¼ di ¼ mi ¼ 1 and adjustable 0 < vi < 1; and
Delobelle’s [17] for vi ¼ mi ¼ 0 and adjustable 0 < ci < 1 and 0 < di < 1.
Replacing Eq 9 into Eq 6, the generalized plastic modulus becomes the form of Eq 11:
X
M X
M
P ¼ ð2=3Þ  vi T  ~
pi ~ n ¼ ð2=3Þ  ~T  ~
pi  ðDri  vi   mi   ci  b nÞ: (11)
i
i¼1 i¼1

The 5D translation direction ~ vi0 of each yield surface from Eq 9 can be separated into
three components: (i) the Prager-Ziegler term, in the normal direction ~ n0 perpendicu-
lar to the yield surface; (ii) the dynamic recovery term, in the opposite direction b ~0
i
of the back stress induced by that yield surface, which acts as a recall term that gradu-
ally erases plastic memory with an intensity proportional to the product
vi  mi  ci  di ; and (iii) the radial return term, in the opposite direction ~ n0 of the
normal vector, which affects MR predictions, calibrated from vi  mi   ci  

ð1  di Þ  ðb ~0 T  ~
n0 Þ.
i
Among the models, Prager’s [18] translation direction ~ vi0 ¼ ~
n0  Dri is not able
to predict UR because it only uses the Prager-Ziegler term, which is linear. For MR,
it only predicts a very short transient that almost immediately arrests, reaching a
shakedown state and highly underestimating MR rates. The translation direction
vi0 ¼ ~
~ n0  Dri  b ~0 from Mróz [5] includes a dynamic recovery term, however in a
i
linear formulation that highly overestimates MR effects. As discussed before, the
Mróz rule cannot predict any UR at all when used in a multi-surface formulation
where the outer surfaces not touched by the current stress point are not allowed to
translate. But when applied to the NLK formulation, where all surfaces translate
during plastic straining, the Mróz translation direction becomes capable of predict-
ing UR, albeit largely overestimating it.
Armstrong and Frederick proposed the use of a ratcheting coefficient
0  ci  1, originally intended to be a scalar function of the plastic strain path, add-
ing nonlinearity to their hardening model [11]. However, in many practical imple-
mentations, ci was assumed as a constant, turning their translation equation
vi0 ¼ ~
~ n0  Dri  ci  b ~0 into a linear rule that suffers the same drawbacks of the Mróz
i
WU AND ZHONG, DOI 10.1520/STP159820160059 233

translation rule in the NLK formulation, with a large overestimation of both uniaxi-
al and MR. Even though ci can calibrate ratcheting rates, with the limit values ci ¼ 0
(Prager’s rule) for no ratcheting and ci ¼ 1 (Mróz rule) for large ratcheting rates,
the linearity associated with a constant ci makes it impossible to predict MR rate
decay and arrest (shakedown) observed in several constant amplitude experiments.
In addition, for constant coefficients ci < 1, the Armstrong-Frederick translation
rule would result in d b ~0 6¼ 0 in the saturated condition, which would allow the
i
surfaces to pass through one another. To avoid this, it has been proposed to simply
enforce ci ¼ 1 in the saturation condition jb ~0 j ¼ Dri , while allowing the use of a cali-
i
~0 j < Dri [19]. Each of the Armstrong-Frederick and Mróz rules
brated ci < 1 for jb i
is a particular case of the proposed generalized surface translation rule from Eq 9
and Eq 10, for vi ¼ mi ¼ 0, di ¼ 1, and an adjustable 0  ci  1 that for Mróz is set
to ci ¼ 1.
Chaboche [12] replaced the constant ratcheting coefficient ci with . a saturation
ratio, which in the proposed 5D formulation is represented asjb ~0 j Dri , ranging
i
from 0 in the unhardened condition to 1 at saturation, eliminating the discontinuity
problem caused by ci = 1. But even though the resulting surface translation rule as
shown in Eq 12 (represented in its 5D version):
.
vi0 ¼ ~
~ ~0 j Dri Þ  b
n0  Dri  ðjb ~0 (12)
i i

is an improvement over the constant ci models such as Mróz and most implementa-
tions of Armstrong-Frederick, it is unable to predict MR rate decay and arrest. This
model predicts a short ratcheting transient followed by a constant ratcheting rate that
never decays, overestimating its effects in multiaxial experiments. Chaboche’s model
is also a particular case of Eq 9 and Eq 10, for vi ¼ 1, mi ¼ 0, and ci ¼ di ¼ 1.
Burlet and Cailletaud noticed that multiaxial experiments generally show low-
er ratcheting rates than the uniaxial ones for equivalent conditions on stress or
strain amplitudes [20]. To lower the MR rate predictions without altering the uni-
axial response, they replaced the dynamic recovery term with a radial return term.
In the 5D formulation adopted in this work, their surface translation direction
becomes the form of Eq 13:

vi0 ¼ ~
~ ~0 T  ~
n0  Dri  ci  ðb n0 Þ  ~
n0 : (13)
i

Burlet-Cailletaud’s strain-hardening model is also obtained from the proposed


Eq 9 and Eq 10, assuming vi ¼ mi ¼ di ¼ 0 and an adjustable 0  ci  1.
The product b ~0 T  ~
n0 used in the radial return term measures the non-coaxiality
i
between the surface back stress b ~0 and the plastic strain increment direction ~
n0 . As a
i
result, it is a measure of the nonproportionality of the loading because parallel b ~0
i
T
and ~ 0
n usually found in proportional loadings result in b ~ ~
0 0
n ¼ 6jb~ j, while 90o
0
i i
out-of-phase loadings where plastic straining happens in a direction ~ n0 perpendicu-
~0 ~0T 0
lar to the surface back stress bi gives bi  ~ n ¼ 0. Such different products allow the
model to predict NP effects in MR.
234 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

Ohno and Wang used the nonproportionality product (calculated in our 5D


general formulation from b ~0 T  ~
n0 ) in a different way [16,21]. For plastic straining in
i
a direction ~ 0
n that makes an obtuse angle with b ~0 , that is, when b ~0 T  ~n0 < 0 (usually
i i
during an elastoplastic unloading process), they assumed that the translation direc-
tion follows Prager’s linear rule ~ vi0 ¼ ~n0  Dri . Otherwise, when b ~0 T  ~n0 0 (usually
i
during an elastoplastic loading process), they introduced in their model a scalar
function (given by 0  vi   ðjb ~0 j=Dri Þvi  1 in the proposed 5D formulation)
i .
and a nonproportionality term (defined in 5D as 0  ðb ~0 T  ~
n0 jb ~0 jÞ  1), result-
i i
ing in the “Ohno-Wang II” (OW-II) surface translation direction, whose 5D version
in the E5s stress space becomes
 . v  T . 
vi0 ¼ ~
~ ~0 j Dri i  b
n0  Dri  jb ~0  ~ ~0 j  b
n0 jb ~0 ð5D Ohno  Wang IIÞ (14)
i i i i

where vi (0  vi  1) is the ratcheting exponent. Surfaces calibrated with a very large


vi (such as in their “Ohno-Wang I” [OW-I] model version that assumes vi ! 1)
have vi ffi vi0 ¼ ~
. 0, which results in Prager’s linear rule ~ n0  Dri for most of the range
~ 0
0  jb i j Dri  1 before saturation, becoming unable to predict UR. The dynamic
recovery term would only.be activated when the surfaces are closer to contacting each
~0 j Dri ffi 1. Lower calibrated values of vi, on the .
other, that is, when jb other hand,
i
allow the dynamic recovery term to be partially operative in the entire jb ~0 j Dri range,
i
increasing the predicted UR rates. So, in summary, lower calibrated values of vi result
in higher UR rate predictions. Note that both OW-I and OW-II models are a particu-
lar case of the proposed Eq 9 and Eq 10, adopting mi ¼ ci ¼ di ¼ 1, and an adjustable
0  vi < 1 that tends to infinity for the OW-I model.
However, the ratcheting parameter vi influences both UR and MR predictions.
When vi is calibrated to fit UR data, the OW-II model ends up overestimating MR.
In addition, although the OW-II model is able to predict MR rate decay, the use of
a single calibration parameter vi renders it unable to model experiments with con-
stant MR rates.
Jiang and Sehitoglu [14,15] improved the OW-II model to.solve this last prob-
lem by simply removing the nonproportionality term (b ~0 T  ~
n0 jb~0 j in our 5D for-
i i
mulation) from its translation rule. When represented in the proposed E5s stress
space, Jiang-Sehitoglu’s surface translation direction is expressed as Eq 15:
.
vi0 ¼ ~
~ ~0 j Dri Þvi  b
n0  Dri  ðjb ~0 (15)
i i

where 0  vi < 1. This equation is used even during an elastoplastic unloading,


~0 T  ~
b n0 < 0, instead of switching to Prager’s linear rule as it had been proposed by
i
Ohno and Wang.
Jiang-Sehitoglu’s equation is also a particular case of Eq 9 and Eq 10, for mi ¼ 0
and ci ¼ di ¼ 1, similar to Chaboche’s model [12], and with an adjustable
0  vi < 1. As a result, Jiang-Sehitoglu’s equation is a generalized version of Cha-
boche’s original model [12], which would be obtained for the particular case vi ¼ 1.
Because Chaboche’s model has the ability to predict constant ratcheting rate for
WU AND ZHONG, DOI 10.1520/STP159820160059 235

both uniaxial and multiaxial loadings, Jiang-Sehitoglu’s equation overcomes the in-
ability of the OW-II model to predict constant MR rates. MR rate decay can also be
predicted if a ratcheting rate exponent vi = 1 is chosen in the calibration. Neverthe-
less, Jiang-Sehitoglu’s model still relies on a single calibration parameter vi to pre-
dict both UR and MR rates.
Calibrating a kinematic hardening model using different parameters to inde-
pendently control UR and MR allows for a much better description of the material
behavior. This separation is necessary because both ratcheting types are caused by
different phenomena: UR is a consequence of anisotropy between the tension and
compression behaviors, while MR is associated with elastoplastic deviatoric stress
increments d~ s 0 that are not parallel to the normal ~
n0 to the yield surface at the cur-
rent state, causing plastic strains not only in the direction of d~ s 0 but also ratcheting
strains in perpendicular directions. For instance, a material with a significant
strength difference between tension and compression could have almost the same
MR behavior as a perfectly isotropic one, even though only the former could suffer
UR. It would be impossible to accurately calibrate both independent behaviors with
a single scalar parameter for each surface, such as vi.
Because Armstrong-Frederick’s model largely underpredicts while Burlet-Cailletaud’s
largely overestimates MR rates, Delobelle, Robinet, and Bocher [17] decided to interpolate
them using an MR coefficient di (0  di  1). In the proposed E5s space, Delobelle’s sur-
face translation direction is shown as Eq 16:

vi0 ¼ ~
~ n0  Dri  ci  ½di  b ~0 T  ~
~0 þ ð1  di Þ  ðb n0 Þ  ~
n0 : (16)
i i

The limit value di ¼ 0 gives the Burlet-Cailletaud model, associated with a large
radial return term and zero dynamic recovery, a “radial evanescence” condition
that results in low MR rates with large rate decay. The other limit value di ¼ 1 gives
the Armstrong-Frederick model, with a large dynamic recovery term and zero radi-
al return, the usual “back stress evanescence” condition that results in overesti-
mated MR without rate decay, as discussed before. If 0 < di < 1, then the
predictions are somewhere in between the two limit cases, with di acting as a
weighting factor to calibrate the MR rates. The value of ci influences both uniaxial
and MR estimations. However, the UR response is not affected by di because, for
uniaxial loadings, b ~ 0 and ~ n0 are always parallel to the uniaxial direction; therefore,
i
T
~0  ~
the relation ci  ðb n0 Þ  ~ ~0 causes a translation direction (Eq 17):
n0 ¼ ci  b
i i

vi0 ¼ ~
~ ~0 ¼ ~
n0  Dri  ci  ½di þ ð1  di Þ  b ~0
n0  Dri  ci  b (17)
i i

that is independent of di. So, to fit UR data, ci must be calibrated first for every surface
and, after that, the di values can be freely calibrated to MR data without affecting the
previous uniaxial calibration. Delobelle’s model is obtained from the generalized sur-
face translation rule from Eq 9 and Eq 10 for vi ¼ mi ¼ 0 and from adjustable
0  ci  1 and 0  di  1 to independently calibrate UR and MR. Note however that,
similar to both the Armstrong-Frederick and Burlet-Cailletaud equations, the
236 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

Delobelle model still overpredicts UR rates because the ratcheting coefficient ci is not
as efficient as the ratcheting exponent vi to model UR rate decay or even growth as a
function of the stress amplitude.
On the other hand, the Chen-Jiao-Kim model [22], in addition to using the bet-
ter parameter vi to calibrate UR, is able to independently calibrate UR and MR
behaviors . by incorporating an MR exponent mi in the nonproportionality term
~0 T  ~
b n 0 ~0 j from the OW-II model. Therefore, for a multiaxial elastoplastic load-
jb
i i
 T . mi
ing process with b ~0 T  ~ ~0  ~
n0 0, the scalar function mi   b ~0 j
n0 j b is used
i i i

to multiply the dynamic recovery term. In the proposed 5D formulation, Chen-


Jiao-Kim’s surface translation direction is shown in Eq 18:
 . v  T . mi
vi0 ¼ ~
~ ~0 j Dri i  b
n0  Dri  jb ~0  ~ ~0 j
n0 j b ~0
b (18)
i i i i

where 0  vi < 1 and 1 < mi < 1. For an elastoplastic unloading process with
~0 T  ~
b n0 < 0, Prager’s linear rule ~vi0 ¼ ~
n0  Dri is used instead. For uniaxial load histo-
i
ries during elastoplastic loading, where the relation b ~0 T  ~
n0 ¼ jb~0 j is always valid, the
i i
  mi
scalar function mi simply becomes mi ¼ ð1Þ ¼ 1; therefore, UR predictions are
not affected by the calibrated value of mi. Thus, vi should be calibrated first for the
yield and every hardening surface to fit UR rate data, and after that the mi values
could be freely calibrated to correctly describe measured MR rates and decays without
affecting the previous uniaxial calibration. Chen-Jiao-Kim’s model can also be
obtained from the general Eq 9 and Eq 10, adopting ci ¼ di ¼ 1 and independently
adjustable 0  vi < 1 and  1  mi  1.
A different approach for obtaining a simultaneous correct description of UR
and MR was adopted in the Chen-Jiao model [23] (Eq 19). This model uses Delo-
belle’s [17] MR coefficient di (0  di  1) instead of the MR exponent mi, incorpo-
rated into Jiang-Sehitoglu’s model to give, in the adopted E5s space version,
.
vi0 ¼ ~
~ ~0 j Dri Þvi  ½di  b
n0  Dri  ðjb ~0 T  ~
~0 þ ð1  di Þ  ðb n0 Þ  ~
n0 : (19)
i i i

As in Delobelle’s model, di can calibrate MR data without affecting UR calcula-


tions. The exponent vi (0  vi < 1) of the yield or every hardening surface should
be calibrated first to accurately match UR data, and only then the di should be fit-
ted to describe MR rates and decay. Chen and Jiao also refined the MR description,
allowing the di parameter from each surface to vary between an initial value and a
target value dti, with an evolution equation ddi ¼ ðdti  di Þ  bCJ  dp controlled by
the equivalent plastic strain increments dp, where bCJ is the Chen-Jiao evolution
rate. Note, however, that this refinement introduces the additional parameters
dti (one for each surface i) and bCJ, which would need to be calibrated in proper
tests. Finally, note that Chen-Jiao’s model is also a particular case of the proposed
Eq 9 and Eq 10 adopting mi ¼ ci ¼ 1 and independently adjustable 0  vi < 1 and
0  di  1.
Note that independent calibration of UR and MR rates can only be achieved
using equations with at least two parameters per surface (i.e., a total of at least 2M
WU AND ZHONG, DOI 10.1520/STP159820160059 237

parameters for M yield and hardening surfaces), such as the Delobelle, Chen-Jiao,
and Chen-Jiao-Kim equations. However, Delobelle’s model still overpredicts UR
rates, due to the use of the ratcheting coefficient ci instead of the better ratcheting
exponent vi to calibrate them. Nevertheless, if the studied load history only causes
significant UR or MR, but not both, then Jiang-Sehitoglu’s equation would also be a
good modeling choice because it can calibrate arbitrary UR or MR rates, including
MR with constant rate or rate decay, using only the M ratcheting exponents vi from
the M surfaces, without requiring the (possibly less robust) calibration of 2M or
more parameters.
The fitting of the generalized plastic modulus coefficients pi from each surface
for a given ri as well as the calibration of the ratcheting coefficients vi, mi, ci, or di
(or a combination thereof), depend on the adopted NLK model. Approximate fit-
ting algorithms for the parameter pairs (pi, vi) or (Dri, vi) are shown for the Jiang-
Sehitoglu model [15], which can be easily adapted from the 6D to the proposed 5D
formulation; however, they are precise only for materials with very large ratcheting
exponents vi.

ISOTROPIC AND NP HARDENING FORMULATION


Isotropic hardening (or softening) accounts for the changes in the material yield
strength associated with the expansion (or contraction) of the yield surface while
gradually changing from the monotonic radius S ¼ SY to the cyclic radius S ¼ SYc.
In addition, NP hardening accounts for the increase in the material yield
strength associated with the expansion of the yield surface caused by NP loading
histories in some materials, especially austenitic stainless steels, increasing the radi-
us toward a target value SYc(1 þ aNPFNP), where 0  FNP  1 is a load-path-
dependent NP factor, while 0  aNP  1 is the material-dependent additional
hardening coefficient.
In this work, instead of changing the yield surface radius, isotropic and NP
hardening are accounted for by gradually changing the generalized plastic modu-
lus coefficients pi, while keeping constant the yield surface radius and all satura-
tion values Dri. This approach is much simpler to implement in a computer code
because a varying yield surface radius would result in a more complex consisten-
cy condition (e.g., a fading function of memorization is used to take into account
the effect of plastic strain amplitude on isotropic hardening) [24]. Moreover, this
approach results in essentially the same stress-strain predictions as the ones with
varying radius, as long as the yield surface is defined for a low plastic strain level,
typically 0.02 % instead of the usual 0.2 %. Because in the NLK formulation the
material behavior inside the yield surface is assumed linear elastic, it is recom-
mended to adopt a yield surface defined between 0.02 % and 0.002 % plastic
strain levels to avoid a discontinuous derivative in the calculated stress-strain
hysteresis loops. Such a yield surface can then be regarded as an elastic limit of
the material, whose radius is usually much smaller that the monotonic or cyclic
yield strengths.
238 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

In the formulation proposed in this work, to account for isotropic and NP


hardening, the generalized plastic modulus coefficients pi(p) evolve from their initial
values pi(0) (from a virgin material) through the Eq 20:
8 91=hc
>
< >
=
pi ðpÞ ¼ pi ð0Þ  ½1 þ aNP  FNP ðpÞ þ ðSY =SYc  1Þ  ehrc p (20)
:|fflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflffl} |fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}>
> ;
NP evolution isotropic evolution

where p  dp is the accumulated plastic strain, hrc is an isotropic hardening rate (typi-
cally between 0.5 and 50 for metals), hc is Ramberg-Osgood’s exponent, and the scalar
function FNP(p) is the transient value of the load-path-dependent NP factor.
To compute Eq 14 for tension-torsion histories, the evolution of FNP(p) is calculated
using a 2D version of Tanaka’s NP hardening model [25], adopting a symmetric 2 by 2
(instead of the usual 5 by 5) polarization tensor [PT], initially equal to zero and with an
evolution equation dictated by the tensor hardening rate hrT through Eq 21:

Pr Prs
½dPT  ¼ ð~ nT  ½PT Þ  hrT  dp; where ½PT  
n ~ : (21)
Prs Ps

If the current plastic straining direction is defined as ~ n  ½ nr ns T , then the


scalar evolution equations for the elements from [PT] can be written as dPr ¼
ðnr 2  Pr Þ  hrT  dp, dPs ¼ ðns 2  Ps Þ  hrT  dp, and dPrs ¼ ðnr ns  Prs Þ  hrT  dp.
Then, from the aforementioned 2D formulation, it is possible to show that
Tanaka’s NP hardening model results in an evolution equation for the transient NP
factor FNP(p) given by Eq 22:
8
>
> dF ðpÞ ¼ ½FNPt  FNP ðpÞ  hrNP  dp;
< NP sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

2Pr 2 ns 2 þ 2Prs 2 þ 2Ps 2 nr 2  4Prs ðPr þ Ps Þnr ns (22)
>
>
: FNPt ¼ Pr 2 þ 2Prs 2 þ Ps 2

where hrNP is a NP hardening rate that dictates how fast the yield surface radius
changes from its current value to its NP-hardened target value FNPt. It is recom-
mended to calibrate Tanaka’s hardening rates hrT and hrNP, satisfying the restriction
hrT  hrNP  2.5hrT, to avoid predicting a transient FNP(p) greater than 1.0, as ver-
ified from extensive simulations.

PROPOSED TRANSLATION RULE


The translation of the yield surface is mainly dominated by the kinematic hardening
model during a plastic strain increment. Summarizing the translation models dis-
cussed here, simultaneously acceptable predictions of both uniaxial and multiaxial
responses are difficult to accomplish. The vi in the Jiang-Sehitoglu model and the di
in the Chen-Jiao or the mi in the Chen-Jiao-Kim model may not lead to
WU AND ZHONG, DOI 10.1520/STP159820160059 239

improvement of MR predictions as displayed in the next section. In order to over-


come this shortcoming, a new translation rule based on the Jiang-Sehitoglu and
Chen-Jiao models has been proposed by introducing an additional multiaxial factor
ki to consider the NP effect and a scalar function Fi that characterizes both the
Jiang-Sehitoglu and Chen-Jiao models, as expressed as in Eq 23:


0  
0 !T ~0 
~vi ¼ k^i 1  abs b i  ~ n =b i

 . vi  
0T
n0  Dri  jb
~ ~0 þ ð1  Fi Þ  !
~0 j Dri  Fi  b b i ~
n ~n ð23Þ
i i

.
0
0
~0 !T ! !T !
where Fi ¼ jb i j Dri and abs b i  n means absolute value of b i  n . The


0  
! T ! !0 
term 1  abs b i  n = b i  can be interpreted as a qualitative measure on mul-

0
!T
tiaxial loading; that is, for uniaxial loading, because abs b i  ! n has the same
 0
! 
direction and value of  b i , the term will become unified and the factor ki will be inac-
tive. As a result, with initial ki ¼ [1,1,1,1,1], the proposed model generates the Jiang-
Sehitoglu model for Fi ¼ 1 and Chen-Jiao model for Fi ¼ di. For multiaxial loading, the
multiaxial parameter ki will influence the ratcheting response. The proper value of vi
can be confirmed by calibrating the experimental data of UR, and the calibration of ki
can be performed after the calibration of vi according to MR data.

Experimental Results and Discussion


To verify the prediction capabilities of the presented numerical framework, tension-torsion
experiments are performed on annealed tubular 316L stainless steel specimens in an MTS
809.25 multiaxial testing machine, shown in Fig. 3. A thin wall of 1.5 mm is usually adopted
in tubular specimens, to avoid having to deal with stress gradient effects across the
thickness. But, since some experiments in this work involve large compression strains, the
minimum wall thickness is increased from 1.5 to 2.0 mm to avoid buckling. On the critical
section, the specimen has an external diameter dext ¼ 16 mm and internal dint ¼ 12 mm.
Engineering stresses and strains are calculated from load/torque cell measurements and
from an MTS 632.68 axial/torsional extensometer and then converted to true stresses and
strains. The engineering shear stresses include the elastoplastic gradient correction
recommended by ASTM E2207-08, Standard Practice for Strain-Controlled Axial-Torsional
Fatigue Testing with Thin-Walled Tubular Specimens [26].
The cyclic properties of this 316L steel are obtained from uniaxial tests, result-
ing in a fitted Ramberg-Osgood uniaxial cyclic hardening coefficient Hc ¼ 874 MPa
and exponent hc ¼ 0.123, with Young’s E ¼ 193 GPa, Poisson ratio n ¼ 0.3, and
G ¼ E/(2 þ 2n) % 74GPa. To improve the calculation accuracy, the back stress is
divided into M ¼ 5 additive components, following Chaboche’s idea [12]. The
proposed hardening model [14] is adopted, using its ci ¼ di ¼ 1, mi ¼ 0, and the
240 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 3 Tension-torsion testing machine and extensometer mounted on a tubular


specimen.

parameter calibration procedure described by Jiang and Sehitoglu [15]: For a chosen
set of generalized plastic modulus coefficients pi ¼ {6176, 786, 100, 12.7, 1.62} MPa,
the resulting saturation values for the back stress components become Dri % {66, 85,
109, 141, 217} MPa. UR experiments are used to calibrate the adjustable exponents
vi ¼ {1, 1, 2.9, 3, 4}, using an RMS fit with respect to the corresponding simulations.
The multiaxial parameter ki ¼ {12 20 50 100 13}.
Strain-controlled tension–torsion tests are then performed adopting strain
paths with 0.02/s strain rate describing elliptical, circular, and crossed paths in the
rxsxyH3 diagram. Fig. 4a–c shows applied elliptical, circular, and crossed strain
paths with different amplitudes. Fig. 4d shows the uniaxial static stress-strain rela-
tionship. Fig. 4e shows the stress-strain relationship for MSR. Fig. 4f shows the
stress-strain relationship for UR. Fig. 4g shows the stress-strain relationship for MR.
The corresponding predictions performed using the presented formulation of Tana-
ka’s NP hardening model together with Jiang-Sehitoglu’s kinematic hardening
model (with ki ¼ [1 1 1 1 1]), result in a good agreement except in the MR case.
Note that the numerical results according to Chen-Jiao and Chen-Jiao-Kim’s model
also have significant deviations in the estimation.
The proposed model has also been applied and the comparisons are shown in
Fig. 5. Compared with the models discussed so far, this model involving additional
multiaxial parameters ki could give more reasonable predictions in the crossed
WU AND ZHONG, DOI 10.1520/STP159820160059 241

FIG. 4 Comparisons of the cyclic strain–stress curves experimentally measured (light blue
lines) and predicted using the simulator (dark red lines) based on Jiang-Sehitoglu’s
model: (a) rxsxyH3 stress paths (response for strain-controlled elliptical path with
axial strain amplitudes 0, 0.2 %, 0.4 %, 0.6 %, 0.7 %, and 0.8 %); (b) rxsxyH3 stress
paths (response for strain-controlled circular path with amplitudes 0.2 %, 0.4 %,
0.6 %, 0.8 % and 1.0 %); (c) rxsxyH3 stress paths (response for strain-controlled
crossed path with amplitudes 0.2 %, 0.4 %, 0.6 % and 0.8 %); (d) static test results;
(e) MSR results; (f) UR results; and (g) MR results.

(a) (b)

(c) (d)

(e) (f)

(g)
242 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 5 Comparisons of the cyclic strain–stress curves experimentally measured (light blue
lines) and predicted using the simulator (dark red lines) based on the proposed
model: (a) rxsxyH3 stress paths (response for strain-controlled elliptical path with
axial strain amplitudes 0, 0.2 %, 0.4 %, 0.6 %, 0.7 %, and 0.8 %); (b) rxsxyH3 stress
paths (response for strain-controlled circular path with amplitudes 0.2 %, 0.4 %,
0.6 %, 0.8 % and 1.0 %); (c) rxsxyH3 stress paths (response for strain-controlled
crossed path with amplitudes 0.2 %, 0.4 %, 0.6 % and 0.8 %); (d) static test results;
(e) MSR results; (f) UR results; and (g) MR results.

(a) (b)

(c) (d)

(e) (f)

(g)
WU AND ZHONG, DOI 10.1520/STP159820160059 243

path case (Fig. 5c) and especially in the MR case (Fig. 5g) with respect to Jiang-
Sehitoglu’s model.

Conclusions
In this paper, an evaluation has been made of the Jiang-Sehitoglu model. This mod-
el is found to perform reasonably well for all considered uniaxial conditions and
most multiaxial conditions—except the MR case. It is difficult to simulate simulta-
neously the responses of UR and MR as well as UMR. A general NLK hardening
model based on the Jiang-Sehitoglu and Chen-Jiao models has been proposed that
is able to give more reasonable simulation in an MR condition than other models.
Several non-proportional tension-torsion experiments with 316L steel tubular
specimens validated the proposed model.

ACKNOWLEDGMENTS
This work was supported by the National Natural Science Foundation of P.R. China
under Grant No. 11302150.

References

[1] Meggiolaro, M. A., Castro, J. T. P., and Wu, H., “A General Class of Non-Linear Kinematic
Models to Predict Mean Stress Relaxation and Multiaxial Ratcheting in Fatigue Problems—
Part I: Ilyushin Spaces,” Int. J. Fatigue, Vol. 82, Part 2, 2016, pp. 158–166.

[2] Masing, G., Eigenspannungen and Verfestigung Beim Messing (“Self Stretching and
Hardening for Brass”), Proceedings of the Second International Congress of Applied
Mechanics, Zurich, Switzerland, 1926, pp. 332–335.

[3] Arcari, A., De Vita, R., and Dowling, N. E., “Mean Stress Relaxation During Cyclic Straining
of High Strength Aluminum Alloys,” Int. J. Fatigue, Vol. 31, Nos. 11–12, pp. 1742–1750.

[4] Landersheim, V., Bruder, T., and Hanselka, H., “Approximation of Mean Stress Relaxation
by Numerical Simulation Using the Jiang Model and Extrapolation of Results,” Procedia
Eng., Vol. 10, 2011, pp. 595–600.

[5] Mróz, Z., “On the Description of Anisotropic Work Hardening,” J. Mech. Phys. Solids,
Vol. 15, No. 3, 1967, pp. 163–175.

[6] Garud, Y. S., “A New Approach to the Evaluation of Fatigue under Multiaxial Loading,”
J. Eng. Mater-T. ASME, Vol. 103, No. 2, 1981, pp. 118–125.

[7] Krieg, R. D., “A Practical Two-Surface Plasticity Theory,” J. Appl. Mech., Vol. 42, No. 3,
1975, pp. 641–646.

[8] Dafalias, Y. F. and Popov, E. P., “Plastic Internal Variables Formalism of Cyclic Plasticity,”
J. Appl. Mech., Vol. 43, No. 4, 1976, pp. 645–651.

[9] Chaboche, J. L., “A Review of Some Plasticity and Viscoplasticity Constitutive Theories,”
Int. J. Plast., Vol. 24, No. 10, 2008, pp. 1642–1693.
244 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

[10] Jiang, Y. and Sehitoglu, H., “Comments on the Mróz Multiple Surface Type Plasticity
Models,” Int. J. Solids Struct., Vol. 33, No. 7, 1996, pp. 1053–1068.

[11] Armstrong, P. J. and Frederick, C. O., A Mathematical Representation of the Multiaxial


Bauschinger Effect, CEGB Report RD/B/N731, Berkeley Nuclear Laboratory, Gloucester-
shire, UK, 1966.

[12] Chaboche, J. L., Dang Van, K., and Cordier, G., Eds., “Modelization of the Strain Memory
Effect on the Cyclic Hardening of 316 Stainless Steel,” Transactions of the Fifth Interna-
tional Conference on Structural Mechanics in Reactor Technology, North-Holland
Publishing Co., The Netherlands, 1979.

[13] Meggiolaro, M. A., Castro, J. T. P., and Wu, H., “A General Class of Non-Linear Kinematic
Models to Predict Mean Stress Relaxation and Multiaxial Ratcheting in Fatigue Prob-
lems,” presented at the International Conference on Fatigue Damage of Structural
Materials X, Hyannis, MA, September 21–26, 2014.

[14] Jiang, Y. and Sehitoglu, H., “Modeling of Cyclic Ratchetting Plasticity, Part I: Develop-
ment of Constitutive Relations,” ASME J. Appl. Mech., Vol. 63, No. 3, 1996, pp. 720–725.

[15] Jiang, Y. and Sehitoglu, H., “Modeling of Cyclic Ratchetting Plasticity, Part II: Comparison
of Model Simulations with Experiments,” ASME J. Appl. Mech., Vol. 63, No. 3, 1996,
pp. 726–733.

[16] Ohno, N. and Wang, J. D., “Kinematic Hardening Rules with Critical State of Dynamic
Recovery, Part I: Formulation and Basic Features for Ratchetting Behavior,” Int. J. Plas-
ticity, Vol. 9, No. 3, 1993, pp. 375–390.

[17] Delobelle, P., Robinet, P., and Bocher, L., “Experimental Study and Phenomenological
Modelization of Ratchet under Uniaxial and Biaxial Loading on an Austenitic Stainless
Steel,” Int. J. Plasticity, Vol. 11, No. 4, 1995, pp. 295–330.

[18] Prager, W., “Recent Developments in the Mathematical Theory of Plasticity,” J. Appl.
Phys., Vol. 20, No. 3, 1949, pp. 235–241.

[19] Abdel Karim, M. and Ohno, N., “Kinematic Hardening Model Suitable for Ratcheting with
Steady-State,” Int. J. Plasticity, Vol. 16, No. 3, 2000, pp. 225–240.

[20] Burlet, H. and Cailletaud, G., “Numerical Techniques for Cyclic Plasticity at Variable
Temperature,” Eng. Computation, Vol. 3, No. 2, 1986, pp. 143–153.

[21] Ohno, N. and Wang, J. D., “Kinematic Hardening Rules with Critical State of Dynamic
Recovery, Part II: Application to Experiments of Ratchetting Behavior,” Int. J. Plasticity,
Vol. 9, No. 3, 1993, pp. 391–403.

[22] Chen, X., Jiao, R., and Kim, K. S., “On the Ohno-Wang Kinematic Hardening Rules for
Multiaxial Ratcheting Modeling of Medium Carbon Steel,” Int. J. Plasticity, Vol. 21, No. 1,
2005, pp. 161–184.

[23] Chen, X. and Jiao, R., “Modified Kinematic Hardening Rule for Multiaxial Ratcheting
Prediction,” Int. J. Plasticity, Vol. 20, Nos. 4–5, 2004, pp. 871–898.

[24] Kang, G., Gao, Q., and Yang, X., 2002, “A Visco-Plastic Constitutive Model Incorporated
with Cyclic Hardening for Uniaxial/Multiaxial Ratcheting of SS304 Stainless Steel at
Room Temperature,” Mech. Mater., Vol. 34, No. 9, pp. 521–531.
WU AND ZHONG, DOI 10.1520/STP159820160059 245

[25] Tanaka, E., “A Non Proportionality Parameter and a Cyclic Viscoplastic Constitutive Mod-
el Taking into Account Amplitude Dependencies and Memory Effects of Isotropic Hard-
ening,” Eur. J. Mech. A-Solid, Vol. 13, No. 2, 1994, pp. 155–173.

[26] ASTM E2207-08, Standard Practice for Strain-Controlled Axial-Torsional Fatigue Testing
with Thin-Walled Tubular Specimens, ASTM International, West Conshohocken, PA,
2013, www.astm.org
246 FATIGUE AND FRACTURE TEST PLANNING, TEST DATA ACQUISITIONS AND ANALYSIS

STP 1598, 2017 / available online at www.astm.org / doi: 10.1520/STP159820160024

Grzegorz Socha1

Fatigue Damage Indicators Based


on Plastic Deformation
Citation
Socha, G., “Fatigue Damage Indicators Based on Plastic Deformation,” Fatigue and Fracture Test
Planning, Test Data Acquisitions and Analysis, ASTM STP1598, Z. Wei, K. Nikbin, P. C. McKeighan, and
D. G. Harlow, Eds., ASTM International, West Conshohocken, PA, 2017, pp. 246–257, http://dx.doi.org/
10.1520/STP1598201600242

ABSTRACT
Fatigue failure of elastic-plastic materials is the result of damage accumulation.
In the initial phase, this process is controlled by plastic deformation. The
standard method for the investigation of fatigue phenomena relies on counting
the applied load (or strain) cycles to final failure or macro crack initiation. This
method does not allow observations of fatigue damage accumulation during
testing. Such observations can only be realized when a measurable physical
quantity (representing damage) is selected for monitoring. Such a quantity,
referred to as the damage indicator, must be proven to change in a monotonic
manner with the progress of the damage. In this paper, two damage indicators
will be proposed and discussed: local inelastic response and accumulated
equivalent plastic strain. Local inelastic response can be referred to as the
relative damage indicator when it changes in a monotonic manner during the
entire life of the material. In the case of such an indicator, its actual value is
compared with its initial value to estimate the amount of accumulated damage.
This kind of damage indicator is useless in the case of materials exhibiting more
complex behavior (e.g., initial cyclic hardening followed by softening). For such
materials, accumulated equivalent plastic strain can be used as the damage
indicator. This kind of indicator can be referred to as a cumulative damage
indicator because it is necessary to use the cumulated value of such a quantity
to estimate the amount of accumulated damage.

Manuscript received January 28, 2016; accepted for publication May 26, 2016.
1
Institute of Aviation, Al. Krakowska 110/114, 02-256 Warsaw, Poland
2
ASTM Symposium on Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis on May 4–5,
2016 in Grand Hyatt, San Antonio, TX.

Copyright V
C 2017 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.
SOCHA, DOI 10.1520/STP159820160024 247

Keywords
fatigue damage, damage indicator, plastic strain

Introduction
Fatigue failure of engineering materials is preceded by accumulation of damage in vari-
ous forms. For elastic-plastic material, plastic deformation at the spots of stress concen-
tration (grain boundary, hard inclusions, etc.) is followed by initiation of physical
discontinuities. Growth of physical discontinuities (micro cracks) leads to formation of a
dominating crack. Eventually, the propagating dominating crack reaches a size for which
the stress intensity factor (SIF) is big enough to trigger unstable crack growth under a ser-
vice load. Quantification of damage should reflect all the aforementioned phenomena
taking place during damage accumulation. One instructive review paper [1] on damage
quantifying parameters divides damage indicators into the following categories:

• Metallurgical parameters measured on metallographic samples


• Surface crack quantifications measured on materials’ surface
• Physical measures (used mostly for nondestructive testing [NDT] inspections)
• Mechanical parameters measured in a strength laboratory

In the case of the first two categories, damage indicators (such as dislocation or slip
band density, crack summary length, or area) are very difficult (or even impossible) to
measure. An attempt to measure dislocation density and its evolution due to fatigue load-
ing was described by Petersmeier et al. [2]. This kind of investigation is rather rare due to
difficulties that are associated with dislocation density determination for the representa-
tive volume element (RVE) of the material. More papers regarding crack density (surface
or internal) can be found in the literature. Most of them discuss surface crack investiga-
tion [3], but some analyze internal micro crack density surface or length [4,5]. Another
difficulty associated with this kind of damage indicator is the fact that different indicators
should be applied to different phases of damage accumulation.
Optical, acoustic, thermal, electrical, or magnetic quantities (physical measures)
often used for NDT inspections are focused mostly on detection of physical discon-
tinuities. There have been many attempts to use physical quantities for measure-
ment of damage in an early stage, before a dominating crack is formed. An example
of investigations that use magnetic measurement can be found in a study by Moor-
thy et al. [6]; thermo-acoustic measurements of fatigue damage are presented by
Meyendorf et al. [7]. Some researchers have tried to quantify damage using a poten-
tial drop technique [8], and others have tried using micro hardness (semidestructive
method) to define damage parameters [9]. The credibility of such measurements is
based on the experimentally determined correlation between the aforementioned
quantities and damage introduced into a material in an early phase.
Damage indicators investigated in a strength laboratory include many material
properties measured during standard and non-standard tests. Typical examples are
248 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

elastic constants. It is assumed by many researchers that the progress of damage


influences elastic constants, and an appropriate continuum mechanics–based theo-
ry was developed to describe this phenomenon [10]. This theory, however, assumes
that physical discontinuities are present in the bulk of the material. This assumption
may not work for the early phase of the damage process, when damage is related to
changes in dislocation density and distribution. In this phase, plastic deformation
can be localized in the vicinity of various material imperfections in the case of high-
cycle fatigue testing, or all of the gage part of the specimen can yield, such as in the
case of low-cycle fatigue (LCF) testing.
Another often-used group of damage indicators is based on observations of ma-
terial response to applied load or deformation. In standard, strain-controlled LCF
testing, it is assumed that, after initial saturation, the hysteresis loop remains stable
until the end of the test. This assumption is not confirmed by experimental observa-
tions [11,12]. In reality, mechanical response of material to applied load or strain
cyclic load varies throughout the entire testing process due to accumulation of dam-
age, which then leads to final failure of the material.
In this paper, investigations of fatigue damage accumulation with the use of
two mechanical parameters (damage indicators) based on local inelastic response to
an applied constant-stress amplitude load program are presented. Proposed damage
indicators can be used for the initial phase of the fatigue damage accumulation pro-
cess. The last phase of the process, when a dominant crack is formed, is analyzed
using laws of fracture mechanics.
The advantage of using these kinds of damage indicators and damage parame-
ters is that we are able to properly define the model used for life prediction. The
damage curve (the plot of the damage parameter as the function of the life fraction)
unambiguously determines which fatigue damage accumulation law must be used
for credible life prediction. Using the linear damage rule (LDR) for life prediction
without experimental verification can lead to enormous errors. Another very
important aspect of using measureable damage indicators and parameters is the fact
that this enables the assessment of material degradation for components in service.

Experimental Technique
According to the continuum damage mechanics approach, a damage parameter is
defined for a selected plane of the representative volume element [10]. If this defini-
tion is to be used for experimental investigations of fatigue damage accumulation,
we develop a testing technique allowing us to monitor changes in the material re-
sponse for the selected plane of the specimen’s gage.
One of the specimens, according to ASTM E466-15, Standard Practice for
Conducting Force Controlled Constant Amplitude Axial Fatigue Tests of Metallic
Materials [13], which is recommended for fatigue testing, is the specimen with a
continuous radius between the ends of its gage portion (a so-called hourglass speci-
men). For such a specimen, the highest stress amplitude during cyclic loading is
SOCHA, DOI 10.1520/STP159820160024 249

achieved for the plane of the smallest cross-section located in the middle of the gage
portion. Therefore, the highest rate of the fatigue damage accumulation process is
expected to occur for this plane. An example of the specimen design used for some
of the investigations performed by the author of this paper is shown in Fig. 1 along
with the coordination system used for descriptions of the stress and strain states.
The specimen shown in Fig. 1 is mounted in the testing machine load train
using self-alignment, a patented gripping system [14]. An axial, fully reversible
(tensile-compressive) cyclic load is applied with a frequency of 20 Hz. All tests are
stress-controlled, and constant stress amplitude loading is applied. Because all the
testing is stress-controlled, there is no need to handle elastic or plastic strain during
testing. For the smallest cross-section of the specimen’s gage, transversal total strain
is measured using a commercially available diametral extensometer. Using an exper-
imentally determined transversal stiffness modulus (initial slope of axial stress mi-
nus transversal total strain curve), the plastic part of the transversal total strain
p
component ey is calculated as the difference between recorded total transversal
strain and elastic transversal strain (axial stress divided by transversal stiffness) for
each recorded data point of the hysteresis loop. Axial plastic strain is then calculated
using a simple formula (Eq 1), assuming material incompressibility and plastic isot-
ropy (coordination system shown in Fig. 1):

epx ¼ 2epy ¼ 2epz : (1)

Finally, the hysteresis loop is plotted for selected cycles of the load and two param-
eters (inelastic response ei and ratcheting strain em) are determined. A schematic of

FIG. 1 Design of specimen, dimensions in mm.


250 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

the recorded hysteresis loop is shown in Fig. 2. Changes in the hysteresis loop and posi-
tion from cycle N-1 to cycle N are exaggerated to illustrate possible variations of the
two parameters that will be investigated in the following paragraphs. More details
regarding this experimental technique can be found in a study by Socha [15].

Material under Investigation


The material under investigation was martensitic chromium-molybdenum alloy
steel P91 (9Cr-1Mo-V, according to ASTM A335/335M, Standard Specification for
Seamless Ferritic Alloy-Steel Pipe for High-Temperature Service), which is frequently
used for high-temperature applications such as hot steam pipelines in power
plants. The chemical composition of the alloy is presented in Table 1. This steel was
investigated in its as-delivered condition, with the following basic mechanical
properties: elasticity limit 455 MPa, 0.2 % offset; yield limit 671 MPa; and tensile
strength 900 MPa.

Inelastic Response as the Relative Damage


Indicator
Inelastic response of the elastic-plastic material is the result of the movement of
dislocations. In the case of cyclic loading, inelastic (or plastic) response is the width

FIG. 2 A scheme of the hysteresis loop changes during the fatigue test.
SOCHA, DOI 10.1520/STP159820160024 251

TABLE 1 Chemical composition of P91 steel.

C Mn Cr Mo V Ni Cu Si S

0.08–0.12 0.3–0.6 8–9.5 0.85–1.1 0.18–0.25 <0.4 <0.3 0.2–0.5 <0.01

Note: C ¼ carbon; Mn ¼ manganese; Cr ¼ chromium; Mo ¼ molybdenum; V ¼ vanadium;


Ni ¼ nickel; Cu ¼ copper; Si ¼ silicon; and S ¼ sulfur.

of the hysteresis loop. If the material is loaded below the yield limit, theoretically,
no inelastic response should be observed. Real materials always consist of imper-
fections in crystal structure such as inclusions, grain boundaries, interface
between phases, and so on. If we carefully observe the local response of the real
material, we can notice some inelastic response even if the material is loaded
below the yield point. This response can change due to accumulation of fatigue
damage in the form of changes in dislocation density, initiation of material
discontinuities, or other changes in the material structure occurring under an
applied, cyclic load. If constant stress amplitude is applied to the material, we
can use inelastic strain amplitude ei (see Fig. 2) as the measure of the inelastic
response. Changes in this quantity, referred to as the damage indicator, will re-
flect accumulation of fatigue damage in various forms. The damage parameter,
according to Lemaitre [10], is defined as the scalar variable bounded by 0 and 1.
Value 0 corresponds to the undamaged material, and Value 1 corresponds to the
failure of the material. If inelastic response changes in a monotonic manner, as
can be seen in Fig. 3, we can use it as the relative damage indicator. In this case,

FIG. 3 Inelastic response of the P91 steel as the function of the load cycle number.
252 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

assuming that the material is initially undamaged, the value of the damage pa-
rameter can be calculated using Eq 2:
ei  eið1Þ
DR ¼ (2)
eið f Þ  eið1Þ

where:
ei(f) ¼ final value of the inelastic response (corresponding to failure of the
specimen),
ei(1) ¼ initial cycle, and
ei ¼ actual value of inelastic response.
Changes in the damage parameter as a function of the life fraction, usually denoted
as the damage curve, are shown in Fig. 4.
It can be seen from Fig. 4 that, for two stress amplitudes (600 and 640 MPa), a
sudden change in the damage accumulation rate can be observed for the value 0.98
of the life ratio N/Nf. The observed change in the fatigue damage accumulation rate
is related to the creation of a dominant crack—after which the stress state in the
gage part of the specimen changes and uniform distribution of the stress in the
smallest cross-section can no longer be assumed. Propagation of the dominant
crack is governed by fracture mechanics laws and is due to the large stress concen-
tration in the vicinity of the crack tip, which is associated with plastic deformation.
Because of this fact, many researchers consider the creation of a dominant crack as
the end of material life.
Assuming that the creation of a dominant crack is the end of material life, we
can find the corresponding inelastic response ei(f) and, using Eq 2, prepare the

FIG. 4 Damage curve based on the relative damage indicator for the P91 steel.
SOCHA, DOI 10.1520/STP159820160024 253

damage curve for the first phase of the damage process (from the start of loading to
the creation of the dominant crack). This plot is shown in Fig. 5. In this figure, the
actual number of load cycles is normalized against the number of cycles corre-
sponding to the creation of the dominant crack Ni instead of normalizing it against
the number of cycles to failure Nf (see Fig. 4). As can be seen in Fig. 5, for the first
phase of the fatigue damage accumulation process, we can use the LDR to calculate
the life of the material subjected to a known history of loading; but if we need more
accurate life prediction, it is advisable to use one of the nonlinear models of fatigue
damage accumulation.

Accumulated Equivalent Plastic Strain as the


Cumulative Damage Indicator
The relative damage indicator discussed in the previous section can be used to
quantify damage in the case of monotonic changes in measured local inelastic re-
sponse during the entire life of the material. Only materials with such a behavior
damage parameter can be unambiguously determined using Eq 2. The material in-
vestigated in this paper exhibits softening during its entire life, but there are many
materials that exhibit mixed behavior (for example, initial hardening followed by
softening). In such cases, other kinds of damage indicators have to be used. In
Socha and Dietrich’s study [16], accumulated plastic strain is used as the cumulative
damage indicator. For a cyclic, axial load applied to the specimen shown in Fig. 1,

FIG. 5 Damage curve for the first phase of damage accumulation (before the dominant
crack is formed).
254 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

the accumulated equivalent plastic strain (damage indicator) at the smallest cross-
section of the gage part is determined using Eq 3:
ð ð ð 
 
u ¼ dep ¼ depx ¼ 2depy  (3)

Ð
where dep is the equivalent plastic strain integrated along the stress path in the stress
space. In this case, the stress path is simple: The stress changes from a minimum to
maximum value and, reaching that value, turns in the opposite direction along the
x axis. For one cycle of the load shown in Fig. 2, the increment of plastic strain is given
by the sum of double the inelastic response and ratcheting strain. For N cycles of load,
the accumulated equivalent plastic strain is given by Eq 4:

X
N     
u¼ 2eiðN Þ  þ emðN Þ  emðN1Þ  : (4)
n¼1

Modern testing machines equipped with digital control and data acquisition
systems even allow us to calculate the damage indicator given by Eq 4 online (dur-
ing the test). Changes in the cumulative damage indicator during the fatigue test,
calculated according to Eq 4, are shown in Fig. 5.
As can be seen in Fig. 6, the amount of accumulated equivalent plastic strain
(equivalent deformation obtained by integrating the plastic strain intensity incre-
ments along the deformation path) is lowest for the highest stress amplitude. This
is related to the fact that, for high-stress amplitude, damage accumulation is faster,
and the formation of micro discontinuities accelerates ratcheting. This ratcheting is

FIG. 6 Accumulated equivalent plastic strain (damage indicator) for the P91 steel as
the function of the load cycle number.
SOCHA, DOI 10.1520/STP159820160024 255

FIG. 7 Damage curve based on the cumulative damage indicator for the P91 steel.

the result of a reduction in the active cross-section area of the specimen’s gage and
the local concentration of stress in the vicinity of micro cracks. For this reason, the
total amount of accumulated equivalent plastic strain depends on the initial config-
uration of imperfections in the material’s structure and surface.
Because there is no accumulated plastic strain in the beginning of loading pro-
cess, a simple definition of damage parameter is assumed, using Eq 5:
ð
1
D ¼ p dep (5)
ef

p
where ef stands for final value of the accumulated equivalent plastic strain (corre-
sponding to the occurrence of the dominant crack).
The damage parameter calculated according to Eq 5 is shown in Fig. 7. As can
be seen in this figure, the damage accumulation in this case is nonlinear, and the
appropriate damage law must be applied to predict fatigue life.

Conclusions
Fatigue damage of elastic-plastic materials is impacted by plastic deformation.
Accumulation of damage is a local process that can be quantified only by using local
methods for measuring material properties. An approach for performing such
measurements is reported in one of the author’s papers [15], where the response of
the material to the applied cyclic load is measured for the smallest cross-section of
the hourglass specimen’s gage. Performed investigations prove that local inelastic
response can be used as the damage indicator in the case of materials exhibiting
256 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

cyclic softening during the entire life. This method fails in the case of materials exhib-
iting more complex deformation behavior (e.g., initial hardening followed by soften-
ing). In such cases, local accumulated equivalent plastic strain should be used as the
damage indicator. Additional conclusions drawn from this study are as follows:
• Defining measureable damage parameters allows calculation of the damage pa-
rameter using, for example, formulas given in this paper (Eq 2 and Eq 5).
• Analysis of changes in the damage parameter during the life of the material
allows the specimen’s fatigue life to be divided into two phases: initiation and
propagation of the macro crack.
• For the first phase of the process, a plot of the damage curve showing changes
in the damage parameter as a function of life fraction can be prepared.
• The damage curve illustrates the manner in which damage progresses for the
investigated material and allows for the selection of the proper accumulation
model for fatigue life calculation.

References

[1] Yang, L. and Fatemi, A., “Cumulative Fatigue Damage Mechanisms and Quantifying
Parameters: A Literature Review,” J. Test. Eval., Vol. 26, No. 2, 1998, pp. 89–100.

[2] Petersmeier, T., Martin, U., Eifler, D., and Oettelt, H., “Cyclic Fatigue Loading and Charac-
terization of Dislocation Evolution in the Ferritic Steel X22CrMoV121,” Int. J. Fatigue,
Vol. 20, No. 3, 1998, pp. 251–255.

[3] Yang, B., Feng, M. F., and Zhai, Z. Y., “Evolutionary Statistical Character of Fatigue Dam-
age of Smooth Surface Samples by an Effective Short Fatigue Crack Criterion,” Int. J.
Damage Mech., Vol. 19, No. 2, 2010, pp. 211–231.

[4] Hall, J. A., “Fatigue Crack Initiation in Alpha-Beta Titanium Alloys,” Int. J. Fatigue, Vol. 19,
No. 1, 1998, pp. S23–S37.

[5] Zhang, M., Yang, P., and Tan, Y., “Micromechanisms of Fatigue Crack Nucleation and
Short Crack Growth in a Low Carbon Steel Under Low Cycle Impact Fatigue Loading,”
Int. J. Fatigue, Vol. 21, No. 8, 1999, pp. 823–830.

[6] Moorthy, V., Choudhary, B. K., Vaidyanathan, S., Jayakumar, T., Bhanu Sankara Rao, K.,
and Raj, B., “An Assessment of Low Cycle Fatigue Damage Using Magnetic Barkhausen
Emission in 9Cr-1Mo Ferritic Steel,” Int. J. Fatigue, Vol. 21, No. 3, 1999, pp. 263–296.

[7] Meyendorf, N. G. H., Roesner, H., Kramb, V., and Sathish, S., “Thermo-Acoustic Fatigue
Characterization,” Ultrasonics, Vol. 40, Nos. 1–8, 2002, pp. 427–434.

[8] Zhang, S., Xia, Q., Li, W., and Zhou, X., “Ductile Damage Measurement and Necking Anal-
ysis of Metal Sheets Based on Digital Image Correlation and Direct Current Potential
Drop Methods,” Int. J. Damage Mech., Vol. 23, No. 8, 2014, pp. 1133–1149.

[9] Ganjiani, M., “Identification of Damage Parameters and Plastic Properties of an Aniso-
tropic Damage Model by Micro-Hardness Measurements,” Int. J. Damage Mech., Vol. 22,
No. 8, 2013, pp. 1089–1108.
SOCHA, DOI 10.1520/STP159820160024 257

[10] Lemaitre, J., A Course on Damage Mechanics, Springer-Verlag, Berlin, 1996.

[11] Wahi, R. P., Auerswald, J., Mukherji, D., Dudka, A., Fecht, H.-J., and Chen, W., “Damage
Mechanisms of Single and Polycrystalline Nickel Base Superalloys SC16 and IN738LC
Under High Temperature LCF Loading,” Int. J. Fatigue, Vol. 19, Supp. No. 1, 1998,
pp. S89–S94.

[12] Yang, X., Li, N., Jin, Z., and Wang, T., “A Continuous Low Cycle Fatigue Damage Model
and Its Application in Engineering Materials,” Int. J. Fatigue, Vol. 19, No. 10, 1998,
pp. 687–692.

[13] ASTM E466-15, Standard Practice for Conducting Force Controlled Constant Amplitude
Axial Fatigue Tests of Metallic Materials, ASTM International, West Conshohocken, PA,
2015, www.astm.org

[14] Socha, G. and Dietrich, L. 2008. Self-alignment gripping system. Polish Patent PL
199–326.

[15] Socha, G., “Experimental Investigations of Fatigue Cracks Nucleation, Growth, and
Coalescence in Structural Steel,” Int. J. Fatigue, Vol. 25, No. 2, 2003, pp. 139–147.

[16] Socha, G. and Dietrich, L., “A Fatigue Damage Indicator Parameter for P91 Chromium-
Molybdenum Alloy Steel and Fatigue Pre-Damaged P54T Carbon Steel,” Fatigue Fract
Engng Mater Struct, Vol. 37, No. 2, 2014, pp. 195–205.
258 FATIGUE AND FRACTURE TEST PLANNING, TEST DATA ACQUISITIONS AND ANALYSIS

STP 1598, 2017 / available online at www.astm.org / doi: 10.1520/STP159820160078

Shizhu Xing1 and Pingsha Dong1

A Fatigue Failure Mode Transition


Criterion for Sizing Load-Carrying
Fillet-Welded Connections
Citation
Xing, S. and Dong, P., “A Fatigue Failure Mode Transition Criterion for Sizing Load-Carrying
Fillet-Welded Connections,” Fatigue and Fracture Test Planning, Test Data Acquisitions and
Analysis, ASTM STP1598, Z. Wei, K. Nikbin, P. C. McKeighan, and D. G. Harlow, Eds., ASTM
International, West Conshohocken, PA, 2017, pp. 258–277, http://dx.doi.org/10.1520/
STP1598201600782

ABSTRACT
In load-carrying fillet-welded connections, two distinct fatigue failure modes are
possible depending upon fillet weld leg size and loading conditions. One is weld
toe cracking through base plate thickness and the other is through weld metal,
often referred to as weld root cracking. Weld root cracking mode has always
been a concern in design and analysis of load-carrying fillet-welded connections.
In the past, it has been noticed that weld root cracking mode could be avoided
by enlarging the fillet weld size or weld penetration, and numerous theoretical
studies had been performed to construct such a design reference to prevent
weld root cracking. However, the theoretically developed design reference
cannot include shop floor practices, which may result in unfavorable fillet weld
size design. In this study, experimentally based analysis on fatigue failure mode
transition behavior is performed based on a recent comprehensive fatigue
testing program in support of construction of lightweight ship structures
reflecting typical shop floor practices and through introducing a newly
developed analytical weld throat stress model and misalignment-induced stress
concentration factors. As a result, a critical weld size reflecting typical shop floor
practice is obtained. It is determined that the actual critical weld size obtained
from test data is much lower than the theoretical one due to the combination

Manuscript received March 21, 2016; accepted for publication August 3, 2016.
1
University of Michigan, 2600 Draper Dr., Ann Arbor, MI 48109
2
ASTM Symposium on Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis on May 4–5,
2016 in Grand Hyatt, San Antonio, TX.

Copyright V
C 2017 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.
XING AND DONG, DOI 10.1520/STP159820160078 259

effect of weld penetration and misalignments. Finally, a statistical procedure is


proposed to construct a mathematical model that is able to provide a confidence
level to avoid weld root cracking mode for designed fillet weld size.

Keywords
load-carrying fillet welds, fatigue failure, weld root cracking, weld toe cracking,
joint misalignment, stress concentration

Introduction
Fatigue crack development through weld metal originating at weld root, often re-
ferred to as weld root cracking or throat cracking, has always been a concern in de-
sign and analysis of load-carrying fillet-welded connections [1–8]. Such a fatigue
failure mode should be avoided for a number of reasons. Chief among them are: (a)
fatigue lives associated with weld root cracking tend to be significantly lower than
those associated with weld toe cracking, which can be attributed to the fact that
defects are often present at weld root, in addition to inherent variability in weld
throat size and penetration status—among other factors [5,6]; and (b) weld root
cracking is particularly problematic for computer-based structural fatigue life evalu-
ation because most of the existing fatigue analysis procedures, such as nominal
stress and hot spot stress methods [9–12], are not developed for treating throat
cracking, as discussed by Dong and Hong [13] and Dong [14]. There exists a great
deal of experimental evidence that once a fillet weld size reaches beyond a critical
size, fatigue failure mode tends to transition from weld root to weld toe failure
[3,5,6]. As a result, to prevent weld root cracking from occurring, both existing
weld sizing criteria and shop floor practices tend to encourage over-sized welds as
discussed recently by Nie and Dong [15] and Huang et al. [16], among others.
In addition to fillet weld size, it has been noticed by numerous researchers
[3,4,17] that weld penetration has a significant effect on failure mode transition
from weld root cracking to weld toe cracking. Gurney [3] suggested a family of
curves that relate critical weld size to weld penetration and base metal plate thick-
ness, derived by equating the fatigue lives of weld toe cracking to weld root cracking
based on fracture mechanics calculations. With respect to a specific plate thickness,
each curve indicates that the critical relative weld size s/t (s denotes weld leg size
and t denotes plate thickness) will drastically decrease as weld penetration increases.
In a similar approach, Noblett and Andrews [4] performed crack propagation cal-
culations for both weld toe and weld root cracking in fillet-welded specimens under
remote cyclic tension and bending. By equating the fatigue lives calculated for weld
root cracking and weld toe cracking, penetration effects on failure mode transition
similar to those reported by Gurney [3] were obtained. Recently, Xing, Dong, and
Threstha [17] suggested a closed-form weld throat stress model, on top of which
the critical plane can be explicitly located within fillet weld metal. In most studies
[such as 3,4], a hypothetical failure path was initially assumed without providing
260 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

theoretical evidence. Taking advantage of the analytical throat stress model, a fail-
ure mode transition behavior was systematically investigated [17]. As a result, it has
been suggested [17] that critical weld size would be independent of base plate thick-
ness as long as relative weld size s/t is considered, which is the major difference
from Gurney’s [3] and Noblett’s [4] findings. Then, a single curve-related critical
weld size to weld penetration is constructed.
Another factor that might affect failure mode transition behavior is joint mis-
alignment, including both axial and angular misalignment. Compared to weld pen-
etration, fewer studies focus on misalignment effects. Jakubczak and Glinka [18]
investigated effects of axial misalignments on failure mode transition of fillet con-
nections by performing fracture mechanics calculations. They reported that axial
misalignments had a much more significant influence on weld toe failure mode
than on weld root failure mode and, more importantly, that an increase in axial
misalignment seemed to promote failure mode transition from weld root cracking
to weld toe cracking for similar fillet weld sizes. The failure mode transition due to
axial misalignment was then observed in their experimental study [18] using a set
of cruciform specimens designed in the same weld size (s=t  1:0) but with varying
amounts of axial misalignments.
Recognizing the three influencing factors (i.e., weld size, weld penetration, and
misalignments) on fatigue failure mode transition, most studies [3,4,17,18] focused
more on theoretical study than on experimental. However, theoretical studies can-
not capture the complex shop floor practices, such as the status of weld penetration
and amount of misalignments. As such, using theoretically developed design refer-
ences may result in unfavorable weld size. Thus, a comprehensive experimental
study including shop floor practices is essential to help understand theoretical find-
ings and to facilitate weld design.
In this study, we start with a large amount of fatigue testing using load-carrying
cruciform joints designed in various relative weld sizes (s/t). Then, an analytical
weld throat stress model developed by Xing, Dong, and Threstha [17] and
misalignment-induced stress concentration factors developed by Xing and Dong
[19] for fatigue test specimens are introduced. Through a comprehensive stress
analysis, weld penetration and misalignment effects on weld toe and root cracking
modes are theoretically identified. Then, S-N-based data analysis is performed
along with a weld throat stress model and misalignment-induced stress concentra-
tion, and it is found that the weld toe cracking data and root cracking data follow
the same S-N mean curve. In addition to the findings noted earlier, failure mode
transition behavior is systematically investigated based on the ratio of fatigue lives
associated with weld toe cracking and weld root cracking modes, and an experi-
mentally based critical weld size is then obtained including the influence of weld
penetration and misalignments. Finally, a statistical analysis method is introduced
for establishing a confidence level with respect to the critical weld size beyond
which weld throat cracking would be deemed unlikely.
XING AND DONG, DOI 10.1520/STP159820160078 261

Fatigue Testing
TEST SPECIMENS
All fatigue test specimens are manufactured as cruciform load-carrying fillet-welded
specimens, as illustrated in Fig. 1, by adopting common shop-floor practices [16] so
that the test data reflect characteristic variability in a typical production environ-
ment. A test matrix summarizing specimen details is given in Table 1. Note that the
base plate and attachment plate are of the same thicknesses (5 mm and 10 mm).
Specimen width is 90 mm to ensure sufficient structural restraints to retain full-
strength weld residual stresses. To examine weld size effects on failure mode transi-
tion, different target (or design) weld fillet sizes (see Table 1) are specified on speci-
men design drawings that are presented to a shipyard production floor [17] that is
responsible for construction of all test specimens. Further specimen fabrication
details can be found in Huang et al. [16].
To facilitate data analysis, actual weld fillet sizes are measured by a laser scanning
device prior to fatigue testing. The detailed measurement procedure can be found in
Xing, Dong, and Threstha [17]. Note that all test results in this study will be presented
using measured fillet weld sizes. As reported in Xing, Dong, and Threstha [17], a cer-
tain amount of weld penetration has been observed in weld macros, even without
weld penetration being designed. As a result of this measurement [17], a lower bound
of the penetration is found at around p=t  0 (p denotes the penetration), and an up-
per bound is at about p=t  0:2. In addition, both axial and angular misalignments
are also measured, and the detailed measurements can be found in Ref. [19].

TEST PROCEDURE
All fatigue tests are performed using an MTS test machine with a load capacity 200
KIPs (890 KN) and are equipped with MTS 647 Hydraulic Wedge Grips. Fig. 2
shows a test specimen in a loaded configuration. In each group of specimens in

FIG. 1 Geometry of cruciform test specimen: (a) three-dimensional (3D) view; (b) 2D
cross-section.
262 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

TABLE 1 Cruciform fatigue specimen test matrix.

Base Plate Attachment Target Weld


Welding Process Weld Wire Thickness [mm]
Materials Materials Size s [mm]

DH-36 DH-36 FCAW 71T1-C 5 3, 5, 8


DH-36 DH-36 FCAW 71T1-C 10 5, 6, 8, 10, 12
HSLA-80 HSLA-80 FCAW 101T-C 5 3, 5, 8
HSLA-80 HSLA-80 FCAW 101T-C 10 5, 6, 8, 10, 12
DH-36 HSLA-80 FCAW MIL 71T-1C 5 3, 5, 8
DH-36 HSLA-80 FCAW MIL 71T-1C 10 5, 6, 8, 10, 12
DH-36 HSLA-80 SAW MIL 100S 5 3, 5, 8
DH-36 HSLA-80 SAW MIL 100S 10 5, 6, 8, 10, 12

Table 1, one-half of the specimens are tested with a remote load range of 30 Ksi (207
MPa) and the other half with a load range of 60 Ksi (414 MPa). A stress ratio of
R ¼ 1 is used for all specimens under load-controlled conditions. It should be
noted that finite element calculations under 30 Ksi (207 MPa) remote stress
under minimum possible root gap conditions (about 0.2 mm through weld macro-
based measurements) show that there is no contact occurring along the weld root
gap. A few pre-test trials indicate that a test frequency of 7 to 8 Hz can be used in
order to maintain load-controlled conditions while maximizing test efficiency.
During testing, both peak load range and displacement range are monitored and

FIG. 2 Fatigue test machine with a specimen mounted prior to testing.


XING AND DONG, DOI 10.1520/STP159820160078 263

recorded. In all test specimens, final failure is defined as when stiffness is reduced
by 50 % or complete separation, whichever occurs first.
In addition to documenting cycle to failure for each specimen according to the
final failure definition given in the previous section, cracking path and origin are
carefully examined and documented by separating test data into two failure catego-
ries: weld toe cracking (Mode A) and weld root cracking (Mode B). A representative
Mode A failure is illustrated in Fig. 3a, and Mode B failure is shown in Fig. 3b.

Analysis
WELD THROAT STRESS ANALYSIS
Analytical Weld Throat Stress Model
For a 2D stress problem, such as the cruciform specimen under tension loading,
along any hypothetical cut plane emanating from weld root at an angle h (see Fig. 4),
two traction stress components are exposed. These two components are referred to as
normal traction stress rs ðhÞ, in terms of its membrane part rm ðhÞ and bending part
rb ðhÞ, and transverse stress sT ðhÞ. The transverse shear stress component sT ðhÞ con-
tains only the membrane part, consistent with the transverse shear stress definition in
structural mechanics, and is referred to as sm ðhÞ hereafter.
To investigate weld root cracking behavior, an analytical weld throat stress
model was developed by Xing, Dong, and Threstha [17] corresponding to angular
section h (Fig. 4), and it is presented as follows:

rs ðhÞ ¼ rm ðhÞ þ rb ðhÞ (1)


rm  s  sin h þ sm  s  cos h
rm ðhÞ ¼ (2)
ah

FIG. 3 Representative failure modes observed from fatigue testing: (a) weld toe
cracking (Mode A) and (b) weld root cracking (Mode B).
264 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 4 Analytical fillet weld model: (a) fillet weld without penetration and (b) fillet weld
with penetration, p.

rb s2  3rm  s2 þ 3ah ðrm  s  sin h þ sm  s  cos hÞ


rb ðhÞ ¼ (3)
ah 2
rm  s  cos h  sm  s  sin h
sm ðhÞ ¼ (4)
ah

where rm , rb , and sm are membrane, bending, and transverse shear traction stress,
respectively, corresponding to h ¼ 90 , which is treated as known and obtained based
on the finite element calculation procedure (see [17]). The term ah can be expressed
as a function of h, as shown in Eq 5:

ah ¼ s=ðsin h þ cos hÞ: (5)

Note that the aforementioned expressions are also valid for load-carrying fillet weld
with a given amount of penetration, designated as p in Fig. 4b, by simply replacing
fillet leg size s with sþp in these equations.
To accommodate the presence of both normal and shear traction stresses, an
effective traction stress (ETS) definition proposed by Dong and Hong [20] and Wei
and Dong [21] is adopted here and, after inserting component stress expressions
given in Eqs 1–4, can be expressed as:
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
re ðhÞ ¼ rs 2 ðhÞ þ bsm 2 ðhÞ ¼
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u !2
u ðrm cos h  sm sin hÞðcos h þ sin hÞþ
u   (6)
u
u 6 r6b  r2m þ 12 rm cos cos hsm sin h
ðcos h þ sin hÞ2 ;
t hþsin h

þbðrm sin h þ sm cos hÞ2 ðcos h þ sin hÞ2


XING AND DONG, DOI 10.1520/STP159820160078 265

which has the same functional form as von Mises criterion by taking b ¼ 3. However,
it should be noted that because fatigue behavior of welded joints is governed by stress
range [21,22], all component stresses in Eq 6 should be replaced by respective compo-
nent ranges for fatigue data correlation purposes.
Taking advantage of the analytical ETS expression, the maximum value of ETS and
the corresponding failure angle, defined as critical plane at hc [17], can be readily evaluat-
ed by setting first derivative of Eq 6 with respect to angle h being zero, that is (Eq 7),

d
re ðhÞ ¼ 0: (7)
dh

For facilitating test data analysis, a series of relative weld sizes and weld pene-
trations are considered for evaluation of the maximum ETS, that is, re ðhc Þ. The
results are illustrated in Fig. 5a.
For comparison purposes, ETS at weld toe is also calculated by a finite element
analysis (FEA) procedure given in Xing, Dong, and Threstha [17]. As shown, effec-
tive traction stress associated to weld root cracking decreases more significantly
(Fig. 5a) than the stress responsible for weld toe cracking (Fig. 5b) as relative pene-
tration p/t increases. Especially in the weld toe cracking mode dominant region
[17], weld penetration shows a negligible effect on ETS at weld toe. As such, the
weld penetration effect can be neglected for weld toe cracking. In addition, it is
worth mentioning that ETS at weld toe is equivalent to normal stress rs due to the
negligible transverse shear stress sm under simple tension loading in this study.

Misalignment-Induced Stress Concentration Factors


For demonstrating misalignment effect on fatigue performance of fillet-welded lab-
oratory specimens, an analytical solution method for misalignment-induced stress
concentration factor had been proposed by Xing and Dong [19] by means of a
potential energy formulation considering end-gripping condition (see Fig. 6). As a

FIG. 5 Weld penetration effect on ETS: (a) weld throat ETS at hc ; (b) weld toe ETS.
266 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 6 Cruciform joint with axial misalignment: (a) dimensions and shape before
clamping and (b) deformed shape after clamping.

result, two misalignment-induced stress concentration factors (SCFs) are developed


for axial and angular misalignments, respectively [19]:
   
rb 3 ð3L3 þ 6L2 Lc þ 60LLc 2  40Lc 3 Þ e
ke ¼ ¼ (8)
rp e 10 L3 t
   
rb 4 ðL4  9L3 Lc þ 39L2 Lc 2  60Lc 3 L þ 30Lc 4 Þ
ka ¼ ¼ a: (9)
rp a 5 L3 t

If both axial and angular misalignments are present, the combined SCF can be cal-
culated as shown in Eq 10 [19]:

keþa ¼ ke þ ka (10)

which is the superposition of Eqs 8 and 9. It should be mentioned that angle a in


Eqs 9 and 10 has the unit of radian, although angular misalignment a is presented
in degrees in this study for purposes of easy visualization.
To relate test specimen geometry (including weld toe position) to dimensional
parameters used in the analytical model described in Fig. 6, Fig. 7 provides an illus-
tration using two actual fatigue test specimens. The resulting analytically calculated
SCFs for the two specimens are compared with finite element results in Fig. 8a
showing negligible differences. As illustrated in Fig. 8b, the misalignment-induced
SCFs at weld root are negligible for both 5-mm- and 10-mm-thick specimens,
XING AND DONG, DOI 10.1520/STP159820160078 267

FIG. 7 Illustration of dimensional relationships between actual test specimens and


analytical model (see Fig. 6): (a) a 5-mm-thick specimen and (b) a 10-mm-thick
specimen.

compared to SCFs at weld toe. As such, misalignment effect on weld root cracking
can be neglected.

S-N DATA ANALYSIS


Reference S-N Curve Generation
As briefly discussed earlier, the specimens involved in this study contain weld penetra-
tion and misalignments. As such, a reference S-N curve without influences from penetra-
tion and misalignments is needed for comparison purposes and for facilitating the
investigation on failure mode transition. Along this point, a set of fatigue data reported
by the Shipbuilding Research Association of Japan [22] is selected. The fatigue tests

FIG. 8 Misalignment-induced SCFs: (a) analytical SCF results versus FEA results and
(b) SCF at weld toe versus SCF at weld root evaluated by FEA.
268 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

reported [22] were carried out using a non-load-carrying cruciform joint, which ensures
no influences from penetration and axial misalignment. The base plate thicknesses vary
from 10 mm to 80 mm, and adopted load ratios vary from 0.06 to 0.2.
Fig. 9a shows all test data reported in the study [22] in terms of conventional
nominal stress methods. It can be clearly seen that the thinner plate data tend to be
situated around the upper bound, and the thicker ones are situated on the lower
side. By introducing the master S-N curve approach, originally developed by Dong,
Prager, and Osage [23] and Dong [24] and adopted by ASME 2007 Div 2 Code [25]
and API 579 [26], the same set of data illustrated in Fig. 9a is presented in the form
of an equivalent traction stress range, that is,

Drs
DSs ¼ 2m 1 (11)
te 2m IðrÞm

versus cycle to failure (N) in Fig. 9b. In Eq 11, Drs represents the structural stress
range calculated using nodal force-based procedure: te ¼ t=2 for symmetric cruci-
form joints; m ¼ 3.6, and r is expressed as r ¼ jDrb j=jDrs j, where Drb ¼ Drs  Drm
and Drm is the membrane part of Drs . For the detailed derivation of Eq 11 and its
application in formulating the master S-N curve method, interested readers may
consult a series of prior publications [23,24,27,28]. In addition, the Welding Re-
search Council Bulletin No. 523 [27] provides a comprehensive documentation on
the method and its detailed validations. Compared to Fig. 9a, the data in Fig. 9b
are situated in a much narrower scatter band with a standard deviation of 0.147—
versus 0.205 in Fig. 9a. The thickness effect is no longer observed in Fig. 9b. Note
that the S-N curve presented in Fig. 9b hereafter will be treated as a reference for
analysis of the fatigue data tested in this study.

Weld Root Cracking Data


Fig. 10a shows the weld root cracking data tested in this study in terms of nominal
stress range. The scatter band is quite big even though only one joint type is

FIG. 9: Reported fatigue data [22]: (a) correlated using nominal stress and (b)
correlated by equivalent stress in Eq 11.
XING AND DONG, DOI 10.1520/STP159820160078 269

involved. Recently, it had been proven by Hong [8] that the master S-N curve meth-
od also is effective for correlating weld root cracking data. Then, the same set of
data shown in Fig. 10a is analyzed using an equivalent stress parameter (Eq 11) by
simply replacing Drs with re ðhc Þ. This results in a much narrower standard devia-
tion of 0.202 (Fig. 10b) versus 0.342 in Fig. 10a. In addition, one may notice that the
weld root cracking data tend to be situated around the upper bound of the reference
S-N curve developed in the previous section based on non-load-carrying test speci-
mens. This should contribute to weld penetration, drastically reducing the weld
throat stress (see Fig. 5). In other words, weld penetration will significantly improve
the fatigue performance of load-carrying fillet connections. Based on penetration
measurements [17], it seems reasonable to set weld penetration as p/t ¼ 0.1 on all of
the data for investigation of the penetration effect. Then, the root cracking data are
reanalyzed by using the weld throat stress given in Fig. 5 corresponding to p/t ¼ 0.1.
As a result, all the data move closer to the reference mean curve (see Fig. 11). This
suggests that weld root cracking mode essentially follows the same mean curve as
the reference fatigue data, even though the scatter band is greater due to variability
induced during the welding process.

Weld Toe Cracking Data


It should be mentioned that the fatigue tests in this study were carried out at three
independent laboratories [17]. The measured misalignments are only available for
part of the fatigue tests. Fig. 12a shows the weld toe cracking data without consider-
ing misalignment. The standard deviation is 0.274. Compared to weld root cracking
data (Fig. 10b and Fig. 11), no misalignment effect is observed. This confirms earlier
findings in stress analysis. When considering misalignment-induced bending stress
by replacing Drs in Eq 11 with Drs þ keþa Drn (rn is nominal stress), the scatter
band is significantly improved to 0.196 (see Fig. 12b). Nearly all the data points are sit-
uated within the scatter band of the reference S-N curve. This suggests both weld

FIG. 10 Analysis of weld root cracking data tested in this study: (a) in terms of nominal
stress and (b) in terms of equivalent stress (Eq 11).
270 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 11 Reanalyzed weld root cracking data considering weld penetration.

root cracking and weld toe cracking data follow the same mean curve (i.e., the refer-
ence mean because they are not influenced by weld penetration and misalignments).

FATIGUE FAILURE MODE TRANSITION ANALYSIS


Fatigue Life-Based Critical Weld Size
With S-N-based data analysis performed, it is time to begin investigating the failure
mode (from root cracking to toe cracking) transition behavior of the experimental
data. Denoting fatigue lives corresponding to weld toe cracking mode (Mode A)

FIG. 12 Part of weld toe cracking data with misalignments measured: (a) without using
misalignment-induced SCF and (b) misalignment-induced SCF.
XING AND DONG, DOI 10.1520/STP159820160078 271

and weld toe cracking mode (Mode B) as N A and N B , respectively, the critical weld
size can be obtained at N B ¼ N A . Instead of solving a complex exponential equa-
tion (S-N curves), an alternative method is to explicitly seek the intersection point
of the two curves:

NB
1 (12)
NA
NA
 1: (13)
NB

As noted earlier, it has been proven that both weld toe cracking and root cracking
modes follow the same S-N curve (i.e., the reference S-N curve) without involving
weld penetration and misalignment. Then, an idealized (specimen conditions) criti-
cal weld size can be evaluated using reference S-N curves along with Eqs 12 and 13.
This results in a critical weld size at s=t  1:16 as illustrated in Fig. 13, which repre-
sents a transition point for idealized specimen conditions. It worth mentioning that
the idealized critical weld size is equivalent to the EETS-based theoretical critical
weld size proposed by Xing, Dong, and Threstha [17]. As seen in Fig. 13, an actual
transition point, s=t  0:76, is evaluated using S-N curves obtained from actual
test data without weld penetration and misalignment corrections. It is much low-
er than that idealized (or theoretical) critical weld size s=t  1:16.

FIG. 13 Comparison among critical weld sizes obtained by considering different


specimen conditions.
272 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

It has been found in earlier sections of this paper and further confirmed by S-N
data analysis, that weld penetration has only minor influence on weld toe cracking,
and misalignments have negligible effects on weld root cracking. Along this point,
the critical weld size influenced by misalignments can be evaluated by using the ref-
erence S-N curve for weld root cracking (assuming without weld penetration) and
the S-N curve from Fig. 12b (weld toe cracking data containing misalignments). As a
result, the critical weld size is at s=t  0:98 (see Fig. 13). In a similar manner, the
critical weld size as weld penetrations exist can be obtained by assuming a reference
S-N curve for weld toe cracking (assuming without misalignments) and the actual
weld root cracking S-N curve obtained from Fig. 10b. This results in a critical weld
size at s=t  0:94 (see Fig. 13).
Based on all the results illustrated in Fig. 13, both weld penetration and mis-
alignments tend to lower the failure mode transition point. This can be interpreted
as: (a) weld penetration increases fatigue lives associated with weld root cracking,
although without much influence on weld toe cracking; and (b) misalignments de-
crease the fatigue lives associated with weld toe cracking without having much in-
fluence on weld root cracking. Both of them can make fatigue failure modes
transition at lower relative weld size (s/t) comparing to idealized joint conditions.
For visualizing the transition points along with the test data, all fatigue data
scaled by a factor f, defined by Xing, Dong, and Threstha [17] to collapse differ-
ent loading levels and base plate thicknesses, are plotted in terms of relative weld
size s/t (see Fig. 14). The idealized critical weld size s=t  1:16 is definitely conser-
vative, over which no test failed at weld root because every specimen in this study
contains a certain amount of weld penetration and misalignments. The actual
critical weld size s=t  0:76 seems like a reasonable one, which divides the whole
data set into a weld root cracking dominant region (on the left) and a weld toe
cracking dominant region (on the right). As seen in Fig. 14, using an idealized
critical size or a theoretical one [17], s=t  1:16 as the design reference will result

FIG. 14 Scaled fatigue lives versus relative weld size.


XING AND DONG, DOI 10.1520/STP159820160078 273

in oversized weld, which may cause severe welding-induced distortion [17]. How-
ever, using the actual transition point s=t  0:76 can only give a 50 % confidence
level to prevent weld root cracking. Such a dilemma will be dealt with in the next
section.

Logistic Regression
It is statistically possible for each load-carrying fillet-welded connection to develop
a fatigue crack either at weld toe or at weld root due to inherent variabilities such as
penetration status, misalignments, weld quality, and so on. The possibility of each
cracking mode is highly related to relative weld size s/t as seen in Fig. 14. Because
there are two failure modes, either weld throat or weld toe cracking, a logistic re-
gression analysis method for analyzing problems with binary outcome [29] should
offer another perspective on the interpretation of the failure mode transition behav-
ior seen in Fig. 14.
For the present application, all data in Fig. 14 are considered as independent
observations of variable pairs (xi, yi), i ¼ 1, 2, …, n, where yi denotes a binary out-
come variable, and xi is the independent variable representing relative weld size s/t,
corresponding to ith observation in a given data population. The binary outcome
variable yi takes a value of either 0 or 1, representing weld root cracking or weld toe
cracking mode, respectively. Then, a logistic regression model can be constructed in
the following form [30]:
 !
p ts s
ln   ¼ b0 þ b  (14)
1  p ts t
 
where, p st and 1  p ts represent the probabilities of weld toe cracking and weld
root cracking, respectively, as a function of an independent variable (i.e., relative weld
size s/t). Parameters b0 and b are to be determined through a curve-fitting process to
be discussed next. From Eq 14, the probability of weld toe failure can be expressed
in Eq 15:
 s 1
p ¼ s : (15)
t 1 þ eðb0 þbtÞ

The corresponding conditional likelihood function [30] of the whole data set
shown in Fig. 15 can be constructed as the product of probability of each test out-

come corresponding to either weld toe cracking—that is, p st , or to weld root
s
cracking—that is, 1  p t :
Yn  s  yi  s1yi
Lðb0 ; bÞ ¼ p 1p : (16)
i¼1
t t
274 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 15 Probabilities of weld toe cracking and root cracking versus weld size.

By introducing the maximum likelihood method [30], that is, finding coefficients
b0 and b that maximizes the conditional likelihood function in Eq 16, the two
coefficients b0 and b were determined as b0 ¼ 14:436 and b ¼ 18:573. Then, the
probability of weld toe failure can be estimated using Eq 15 for any relative weld size
(s/t), as graphically presented in Fig. 15. The corresponding weld throat failure probabili-

ty 1  p st is also presented in Fig. 15 for illustration purposes.
The two vertical dashed lines (in Fig. 15) show both 50 % and 95 % confidence
levels for developing weld toe failure when relative weld size exceeds s=t  0:78 and
s=t  0:93. Note that s=t  0:78 at 50 % confidence level agrees well with the actual
critical weld sizes s=t  0:76 obtained earlier. A critical weld size of s=t  0:93 at
95 % confidence level from logistic analysis results (Fig. 15) seems conservative
enough for design consideration. A more detailed discussion of the implications of
the logistic regression results on weld sizing in fatigue design is given in [31].

Conclusions
Based on the large number of fatigue tests performed in this study and on a com-
prehensive data analysis, the following conclusions can be drawn:
1. Weld root cracking data and weld toe cracking data have been shown to follow
the same mean S-N curve, differing in scatter band.
2. Weld penetration has significant effects on weld root cracking and must be
taken into account in the development of failure mode transition criteria. For
this purpose, the proposed analytical throat stress model proves to be effective
in capturing weld penetration effects on weld throat stress in a closed-form
solution.
XING AND DONG, DOI 10.1520/STP159820160078 275

3. Joint misalignments have been shown to have the most significant effects on
weld toe cracking. The analytical solution developed for modeling joint
misalignment provides an effective means for taking misalignment effects into
account in weld toe failure mode S-N based data interpretation.
4. Assuming there is no weld penetration and misalignment, a theoretical fillet
weld sizing beyond which weld toe cracking dominates is found to be
s=t  1:16. Considering actual test data containing penetration and misalign-
ments resulting from shop-floor practices, a fillet weld sizing is found to be
s=t  0:76.
5. Test data obtained in this study (after performing logistic regression) is found
to be s=t  0:78 with respect to 50 % confidence, which is consistent with the
fatigue life-based critical weld size s=t  0:76. The critical weld size at this
level should be recommended as a lower limit of sizing criterion to ensure the
target weld size is located in weld toe fatigue cracking dominant region. In
practice, if a 95 % confidence level is considered, preventing weld root
cracking from happening is found to be s=t  0:93. It is important to note
that the sizing criteria are developed for lightweight ship structures or any oth-
er lightweight structure manufactured under similar conditions.

ACKNOWLEDGMENTS
The authors acknowledge the support of this work through a grant from the National
Research Foundation of Korea (NRF) and funding from the Korean government
(MEST) through GCRC-SOP at the University of Michigan under Project 2-1:
Reliability and Strength Assessment of Core Parts and Material System.

References

[1] Maddox, S. J., Fatigue Strength of Welded Structures, Woodhead, Cambridge, UK, 1991.

[2] Maddox, S. J., “Status Review on Fatigue Performance of Fillet Welds,” J. Offshore Mech.
Arct., Vol. 130, No. 3, 2008, pp. 537–550.

[3] Gurney, T. R., Fatigue of Welded Structures, CUP Archive, Cambridge, UK, 1979.

[4] Noblett, J. E. and Andrews, R. M., A Stress Intensity Factor Solution for Root Defects in
Fillet and Partial Penetration Welds, TWI, Cambridge, UK, 2000, p. 120.

[5] Kainuma, S. and Mori, T., “A Fatigue Strength Evaluation Method for Load-Carrying Fillet
Welded Cruciform Joints,” Int. J. Fatigue, Vol. 28, No. 8, 2006, pp. 864–872.

[6] Kainuma, S. and Mori, T., “A Study on Fatigue Crack Initiation Point of Load-Carrying
Fillet Welded Cruciform Joints,” Int. J. Fatigue, Vol. 30, No. 9, 2008, pp. 1669–1677.

[7] Balasubramanian, V. and Guha, B., “Establishing Criteria for Root and Toe Cracking of
Load Carrying Cruciform Joints of Pressure Vessel Grade Steel,” Eng. Fail. Anal., Vol. 11,
No. 6, 2004, pp. 967–974.
276 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

[8] Hong, J. K., “Evaluation of Weld Root Failure Using Battelle Structural Stress Method,”
J. Offshore Mech. Arct., Vol. 135, No. 2, 2013, pp. 73–80.

[9] Hobbacher, A., ed., Fatigue Design of Welded Joints and Components: Recommenda-
tions of IIW Joint Working Group XIII–XV, Abington Publishing, Abington, Cambridge,
UK, 1996.

[10] Hobbacher, A., Recommendations for Fatigue Design of Welded Joints and Components,
Welding Research Council Bulletin 520, The Welding Research Council, Inc., Shaker
Heights, OH, 2009.

[11] BS 7608, Guide Fatigue Design and Assessment of Steel Structures, British Standards
Institution, London, 1993.

[12] Design of Steel Structures—Part 1-1. ENV 1993-1-1, Eurocode 3, European Committee for
Standardization, Brussels, 1992.

[13] Dong, P. and Hong, J. K., “Analysis of Hot Spot Stress and Alternative Structural
Stress Methods,” presented at the 22nd International Conference on Offshore Me-
chanics and Arctic Engineering, Cancun, Mexico, June 8–13, 2001, ASME, New York,
2003, pp. 213–224.

[14] Dong, P., “A Structural Stress Definition and Numerical Implementation for Fatigue
Analysis of Welded Joints,” Int. J. Fatigue, Vol. 23, No. 1, 2001, pp. 865–876.

[15] Nie, C. and Dong, P., “A Traction Stress Based Shear Strength Definition for Fillet Welds,”
J. Strain Anal. Eng., Vol. 47, No. 8, 2012, pp. 562–575.

[16] Huang, T. D., Harbison, M., Kvidahl, L., Niolet, D., Walks, J., Stefanick, K., Phillippi, M.,
Dong, P., DeCan, L., Caccese, V., Blomquist, P., Kihl, D., Wong, R., Nappi, N., Gardner, J.,
Wong, C., Bjornson, M., and Manuel, A., “Reduction of Overwelding and Distortion for
Naval Surface Combatants, Part 1: Optimized Weld Sizing for Lightweight Ship
Structures,” J. Ship Prod Des., Vol. 30, No. 4, 2014, pp. 184–193.

[17] Xing, S., Dong, P., and Threstha, A., “Analysis of Fatigue Failure Mode Transition in Load-
Carrying Fillet-Welded Connections,” Mar. Struct., Vol. 46, 2016, pp. 102–126.

[18] Jakubczak, H. and Glinka, G., “Fatigue Analysis of Manufacturing Defects in Weldments,”
Int. J. Fatigue, Vol. 8, No. 2, 1986, pp. 51–57.

[19] Xing, S. and Dong, P., “An Analytical SCF Solution Method for Joint Misalignments and
Application in Fatigue Test Data Interpretation,” Mar. Struct., Vol. 50, 2016, pp. 143–161.

[20] Dong, P. and Hong, J. K., “A Robust Structural Stress Parameter for Evaluation of
Multiaxial Fatigue of Weldments,” J. ASTM Int., Vol. 3, No. 7, 2006, pp. 1–17, https://
doi.org/10.1520/JAI100348

[21] Wei, Z. and Dong, P., “A Generalized Cycle Counting Criterion for Arbitrary Multi-Axial
Fatigue Loading Conditions,” J. Strain Anal. Eng., Vol. 49, No. 5, 2014, pp. 325–341.

[22] Fatigue Design and Quality Control for Offshore Structures, SR202, Shipbuilding
Research Association of Japan, 1991.

[23] Dong, P., Prager, M., and Osage, D., “The Design Master SN Curve in ASME DIV 2 Rewrite
and Its Validations,” Weld. World, Vol. 51, Nos. 5–6, 2007, pp. 53–63.
XING AND DONG, DOI 10.1520/STP159820160078 277

[24] Dong, P., “A Robust Structural Stress Method for Fatigue Analysis of Offshore/Marine
Structures,” J. Offshore Mech. Arct., Vol. 127, No. 1, 2005, pp. 68–74.

[25] ASME Boiler and Pressure Vessel Code, Section VIII, Division 2, ASME, New York, 2007.
[26] API 579 RP-1/ASME FFS-1, Fitness-For-Service, American Petroleum Institute, Washington,
DC, 2007.

[27] Dong, P., Hong, J. K., Osage, D. A., Dewees, D., and Prager, M., The Master S-N Curve
Method: An Implementation for Fatigue Evaluation of Welded Components in the ASME
B&PV Code, Section VIII, Division 2 and API 579-1/ASME FFS-1, Welding Research
Council Bulletin 523, The Welding Research Council, Inc., Shaker Heights, OH, 2010.

[28] Dong, P., Pei, X., Xing, S., and Kim, M. H., “A Structural Strain Method for Low-Cycle
Fatigue Evaluation of Welded Components,” Int. J. Press. Ves. Pip., Vol. 119, 2014,
pp. 39–51.

[29] Mesa, J. L., “Understanding Data in Clinical Research: A Simple Graphical Display for
Plotting Data (Up to Four Independent Variables) after Binary Logistic Regression
Analysis,” Med. Hypotheses, Vol. 62, No. 2, 2004, pp. 228–232.

[30] Hosmer, D. W. and Lemeshow, S., Applied Logistic Regression, Wiley, New York, 2004.

[31] Xing, S., Dong, P., and Wang, P. A., “Quantitative Weld Sizing Criterion for Fatigue
Design of Load-Carrying Fillet-Welded Connections,” January, 2017, International Jour-
nal of Fatigue, http://dx.doi.org/10.1016/j.ijfatigue.2017.01.003
278 FATIGUE AND FRACTURE TEST PLANNING, TEST DATA ACQUISITIONS AND ANALYSIS

STP 1598, 2017 / available online at www.astm.org / doi: 10.1520/STP159820160079

Jifa Mei1 and Pingsha Dong1

Analysis of Nonproportional
Multiaxial Fatigue Test Data of
Various Aluminum Alloys Using a
New Damage Parameter
Citation
Mei, J. and Dong, P., “Analysis of Nonproportional Multiaxial Fatigue Test Data of Various
Aluminum Alloys Using a New Damage Parameter,” Fatigue and Fracture Test Planning, Test
Data Acquisitions and Analysis, ASTM STP1598, Z. Wei, K. Nikbin, P. C. McKeighan, and
D. G. Harlow, Eds., ASTM International, West Conshohocken, PA, 2017, pp. 278–298, http://
dx.doi.org/10.1520/STP1598201600792

ABSTRACT
This paper presents a detailed investigation into nonproportional loading-
induced multi–axial fatigue damage in a series of aluminum alloys using a
recently developed moment of load path (MLP) model that contains a material
sensitivity parameter. A generalized procedure is presented for determining
material sensitivity to nonproportional load path and is demonstrated for
characterizing multi–axial fatigue damage in these aluminum alloys under
nonproportional loading conditions. Major findings are: (a) All of the alloys
studied in this paper show a lesser degree of sensitivity to nonproportional
loading when comparing structural steels recently reported by the same authors
with material sensitivity parameter (a defined with respect to stress plane or ae
as to strain plane) ranging from 0.35 to 0.5, which is much smaller than that of
structural steels that were found to be around unity (i.e., a  ae  1); (b) within
the same aluminum alloy type, it is found that material sensitivity parameter (a)
calculated on stress plane using stress–life test data is close to the values
calculated on strain plane using strain–life data (i.e., ae ); and (c) with the

Manuscript received March 22, 2016; accepted for publication October 5, 2016.
1
University of Michigan, Dept. of Naval Architecture and Marine Engineering, 2600 Draper Dr., Ann Arbor, MI 48109
2
ASTM Symposium on Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis on May 4–5,
2016 in Grand Hyatt, San Antonio, TX.

Copyright V
C 2017 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.
MEI AND DONG, DOI 10.1520/STP159820160079 279

material sensitivity parameter (a or ae ) being determined, MLP–based


pffiffiffiffiffi equivalent
pffiffiffi
stress (defined on r  bs plane) or equivalent strain (on e  be c) are shown
to be effective in correlating a large amount of multi–axial fatigue test data
subjected to a varying degree of load–path nonproportionality.

Keywords
multi-axial fatigue, aluminum alloys, nonproportional loading, fatigue damage
modeling, damage parameter, moment of load path, material sensitivity

Introduction
Multi-axial fatigue cases can be grouped into those that are proportional and non-
proportional, depending on whether or not the principal stress directions rotate dur-
ing cyclic fatigue loading [1,2]. For proportional multi-axial fatigue, the effective
stress range defined by following conventional yield criteria (i.e., von Mises yield cri-
teria) in terms of ranges of each stress component is generally applicable. However,
nonproportional multi-axial loading is much more complex than proportional load-
ing and unconservative fatigue life prediction will occur if nonproportionality is not
taken into account properly [3–8].
Over the past few decades, there has been a great deal of experimental evidence
in the literature showing that load-path nonproportionality can introduce addition-
al fatigue damage under multi-axial cyclic loading conditions in a large class of
structural materials such as structural steels [3–8], various types of aluminum alloys
[9–15], titanium alloys [16–19], and magnesium alloys [20,21]. Some typical inves-
tigations include systematic nonproportional fatigue tests following different load
paths carried out by Itoh et al. [7] on stainless steel (SS 304) and aluminum alloys
[12]. In order to take into account a nonproportional hardening effect of materials,
which is believed to be the root cause of additional fatigue damage under nonpro-
portional loading, Itoh et al. [7] introduced a material-hardening parameter in their
proposed equivalent strain range as a fatigue damage parameter. However, it should
be noted that additional hardening of materials during nonproportional loading is
not the only cause of nonproportional damage because reduced fatigue life is also
observed for metal alloys [22,23] without exhibiting nonproportional path-induced
hardening behavior. Considering the rotation of maximum shear-stress plane, Son-
sino and Kueppers [6] showed that nonproportionality-induced fatigue damage can
be captured by an integral formulation of shear stress over all planes, referred to as
an effective equivalent stress hypothesis (EESH). The drawback of the proposed
parameter is that it can only be applied to constant amplitude sinusoidal loading
with a known phase angle between two stress components. There are also several
phenomenological models [24,25], such as minimum circumscribed circle (MCC)
and minimum circumscribed ellipse (MCE), proposed so far that are trying to cap-
ture the nonproportional damage effect without offering any mechanics-based
interpretation. For nonproportional fatigue of welded joints, Dong and Hong [26]
280 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

proposed a modified Gough’s ellipse model in which a fatigue damage parameter is


analytically formulated as a function of phase angle (d) shift between normal and
shear-stress histories, if both stress histories can be expressed as synchronous sinu-
soidal wave forms. A more compressive literature review of nonproportional fatigue
and related multi-axial cycle counting methods can be found in Mei and Dong [27].
When it comes to nonproportional multi-axial fatigue of aluminum alloys,
which is the focus of this paper, nonproportional fatigue damage is also observed in
different aluminum alloys. However, they tend to exhibit less sensitivity to load-
path nonproportionality when compared with structural steels. This category of
data includes those obtained by Kueppers and Sonsino [28,29] and Sonsino [30] on
aluminum weld joints as well as those collected on non-welded tubular specimens
by various researchers [9–15]. However, a comprehensive analysis of various types
of aluminum alloys in a consistent manner is still lacking.
Based on our interpretation of nonproportional multi-axial fatigue [27], an
effective fatigue damage parameter in conjunction with a consistent fatigue defini-
tion is essential in order to effectively characterize how load-path nonproportion-
ality impacts multi-axial fatigue behavior. Unfortunately, this has been an
ongoing research area without a clear consensus, as recently discussed by Dong,
Wei, and Hong [31], Wei and Dong [32,33], and Mei and Dong [27], particularly
when dealing with complex load paths or variable amplitude multi-axial loading
histories in which stress components vary independently with respect to one and
another. In recognizing such a need, the authors have recently proposed a
“moment of load path” (MLP) multi-axial fatigue damage model for which a fa-
tigue cycle is defined by the path-dependent maximum range (PDMR) cycle
counting method proposed by Dong, Wei, and Hong [31] and Wei and Dong
[32,33]. Recent applications [27] of this MLP model in analyzing a large amount
of test data of structural steels available in the literature have shown its promise in
effectively correlating test data in terms of both stress and strain. The work [27]
along this line has demonstrated that load-path proportionality-induced fatigue
damage can be effectively captured by the MLP model. However, questions re-
main on how such a model can be used to model multi-axial fatigue damage in
materials that have been shown to be less sensitive to or even possibly to benefit
from the presence of load-path proportionality.
The purpose of this paper is to address how the MLP model can be effectively
used for characterizing effects of load-path proportionality on multi-axial fatigue
damage in a series of aluminum alloys for which well-documented test data are
available. These include 2000, 6000, and 7000 series of aluminum alloys tested
under simple proportional and nonproportional load paths. The emphasis will be
on how material sensitivity parameters (available in MLP model) to load-path non-
proportionality can be consistently determined for each aluminum alloy type and
applied for correlating test data generated using more complex load paths or time
histories.
MEI AND DONG, DOI 10.1520/STP159820160079 281

Multi-Axial Fatigue Damage Parameter


MLP-BASED MODEL
As presented in detail by Mei and Dong [27], consider a nonproportional load path
pffiffiffi
f on r  bs plane illustrated in Fig. 1, which can be shown to
from A to B, that is, AB,
constitute one-half cycle, according to the PDMR cycle-counting procedure [31–33].
pffiffiffi
Note that b represents a fatigue equivalence parameter between S-N curves
obtained by performing pure normal stress and pure shear-stress cyclic testing (see
more detailed discussions by Dong, Wei, and Hong [31] and Wei and Dong [32,33]
in the context of the PDMR cycle-counting method). It has been shown by a large
amount of data in recent work [27] that b can be taken as 3, corresponding to the fact
pffiffiffi
that uniaxial normal stress r and pure torsion stress 3s are equivalent in terms of
von Mises stress definition. A multi-axial fatigue damage parameter D for any given
nonproportional load path AB f can be decomposed into two parts, as shown in Eq 1:

D ¼ DP þ DNP (1)

in which DP represents damage caused by the reference proportional loading event


from A to B (i.e., AB), which can be directly related to distance from A to B, or the

f nonproportional circular load path AB


FIG. 1 Nonproportional load path AB, c and
reference (proportional) load path AB.
282 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

effective stress range Dre , and DNP represents load-path nonproportionality caused
damage due to any excursions of load path AB f deviating from the reference propor-
tional load path (AB). Therefore, one possible way of representing the load-path non-
f can be stated as follows, with respect to
proportionality related damage along Path AB
the local x0  y0 coordinate system:
ð
DNP ¼ r 0  jsinðhÞjds0 : (2)
e
AB

It can then be shown [27] that a dimensionless form of load-path nonpropor-


tionality-induced damage parameter can be expressed as follows:
Ð 0 Ð 0
r jsinðhÞjds0 r jsinðhÞjds0
DNP e e
gNP ¼ ¼ AB
Ð ¼ AB (3)
DMax RjsinðhÞjds0 2R2
b
AB

where DMax represents the maximum possible nonproportional fatigue damage caused
by the half-circular load path represented by dashed lines in Fig. 1. Then, gNP in Eq 3
can be referred to as a normalized load-path nonproportionality damage factor with
respect to the maximum possible damage DMax , noting that gNP varies from zero
(corresponding to the proportional load path AB) to unity (corresponding to the half-
c Then, an equivalent stress, taking into account nonpropor-
circular load path AB).
tional load path-induced damage, can be written as:

DrNP ¼ Dre ð1 þ a  gNP Þ: (4)

Note that a material sensitivity parameter a is inserted in Eq 4 to accommodate


the fact that some materials are more sensitive to nonproportional multi-axial fa-
tigue loading than others, as has been observed [28–30].
Eq 4 can also be rephrased with respect to strain
pffiffiffiffiffi space for interpretation of
strain-life-based multi-axial test data (e.g., e  be c plane, in which be is a
fatigue equivalency parameter determined by comparing two e  N curves gener-
ated using pure normal strain e and pure shear strain c cyclic tests, respectively).
Consistent with b ¼ 3 in stress space, be ¼ 1=3 has been shown to be valid for
multi-axial strain-life-based test data correlation. Then, an equivalent effective
strain parameter incorporating nonproportional load-path-induced damage can be
written as:
 
DeNP ¼ Dee 1 þ ae  gNP
e
: (5)

In Eq 5, Dee is effective
pffiffiffiffistrain
ffi range, which can be directly related to distance
e
between two points on  b c plane; ae is a material-dependent, nonproportional-
e
ity sensitivity parameter defined in terms ofpcyclic
ffiffiffiffieffi strain fatigue test results; and gNP
can be evaluated according to Eq 3 on e  b c plane.
MEI AND DONG, DOI 10.1520/STP159820160079 283

ANALYTICAL TREATMENT OF ELLIPTICAL LOAD PATHS


In performing lab multi-axial fatigue tests, it is common that both normal and shear
components are represented in a sinusoidal wave form with a phase shift angle (d),
for example,

r ¼ r0 sinðhÞ
(6)
s ¼ s0 sinðh  dÞ
pffiffiffi
which form an elliptical load path in r  bs plane (see Fig. 2). The load-path pro-
portionality fatigue damage parameter gNP given in Eq 3 be expressed as a closed form
solution as described as follows.
pffiffiffi
Consider a nonproportional load case, that is, d 6¼ 0. Let r0 ¼ 1; s0 ¼ 1= 3,
pffiffiffi
and b ¼ 3. The resulting load paths according to Eq 6 in r  bs plane become

elliptical load paths, as shown in Fig. 2 for phase shift of d ¼ 60 as an illustration.
It can be shown that the major semi-axis length A (see Eq 7) and minor semi-
axis length B (see Eq 8) are determined as a function of r0 ; s0 ; and d:
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u 2br20 s20 sin2 d
u
A¼t qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi (7)
2
ðr20 þ bs20 Þ  ðr20 þ bs20 Þ  4br20 s20 sin2 d
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u 2br20 s20 sin2 d
u
B¼t qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi : (8)
2
ðr20 þ bs20 Þ þ ðr20 þ bs20 Þ  4br20 s20 sin2 d

FIG. 2 Illustration of elliptical load path with the long axis and short axis of ellipse
located on local coordinate system x0  y0 .
284 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

Then, the elliptical load path, as illustrate in Fig. 2, can be described with respect to a
local coordinate x0 y0 , shown in Fig. 2, as:
 0 2  0  2
x y
þ ¼1 (9)
A B

with
x0 ¼ A cosðhÞ
(10)
y0 ¼ B sinðhÞ

where h in Eq 10 is a polar coordinate with respect to x0 axis. Then, the integral form
of nonproportional damage parameter DNP according to Eq 2 becomes:

ðp ðp ðp qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
DNP ¼ r sinðhÞds ¼ yds ¼ B sinðhÞ A2 sin2 ðhÞ þ B2 cos2 ðhÞ dh: (11)
0 0 0

Integrating Eq 11 leads to:


ð 1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffi 
2 2 2
arcsinðeÞ
DNP ¼ 2AB 1  e x dx ¼ AB 1e þ (12)
0 e
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
where e ¼ ðA2  B2 Þ=A2 is the eccentricity of ellipse. Substituting Eq 12 into Eq 3
results in:
pffiffiffiffiffiffiffiffiffiffiffiffiffi
g arcsinð 1  g2 Þ
gNP ðr0 ; s0 ; dÞ ¼ ðg þ pffiffiffiffiffiffiffiffiffiffiffiffiffi Þ (13)
2 1  g2

where
0 10:5
 2 2
 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2 2 2 2 2 2
B B r0 þ bs0  ðr0 þ bs0 Þ  4br0 s0 sin dC
g¼ ¼@ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi A : (14)
A 2
ðr20 þ bs20 Þ þ ðr20 þ bs20 Þ  4br20 s20 sin2 d

It is important to note that the closed form solution given in Eq 13 makes the
determination of material sensitivity parameter a rather convenient once fatigue
test data under proportional loading (represented by Dre ) and nonproportional
loading (represented by DrNP ) become available. Applications of Eq 14 will be dis-
cussed in the next section.

PROCEDURE FOR DETERMINING a OR ae


As shown in Fig. 3, one only needs to conduct two sets of relatively simple multi-
axial fatigue tests: in-phase (proportional) tests labeled as (P) and out of phase (e.g.,
with a phase shift angle of 90 ) tests labeled as (Q), with the two respective load
paths illustrated in Fig. 3a. Note that the two load cases considered in Fig. 3a do not
ðPÞ ðQÞ
require that they are of the same effective stress range, that is, Dre ¼ Dre is not
required. When their corresponding S-N test data are plotted as illustrated in
MEI AND DONG, DOI 10.1520/STP159820160079 285

FIG. 3 Procedure for determining material sensitivity parameter a: (a) proportional and
nonproportional circle load paths for fatigue testing; and (b) calculation of
material sensitivity parameter a with respect to reference fatigue life Nref .

Fig. 3b in terms of effective stress range (Dre ), which is defined as the maximum dis-
pffiffiffiffiffi
tance traversed in r  bs plane shown in Fig. 3a, two separate S-N curves can then be
determined by curve fitting. Now consider a reference fatigue life (Nref ), the MLP-based

equivalent stress range for both in-phase case (A) and 90 out-of-phase loading case
ð AÞ ðBÞ
(B) should be of the same values, each represented by DrNP and DrNP ; respectively.
ð AÞ ð AÞ
For the in-phase case (A), because gNP ¼ 0, DrNP can be written, according to Eq 4, as:
 
ðAÞ ð AÞ
DrNP ¼ DrðeAÞ 1 þ a  gNP ¼ DrðeAÞ : (15)

ðBÞ
For the out-of-phase case (B) for which gNP ¼ 1 by definition (see Eq 3 as
applied to the circular load path shown in Fig. 3a),
 
ðB Þ ðB Þ
DrNP ¼ DrðeBÞ 1 þ a  gNP ¼ DrðeBÞ ð1 þ aÞ: (16)

Because a is introduced in Eq 4 as a material sensitivity parameter that equates


nonproportional loading-induced effective stress to that corresponding to propor-
ð AÞ ðBÞ
tional loading, it follows that DrNP ¼ DrNP , resulting in:
ð AÞ
Dre
a¼ ðBÞ
1 (17)
Dre

for the given reference cycle (Nref ) to failure, as illustrated in Fig. 3b.
ð AÞ ðBÞ
As can be seen, when Dre >Dre , a > 0, material is considered as being sen-
ð AÞ ðBÞ
sitive to nonproportional loading. When Dre Dre , a  0 is neutral to or even
benefits from the presence of nonproportional loading. It is important to point out
286 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

that the aforementioned procedure for determining a is not restricted to the avail-
ability of test data obtained using a circular load path for which gNP ¼ 1, which is
required by almost all nonproportional fatigue damage models available to date
(e.g., Itoh et al. [7]; Lee, Tjhung, and Jordan [34]; and others [35]). The procedure
presented here is applicable for any elliptical load path as long as the out of phase
angle d is known in Eq 6. As such, Eq 17 can be generalized as:
 ðAÞ 
Dre
ðBÞ  1
Dre
a¼ (18)
gNP

in which gNP is determined through Eq 13 for given r0 , s0 , and d. Note that the
material sensitivity parameter given in Eq 18 has been shown to have a direct correla-
tion with material ductility in [36]. Its application for formulating an effective stress
parameter for fatigue evaluation of welded components are given in [37].
For strain-life-based multi-axial data correlation, the corresponding material
sensitivity parameter ae p
given
ffiffiffiffiffi in Eq 5 can be determined in exactly the same manner
but with respect to e  be c plane. It would be desirable to establish a direct rela-
tionship between a and ae so that one determined from stress space (i.e., a) can be
used to infer nonproportional fatigue damage behavior in strain space (ae ) or vice
versa. Such a relationship would require considerations of the material cyclic consti-
tutive relationship under nonproportional loading conditions in both high-cycle and
low-cycle regimes, which will be a subject of an ongoing study by the authors.

Applications in Aluminum Alloys


EXTRACTION OF a OR ae
Six groups [9–15] of multi-axial (including proportional and nonproportional) test
data obtained from three types of aluminum alloys are considered here. These alu-
minum alloys are 2000, 6000, and 7000 series. As an illustration for applying the
procedure described in the previous section, Fig. 4 is a plot of strain-life curves for
hollow-tube specimen (see Fig. 5) test results obtained on aluminum alloy 7075-T6
[9,10] in terms of effective strain range
pffiffiffiffiffiversus cycle to failure, in which the effective
strain range is the distance on e  be c plane, where be ¼ 1=3, as discussed in the
earlier section on multi-axial fatigue damage parameters. The two sets of test data
can then be represented by the two lines through curve fitting. Because the two lines
have slightly different slopes, a mid-range life (i.e., Nref ¼ 100Þ is chosen here for
determining ae that will be used for analyzing test data obtained under more com-
ð AÞ ðBÞ
plex nonproportional load paths. Substituting ee and ee into exactly the same
form of Eq 18 but in terms of strain, a is found to be 0.35 (i.e.; ae ¼ 0:35).
e

Following the same procedure, ae for another set of 7075-T651 multi-axial


fatigue tests data [11] done on tubular specimens is also determined as ae ¼ 0:35.
Table 1 summarizes material sensitivity parameter a (from load-controlled tests) or
ae (from strain-controlled tests) calculated by comparing simple proportional and
MEI AND DONG, DOI 10.1520/STP159820160079 287

FIG. 4 Determination of ae using proportional and nonproportional (phase shift angle


of 90 ) test data, aluminum alloy 7075-T6.

nonproportional multi-axial fatigue test results for all three groups of aluminum
alloys, except for the test data from Itoh et al. [12] in which pure torsion test data
are compared with out-of-phase data because few in-phase data are available. Note
that within each of the aluminum alloy groups, a or ae is extracted from indepen-
dent sources when available. The following interesting observations can be made:

FIG. 5 Thin-walled tubular specimen test by Zamrick [9] and Zamrick and Frishmuth
[10].
288 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

TABLE 1 Summary of material sensitivity parameter a or ae determined for three types of


aluminum alloys.

Series of Aluminum Aluminum Alloys Loading Types a or ae


e
7000 series alloys 7075-T651 [9,10] Strain-controlled a ¼ 0.35
7075-T6 [11] Strain-controlled ae ¼ 0.35
6000 series alloys 6061 [12] Strain-controlled ae ¼ 0.4
6082-T6 [13] Load-controlled a ¼ 0.35
2000 series alloys 2024-T4 [14] Load-controlled a ¼ 0.5
2024-T4 [15] Strain-controlled ae ¼ 0.5

(a) Material sensitivity parameter a determined from stress-life data or ae deter-


mined from strain-life data seem to have a similar value, as shown for 2000
and 6000 series alloys.
(b) For each alloy type, there exists a reasonable degree of consistency in a or ae
values determined from tests conducted independently by different researchers,
suggesting that a or ae indeed is a material constant.
(c) For the three types of aluminum alloys investigated, the values of a and ae
are found to be in the range of 0.35–0.55, much less than those found for
structural steel and steel weldment tests, which can be treated as unity (i.e.,
a ¼ ae ¼ 1), as shown in Mei and Dong [27].
The validity of the material sensitivity parameter in terms of either a or ae sum-
marized in Table 1 will be further substantiated in the next section by applying the
resulting MLP model for correlating multi-axial fatigue tests under more complex
nonproportional loading conditions for components made of the three alloys.

CORRELATION OF TEST DATA WITH COMPLEX LOAD PATHS


7000 Series Aluminum Alloy
Zamrick [9] and Zamrick and Frishmuth [10] studied nonproportional multi-axial
low-cycle fatigue of 7075-T6 aluminum alloys by testing a large number of hollow-
tube specimens, as shown in Fig. 5. The tests involved pure tension, pure torsion, com-
bined tension, and torsion cyclic tests with a phase shift angle d being 0o, 30 , 45 ,
60 , and 90 (see Eq 6). To analyze these strain-life test results, the material sensitivity
parameters ae are already obtained and given in Table 1, and the corresponding load-
path nonproportionality factor damage factor for each phase angle d can be deter-
mined by Eq 13. Then Eq 5 can be used to calculate the MLP-based equivalent strain
DeNP corresponding to applied effective strain Dee without incorporating load-path
nonproportionality effects. The MLP-based equivalent strain corresponding to all test
data reported by Zamrik [9] are shown in Fig. 6, indicating an excellent correlation
among all test data with a standard deviation of 0.26 and a correlation coefficient of
0.86. It is worth noting that most of the fatigue life estimated by the MLP model is
MEI AND DONG, DOI 10.1520/STP159820160079 289

FIG. 6 Data correlation using MLP model–aluminum alloy 7075-T6 test data reported
by Zamrick [9] and Zamrick and Frishmuth [10].

1.0E+00
+2*std -2*std
Mean N prop loading
Pure tension Pure torsion test
90 out of phase 30 phase angle
MLP-based equivalent strain

45 phase angle 60 phase angle


1.0E-01

1.0E-02
STD = 0.26
R^2 = 0.86

1.0E-03
1.00E+00 1.00E+01 1.00E+02 1.00E+03 1.00E+04 1.00E+05
cycles to failures (N)

within a factor of two of tested fatigue life, except for those extremely low-cycle life
(N < 10) cases that may have tube buckling involved [9].
More recently, Zhao and Jiang [11] reported a series of nonproportional multi-
axial fatigue tests on 7075-T651 aluminum alloys using hollow-tube specimens with
fully reversed strain-controlled load paths, as shown in Fig. 7. Load paths (d) and (e)
in Fig. 7 correspond to asynchronous loading with torsion/tension frequency ratios of
and 4, respectively. For these complex load paths, the PDMR cycle-counting
procedure can be used to break the load blocks, that is, (e) and (f), into multiple cycles
(see [27,31–33]) and corresponding based effective stress ranges. These counted cycles
is then treated as one equivalent cycle corresponding to Miner’s rule based equivalent
strain range for data plotting and correlation purpose. Data correlation by the MLP-
based equivalent strain is shown in Fig. 8. It can be seen that nonproportional test data,
especially the data corresponding to asynchronous loading paths, fall in the neighbor-
hood of the mean curve, demonstrating the effectiveness of the MLP-based fatigue
damage model.
An alternate way of demonstrating the effectiveness of the MLP-based model is
by comparing MLP-based prediction with Smith-Watson-Topper (SWT) model
prediction [11], as shown in Fig. 9 and Fig. 10, respectively. Note that the dotted
lines in Fig. 9 and Fig. 10 represent factor-of-five in-life positions from the mean
line. It can be clearly seen that all but one data point fall within a factor of five in
290 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 7 Load paths for axial-torsion loading using tubular specimens [11].

FIG. 8 MLP-based equivalent strain correlation of Zhao and Jiang’s 7075-T651 [11]
aluminum alloy test data.

STD = 0.32
R^2 = 0.93
MEI AND DONG, DOI 10.1520/STP159820160079 291

FIG. 9 MLP-based fatigue life prediction vs. actual tests data [11] on 7075-T651
aluminum alloy.

Mean N
1.0E+07 prop loading
circular load path

asynchronous loading with torsion-axial load


1.0E+06 frequency rao of 2
asynchronous loading with torsion-axial
Predicted fague life (N)

load frequency rao of 4


Pure torsion test

1.0E+05

1.0E+04

1.0E+03

1.0E+02
1.00E+02 1.00E+03 1.00E+04 1.00E+05 1.00E+06 1.00E+07
Observed fague life (N)

FIG. 10 SWT-based fatigue life prediction given in Zhao and Jiang [11] versus tests on
7075-T651 aluminum alloy [11].

1.0E+07 Mean N
prop loading
circular load path
asynchronous loading with torsion-axial load
1.0E+06 frequency rao of 2
Predicted fague life (N)

asynchronous loading with torsion-axial


load frequency rao of 4
Pure torsion test
1.0E+05

1.0E+04

1.0E+03

1.0E+02
1.00E+02 1.00E+03 1.00E+04 1.00E+05 1.00E+06 1.00E+07
Observed fague life (N)
292 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 11 Strain path employed in Itoh et al. test of 6061 aluminum alloy [12].

FIG. 12 MLP-based equivalent strain correlation of 6061 aluminum alloy test data from
Itoh et al. [12].
MEI AND DONG, DOI 10.1520/STP159820160079 293

FIG. 13 Solid round bar specimen used by Susmel and Petrone [13].

life when the MLP model is used, showing a significant improvement over SWT-
based prediction presented by Zhao and Jiang [11].

6000 Series Aluminum Alloy


Itoh et al. [12] examined nonproportional low-cycle fatigue behaviors of 6061 alumi-
num alloys by conducting strain-life testing using 14 strain paths, as shown in Fig. 11.
As discussed in the section on extraction of a or ae , the material sensitivity

FIG. 14 MLP-based equivalent stress correlation of Susmel and Petrone’s 6082-T6


aluminum alloy test data [13].

pure torsion +2*std


In phase -2*std
Pure bending Out of phase
1.0E+03
MLP-based eq stress (MPa)

Mean N 126 degree out of phase loading

STD = 0.21
R^2 = 0.86
1.0E+02
1.00E+03 1.00E+04 1.00E+05 1.00E+06
cycles to failures (N)
294 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

parameter ae is calculated using test data according to Case 0 and Case 13. The use
of Case 0 rather than Case 5 is due to the consideration of the fact that there are
only two data points available, which is statistically insufficient for the purpose of
determining ae . Note that the PDMR cycle-counting procedure is used for Case 1
to Case 4, and illustrations for the PDMR cycle-counting details can be found in
Dong, Wei, and Hong [31] and in Wei and Dong [32,33]. The data correlation
results using MLP-based equivalent strain range are summarized in Fig. 12, again
showing an excellent correlation among all test data, with a standard deviation of
0.17 (STD ¼ 0.17). Among a total of 21 tests, 19 of them fall within a factor of 2 of
the mean curve.
By considering a similar aluminum alloy, 6082-T6, Susmel and Petrone [13]
performed a series of load-controlled multi-axial tests (pure torsion, bending,
combined loading considering both in-phase and out-of-phase with a phase angle
shift of roughly 90 and 126 ) using solid round bar specimens (see Fig. 13). The
MLP-based effective stress range is used in Fig. 14 for correlating all test data
reported by Susmel and Petrone [13]. Again, an excellent correlation among the
test data can be seen, with a standard deviation of around 0.21.

2000 Series Aluminum Alloy


Xia and Yao [14] recently conducted load-controlled nonproportional fatigue tests on
2024-T4 aluminum alloys with phase angles varying from 30o, 45 , 60 , to 90 between

FIG. 15 MLP-based equivalent stress correlation of Xia and Yao’s 2024-T4 aluminum
alloy [14] test data.

Pure tension Pure torsion


Proporonal 90 degree out of phase
30 degree out of phase 45 degree out of phase
1.0E+03
MLP-based equivalaent stress

60 degree out of phase Mean

STD = 0.14
R^2 = 0.92

1.0E+02
2.00E+03 2.00E+04 2.00E+05
cycles to failure
MEI AND DONG, DOI 10.1520/STP159820160079 295

FIG. 16 MLP-based equivalent strain correlation of 2024-T4 aluminum alloy test data of
Wang et al. [15].

1.0E-01
+2*std -2*std
Mean N pure axial
MLP-based equivalent strain

pure shear In phase


90 out of phase

1.0E-02

STD = 0.3
R^2 = 0.94

1.0E-03
5.00E+01 5.00E+02 5.00E+03 5.00E+04 5.00E+05
cycles to failure

tension and torsion. Their objectives mainly focused on the comparison of cumulative
damage rules for dealing with multi-axial block-loading spectrums. For the purpose of
this investigation, only “single-stage” test data [14] are considered here. A good correla-
tion can be seen in Fig. 15 once the MLP-based equivalent stress range is used in analyz-
ing these data. For the same type of aluminum alloy, strain-controlled nonproportional
test data were reported by Wang et al. [15]. The MLP-based equivalent strain versus
cycle to failure for this set of strain-controlled data is given in Fig. 16. Note thatpeven
ffiffiffiffiffi
pffiffiffi
though these two groups of test data are analyzed in r  bs plane and e  be c
plane, respectively and independently, the material-sensitivity parameter calculated
following the procedures in the earlier section on the multi-axial fatigue damage
parameter are of the same values for both cases, with a ¼ ae ¼ 0:5. What is more, sat-
isfactory correlation is achieved by both cases. The predicted fatigue life is within a fac-
tor of two of Xia and Yao’s [14] test results for all load cases, while a factor of three is
found between predicted life and the test results of Wang et al. [15] for most cases.

Conclusions
In this study, the application of a recently proposed MLP model for investigating
nonproportional multiaxial fatigue damage in a series of aluminum alloys is exam-
ined. Major findings are:
(a) All of the alloys studied in this paper show a lesser degree of sensitivity to
nonproportional loading when comparing structural steels recently reported
by the same authors, with material sensitivity parameters (a defined with
296 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

respect to stress plane or ae strain plane) ranging from 0.35 to 0.5. This is
much smaller than that of structural steels, which were found to be around
unity (i.e., a  ae  1).
(b) Within the same aluminum alloy type, it is found that the material sensitivity
parameter (a) calculated on stress plane using stress-life test data is rather
close to the values calculated in strain plane using strain-life data (i.e., ae ).
(c) With the material sensitivity parameter (a or ae ) being determined, MLP-
pffiffiffi
based equivalent
pffiffiffiffieffi stress (defined on r  bs plane) or equivalent strain
(on e  b c) are shown to be effective in correlating a large amount of
multiaxial fatigue test data subjected to a varying degree of load-path
nonproportionality.

References

[1] Socie, D. F. and Marquis, G., Multiaxial Fatigue, Society of Automotive Engineers, War-
rendale, PA, 1999.

[2] Skibicki, D., Phenomena and Computational Models of Nonproportional Fatigue of Mate-
rials, Springer, New York, 2014.

[3] Noban, M., Jahed, H., Ibrahim, E., and Ince, A., “Load Path Sensitivity and Fatigue Life Es-
timation of 30CrNiMo8HH,” Int. J. Fatigue, Vol. 37, 2012, pp. 123–133.

[4] Anes, V., Reis, L., Li, B., Fonte, M., and De Freitas, M., “New Approach for Analysis of
Complex Multiaxial Loading Paths,” Int. J. Fatigue, Vol. 62, 2014, pp. 21–33.

[5] Fatemi, A. and Shamsaei, N., “Multiaxial Fatigue: An Overview and Some Approximation
Models for Life Estimation,” Int. J. Fatigue, Vol. 33, No. 8, 2011, pp. 948–958.

[6] Sonsino, C. M. and Kueppers, M., “Multiaxial Fatigue of Welded Joints under Constant
and Variable Amplitude Loadings,” Fatigue Fract. Eng. M., Vol. 24, No. 5, 2001,
pp. 309–327.

[7] Itoh, T., Sakane, M., Ohnami, M., and Socie, D. F., “Nonproportional Low Cycle Fatigue
Criterion for Type 304 Stainless Steel,” J. Eng. Mater-T. ASME, Vol. 117, No. 3, 1995,
pp. 285–292.

[8] Shamsaei, N., Fatemi, A., and Socie, D. F., “Multiaxial Fatigue Evaluation Using Discrimi-
nating Strain Paths,” Int. J. Fatigue, Vol. 33, No. 4, 2011, pp. 597–609.

[9] Zamrik, S. Y., An Investigation of Strain Cycling Behavior of 7075-T6 Aluminum under
Combined State of Strain—The Effects of Out-of-Phase, Bi-Axial Strain Cycling on Low
Cycle Fatigue, NASA Technical Report CR-72843, National Aeronautics and Space
Administration, Washington, DC, 1972.

[10] Zamrik, S. Y. and Frishmuth, R. E., “The Effects of Out-of-Phase Biaxial-Strain Cycling on
Low-Cycle Fatigue,” Exp. Mech., Vol. 13, No. 5, 1973, pp. 204–208.

[11] Zhao, T. and Jiang, Y., “Fatigue of 7075-T651 Aluminum Alloy,” Int. J. Fatigue, Vol. 30,
No. 5, 2008, pp. 834–849.
MEI AND DONG, DOI 10.1520/STP159820160079 297

[12] Itoh, T., Nakata, T., Sakane, M., and Ohnami, M., “Nonproportional Low Cycle Fatigue of
6061 Aluminum Alloy under 14 Strain Paths,” ESIS Publ., Vol. 25, 1999, pp. 41–54.

[13] Susmel, L. and Petrone, N., “Multiaxial Fatigue Life Estimations for 6082-T6 Cylindrical
Specimens under In-Phase and Out-of-Phase Biaxial Loadings,” ESIS Publ., Vol. 31,
2003, pp. 83–104.

[14] Xia, T. X. and Yao, W. X., “Comparative Research on the Accumulative Damage Rules un-
der Multiaxial Block Loading Spectrum for 2024-T4 Aluminum Alloy,” Int. J. Fatigue, Vol.
48, 2013, pp. 257–265.

[15] Wang, X., Gao, Z., Qiu, B., Wang, L., and Jiang, Y., “Multi-Axial Fatigue of 2024-T4 Alumi-
num Alloy,” Chin. J. Mech. Eng., Vol. 24, No. 2, 2011, pp. 1–6.

[16] Wu, M., Itoh, T., Shimizu, Y., Nakamura, H., and Takanashi, M., “Low Cycle Fatigue Life
of Ti-6al-4v Alloy under Non-Proportional Loading,” Int. J. Fatigue, Vol. 44, 2012,
pp. 14–20, http://dx.doi.org/10.1016/j.ijfatigue.2012.06.006

[17] Fatemi, A. and Shamsaei, N., “Multiaxial Fatigue: An Overview and Some Approximation
Models for Life Estimation,” Int. J. Fatigue, Vol. 33, No. 8, 2011, pp. 948–958.

[18] Nakamura, H., Takanashi, M., Itoh, T., Wu, M., and Shimizu, Y., “Fatigue Crack Initiation
and Growth Behavior of Ti–6Al–4V under Non-Proportional Multiaxial Loading,” Int. J.
Fatigue, Vol. 33, No. 7, 2011, pp. 842–848.

[19] Shamsaei, N., Gladskyi, M., Panasovskyi, K., Shukaev, S., and Fatemi, A., “Multiaxial
Fatigue of Titanium Including Step Loading and Load Path Alteration and Sequence
Effects,” Int. J. Fatigue, Vol. 32, No. 11, 2010, pp. 1862–1874.

[20] Xiong, Y., Yu, Q., and Jiang, Y., “Multiaxial Fatigue of Extruded AZ31B Magnesium Alloy,”
Mater. Sci. Eng. A, Vol. 546, 2012, pp. 119–128.

[21] Wu, Z. R., Hu, X. T., and Song, Y. D., “Multiaxial Fatigue Life Prediction for Titanium Alloy
TC4 under Proportional and Nonproportional Loading,” Int. J. Fatigue, Vol. 59, 2014,
pp. 170–175.

[22] Shamsaei, N., Gladskyi, M., Panasovskyi, K., Shukaev, S., and Fatemi, A., “Multiaxial Fa-
tigue of Titanium Including Step Loading and Load Path Alteration and Sequence
Effects,” Int. J. Fatigue, Vol. 32, No. 11, pp. 1862–1874.

[23] Shamsaei, N., Fatemi, A., and Socie, D. F., “Multiaxial Fatigue Evaluation Using Discrimi-
nating Strain Paths,” Int. J. Fatigue, Vol.32, No. 11, pp. 597–609.

[24] Papadopoulos, I., “Critical Plane Approaches in High-Cycle Fatigue: On the Definition of
the Amplitude and Mean Value of the Shear Stress Acting on the Critical Plane,” Fatigue
Fract. Eng. M., Vol. 21, 1998, pp. 269–285.

[25] Li, B., Reis, L., and de Freitas, M., “Comparative Study of Multiaxial Fatigue Damage
Models for Ductile Structural Steels and Brittle Materials,” Int. J. Fatigue, Vol. 31, Nos.
11–12, 2009, pp. 1895–1906.

[26] Dong, P. and Hong, J., “A Robust Structural Stress Parameter for Evaluation of Multiaxial
Fatigue of Weldments,” J. ASTM Int., Vol. 3, No. 7, 2006, pp. 1–17, https://doi.org/
10.1520/JAI100348
298 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

[27] Mei, J. and Dong, P., “A New Path-Dependent Fatigue Damage Model for Non-Proportional
Multi-Axial Loading,” Int. J. Fatigue, Vol. 90, 2016, pp. 210–221.

[28] Kueppers, M. and Sonsino, C. M., “Critical Plane Approach for the Assessment of the
Fatigue Behaviour of Welded Aluminium under Multiaxial Loading,” Fatigue Fract. Eng.
M., Vol. 26, No. 6, 2003, pp. 507–513.

[29] Kueppers, M. and Sonsino, C. M., “Assessment of the Fatigue Behaviour of Welded
Aluminium Joints under Multiaxial Spectrum Loading by a Critical Plane Approach,” Int. J.
Fatigue, Vol.28, No. 5, 2006, pp. 540–546.

[30] Sonsino, C. M., “Influence of Material’s Ductility and Local Deformation Mode on Multiax-
ial Fatigue Response,” Int. J. Fatigue, Vol. 33, No. 8, 2011, pp. 930–947.

[31] Dong, P., Wei, Z., and Hong, J. K., “A Path-Dependent Cycle Counting Method for
Variable-Amplitude Multi-Axial Loading,” Int. J. Fatigue, Vol. 32, No. 4, 2010, pp. 720–734.

[32] Wei, Z. and Dong, P., “A Generalized Cycle Counting Criterion for Arbitrary Multi-Axial
Fatigue Loading Conditions,” J. Strain. Anal. Eng. Des., Vol. 49, 2014, pp. 325–341.

[33] Wei, Z. and Dong, P., “Multiaxial Fatigue Life Assessment of Welde,” Int. J. Fatigued
Structures, Eng. Fract. Mech., Vol. 77, No. 15, 2010, pp. 3011–3021.

[34] Lee, Y. L., Tjhung, T., and Jordan, A., “A Life Prediction Model for Welded Joints under
Multiaxial Variable Amplitude Loading Histories,” Int. J. Fatigue, Vol. 29, No. 6, 2007,
pp. 1162–1173.

[35] Babaei, S. and Ghasemi-Ghalebahman, A., “Damage-Based Modification for Fatigue Life
Prediction under Non-Proportional Loadings,” Int. J. Fatigue, Vol. 77, 2015, pp. 86–94.

[36] Mei J. and Dong P., “Modeling of Path-Dependent Multi-Axial Fatigue Damage in Alumi-
num Alloys,” Int. J. Fatigue, Vol. 95, 2017, pp. 252–263.

[37] Mei J. and Dong P., “An Equivalent Stress Parameter for Multi-Axial Fatigue Evaluation
of Welded Components Including Non-Proportional Loading Effects,” Int. J. Fatigue,
2017 (in press), http://dx.doi.org/10.1016/j.ijfatigue.2017.01.006
FATIGUE AND FRACTURE TEST PLANNING, TEST DATA ACQUISITIONS AND ANALYSIS 299

STP 1598, 2017 / available online at www.astm.org / doi: 10.1520/STP159820150099

Hoda Eskandari1 and Ho Sung Kim2

A Theory for Mathematical


Framework and Fatigue Damage
Function for the S-N Plane
Citation
Eskandari, H. and Kim, H. S., “A Theory for Mathematical Framework and Fatigue Damage
Function for the S-N Plane,” Fatigue and Fracture Test Planning, Test Data Acquisitions and
Analysis, ASTM STP1598, Z. Wei, K. Nikbin, P. C. McKeighan, and D. G. Harlow, Eds., ASTM
International, West Conshohocken, PA, 2017, pp. 299–336, http://dx.doi.org/10.1520/
STP1598201500993

ABSTRACT
Fatigue life predictions associated with the S-N curve have been largely based
on empiricism or unfounded theories, or both, due to the complexity involving
multiple variables such as fatigue life, applied stress, number of loading cycles,
and stress ratio. A damage model proposed by Palmgren in 1924 and
popularized by Miner in 1945 may be one of the most important milestones in
the fatigue damage research history. Its validity and principle adopted, however,
have been problematic, and many researchers have attempted refining the
model without much success. The ultimate objectives of the current work are to
provide a theory for mathematical validity framework of fatigue damage
associated with the S-N curve and to derive a damage function capable of
predicting the fatigue life. The validity framework is designed as the fundamental
basis to ensure the validity of a damage function in the development process. In
this paper, a theory for validity framework, consisting of axioms, relative
conditions for compatibility, and boundary conditions, is developed for fatigue
damage on the S-N plane at a stress ratio of zero. Also, compatibility conditions
based on the new concept for fatigue damage are developed. Manifestation
points for accumulated damage are defined and conceptualized for boundary

Manuscript received December 22, 2015; accepted for publication March 23, 2016.
1
The University of Newcastle, School of Engineering, University Drive, Callaghan, NSW 2308, Australia
2
The University of Newcastle, School of Engineering, University Drive, Callaghan, NSW 2308, Australia
http://orcid.org/0000-0001-9075-9069
3
ASTM Symposium on Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis on May 4–5,
2016 in Grand Hyatt, San Antonio, TX.

Copyright V
C 2017 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.
300 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

conditions by differentiating between damage accumulated before failure and


failure caused by damage. A selected equation for damage at failure as the
reference damage is theoretically validated for further validation of damage on
the S-N plane. Subsequently, a damage function capable of predicting fatigue
damage is proposed following the validation process. Comparisons between
experimental results from two stress level sequence loading and theoretical
fatigue life predictions are made and a close agreement between them was
found. With the benefits of the new compatibility conditions and criteria we
developed, an evaluative review on the fatigue damage models is presented.

Keywords
fatigue, S-N curve, damage, validity, damage function, theory, model, prediction,
remaining fatigue life, compatibility

Introduction
Fatigue is a phenomenon that causes progressive damage to materials subjected to
repeatedly fluctuating stress. Fatigue damage consists of cracks and localized plastic
deformation. It is the key to understanding fatigue behaviors of materials affected
by various factors including number of loading cycles, stress amplitude, stress ratio,
loading frequency, and temperature. Factors such as stress amplitude and stress
ratio may be the ones that are directly responsible for damage mechanisms leading
to failure at a certain fatigue life while loading frequency and temperature affect
the material properties prior to creating damage. Fatigue damage mechanisms,
therefore, are directly dependent on fatigue loading types such as tension-tension,
compression-compression, and tension-compression, which are dictated by stress
amplitude and stress ratio.
One of the most basic S-N curve data points can be experimentally obtained at
failure with a constant stress ratio with a given constant stress amplitude. When dif-
ferent stress amplitudes are employed at a constant stress ratio, an S-N curve can be
obtained from collected data points at failure. Further, when the stress ratio is var-
ied, different S-N curves can be experimentally generated. More generally, if stress
ratio and amplitude can be varied before failure, a spectrum or a random loading
pattern may be considered. Once an experimental S-N curve at a constant stress
ratio is found, a remaining fatigue life after being subjected to a certain number of
loading cycles at a given constant stress amplitude with a constant stress ratio can
be readily predicted. However, if a remaining fatigue life is to be predicted for a dif-
ferent stress amplitude but with the same stress ratio, fatigue damage is required to
be conceptualized for quantification associated with the prediction of the remaining
lives. If a remaining fatigue life is to be predicted for a different stress amplitude
with a different stress ratio, the constant fatigue life diagram [1] may be useful in
conjunction with a fatigue damage model. If a remaining fatigue life is to be pre-
dicted at a variable stress amplitude with a variable stress ratio for spectrum or
ESKANDARI AND KIM, DOI 10.1520/STP159820150099 301

random loading, a relatively complicated modeling involving more independent


variables may be required. Thus, using S-N curve(s) obtained at constant stress ra-
tio(s) with constant stress amplitude(s), the different stages corresponding to differ-
ent levels of difficulty for the remaining fatigue life predictions may be summarized
as follows:
Stage 1: For the same constant stress ratio but with a different constant stress ampli-
tude after being subjected to a constant stress amplitude with a constant stress
ratio (see Fig. 1a)
Stage 2: For a different constant stress ratio with a different constant stress amplitude
after being subjected to a constant stress amplitude with a constant stress ratio
(see Fig. 1b)

FIG. 1 Different stages for the remaining life predictions: (a) for the same constant
stress ratio but a different constant stress amplitude after being subjected to a
constant stress amplitude with a constant stress ratio; (b) for a different
constant stress ratio and a different constant stress amplitude after being
subjected to a constant stress amplitude with a constant stress ratio; (c) for the
same constant stress ratio and a variable stress amplitude after being subjected
to a variable stress amplitude with a constant stress ratio; and (d) for a variable
stress ratio with a variable constant stress amplitude after being subjected to a
variable stress amplitude with a variable stress ratio.

Previous fague
loading Remaining fague life
Previous fague
loading Remaining fague life
smax smax

Number of cycles (N) Number of cycles (N)

(a) (b)

Previous fague
Previous fague Remaining fague life
loading
loading Remaining fague life

s s

Number of cycles (N) Number of cycles (N)

(c) (d)
302 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

Stage 3: For the same constant stress ratio with a variable stress amplitude after being
subjected to a variable stress amplitude with a constant stress ratio (see Fig. 1c)
Stage 4: For a variable stress ratio with a variable constant stress amplitude after being
subjected to a variable stress amplitude with a variable stress ratio (see Fig. 1d)
The literature suggests that the life prediction models and theories produced
over much of the past century have not settled down for Stage 1 despite numerous
attempts to predict the remaining fatigue lives for materials subjected to various
loading conditions—even including spectrum or random loadings.
Since the dawn of fatigue history [2], the research on fatigue life predictions as-
sociated with an S-N curve has been largely based on empiricism or unfounded the-
ories, or both. The empiricism or unfounded theories may be due to the complexity
involving multiple variables (e.g., fatigue life, applied stress, number of loading
cycles, and stress ratio). A damage model proposed by Palmgren in 1924 [3] and
popularized by Miner in 1945 [4] may be one of the most important milestones in
fatigue damage research. Its validity and principle adopted in derivation, however,
have been problematic despite the fact that it might have inspired many researchers
in refining the model (e.g., [5–13]). In 1954, Marco and Starkey [5] brought to at-
tention the problem with the cumulative fatigue damage or fraction of damage
(N/Nf ¼ number of cycles/number of cycles at failure) proposed by Palmgren [3].
They not only pointed out the inaccuracy that stemmed from the model’s linearity
but also raised the conceptual problem of “damage.” Hwang and Han [13] also
pointed out the difficulty in defining fatigue damage. In fact, Miner [4] did not explic-
itly define “cumulative fatigue damage” or “fatigue damage” in terms of consequential
material properties despite its long-standing popularity and uses. Hwang and Han
[13] went on to compare 20 different empirical damage models and then produced
more models. In their conclusion, however, they expressed some frustration, stating
“It seems impossible to establish a universal fatigue damage model right now. But
the effort must be given to find it.” Fatemi and Yang [14] surveyed more than 50
different damage models for their review on fatigue. Even today, reportedly “new”
damage models (e.g., [15,16]) are still being produced, although some researchers
still adopt Palmgren’s damage model [3] for developing other applicative models
(e.g., [17,18]). At the same time, some theories available in the literature are without
full legitimacy in their theoretical foundations. The plethora of damage models may
be attributed to the failures of empiricism and unfounded theories, meaning that no
one model can serve as “the model.” Most damage predictions, though, have been
experimentally verified in one way or another—but at most for particular materials
and loading conditions. Philippidis and Passipoularidis [19] evaluated nine selected
damage models involving residual strength degradation. They found that no model
is consistent in accurately predicting the residual strength degradation behavior of
different composite laminates. Post, Case, and Lesko [20] also selectively evaluated
twelve fatigue life prediction models for spectrum loading applications; they sur-
mised that most of the investigators who proposed models simply fit them to their
experimental results with their predictive ability remaining uncertain.
ESKANDARI AND KIM, DOI 10.1520/STP159820150099 303

Unreliable fatigue life predictions of structural elements lead to various types of


accidents and cost human lives, resources, and productivity in various industries
associated with aircraft, nuclear power plants, pressure vessels, wind turbines, and
so on. One of the most important and difficult unsolved problems in S-N fatigue is
“fatigue damage function” for Stage 1 (Fig. 1a); another related problem is “stress ra-
tio effect” for Stage 2 (Fig. 1b). Some researchers (e.g., [21,22]), though, have
assumed a simplistic approach for stress ratio effect on fatigue damage by ignoring
its limitations, leading to the premature generalizations of their damage model. On
the other hand, the modeling of stress ratio for fatigue has been a discrete research
topic [23–26] ever since 1874 [1], even without considering damage modeling.
A missing link, as such, has been created between the two streams of research (i.e.,
damage and stress ratio). Also, other important missing elements in research appear
to be the fundamental principles for validation prior to experimental verification
for particular cases.
In light of the deficiencies in past approaches and methodology, quantification
of fatigue damage and relativity conditions for compatibility of variables are con-
ceptualized in this paper to deal with the long-standing fundamental problems. The
ultimate objectives of the current work are to provide a theory for validity frame-
work of fatigue damage associated with the S-N curve and to derive a damage func-
tion capable of predicting the fatigue life at a stress ratio of zero. Also, an evaluative
review on selected fatigue damage models is given having the benefits of the new
compatibility concept and criteria developed in this paper. The validity framework
is designed as the fundamental basis for ensuring the validity of a damage function
in the development process.

Axioms for Fatigue Damage Quantification and


4
Compatibility Concept
The structural change within a material after being subjected to fatigue loading may
appear in various forms such as localized plastic deformation, cracking, delamina-
tion (if a composite), and so on. Such changes affected by loading conditions may
be called “damage” and are reflected in the consequential material properties (e.g.,
residual strength). There may be two different ways to quantify damage. One way is
to deal with individual cracks and deformation using fracture mechanics parame-
ters such as the stress intensity factor [27]. This may be useful if individual cracking
behaviors are reproducible for damage quantity every time an unnotched fatigue
specimen is tested. However, on a smaller scale, such behaviors of some composite
materials are neither reproducible nor predictable because of uncertainties in inher-
ent material defects [28]. The other way of quantifying damage is in a collective

4
A preliminary discussion contained in part of this section was presented at the 13th International Confer-
ence on Fracture and Damage Mechanics, September 23–25, 2014, Ponta Delgado, Azores, Portugal.
304 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

manner at a larger scale for a better predictability such that the damage magnitude
caused by multiple “small” cracks [29] is more reproducible as the scale size
increases. Damage, in this paper, will refer to the latter for a larger volume of
unnotched fatigue specimens containing multiple “small” cracks with small-scale
plastic deformation.
The quantification in a general complex problem is to find the relative differ-
ences within a given set of variables for making comparisons. Similarly, the damage
quantity does not necessarily have to be absolute but relative for comparison pur-
poses. A specific purpose may be adopted for the fatigue damage as follows. The
purpose of damage quantification is to provide a relative quantity resulting from the
property change and to predict the relative capability of resisting fatigue. According-
ly, understanding the relativity of variables, constants, and boundary conditions
may be important for damage quantification.
The quantification of damage for fatigue life prediction may now be consid-
ered for possible different systems as shown in Fig. 2, depending on how we select
system elements (i.e., variables and constants). An open system consists of ele-
ments for input, output, and system constants [30,31], in which two different
types of variables and one set of system constants may be identified for developing
a predictive damage function. Dependent variables (e.g., residual strength, residu-
al modulus, and number of cycles at failure Nf) form one of the types and are use-
ful for description of the consequential property change. Independent variables
form another type and are useful for description of the loading conditions (e.g.,
applied peak stress level [rmax] and number of applied loading cycles [N]). The
system constants are also useful not only for loading conditions (e.g., stress ratio
and frequency) but also can be selected as independent variables depending on

FIG. 2 A fatigue system structure consisting of input and output with system
constants.

Input (Independent
variables for
loading condition Consequential properties
such as N and smax, (dependent variables e.g., residual
etc.) strength, modulus, etc.)

System constants (e.g., stress ratio, Output (fatigue damage


frequency, etc.) quantity for fatigue life
prediction)
ESKANDARI AND KIM, DOI 10.1520/STP159820150099 305

the purpose of a system set-up. Some dependent variables may also be selected as
system constants for a simple analysis. In a system, input may consist of indepen-
dent variables for loading condition while the output is damage quantity from
damage function (D) for fatigue life prediction. Thus, the output is dependent on
the independent variables and system constants we select and on what damage
function is used. Discovering valid damage function may not necessarily be
straightforward as learned from the fatigue history, and the difficulty increases
with an increasing number of independent variables. If we reduce the number of
independent variables but increase the number of system constants, damage
quantification becomes simpler.
For conceptualizing compatibility, we may consider some previously proposed
damage equations/definitions used by researchers in the past. The equation
D ¼ N=Nf [4] used by many researchers is a function of a single independent vari-
able (N) with the number of cycles to failure (Nf) as the system constant or bound-
ary condition. It may be valid for damage quantification for fatigue life prediction
but only at a given rmax and system constants. If we choose N and rmax as two inde-
P
pendent input variables, we may consider Df ¼ Ni =Nfi ¼ 1 (where subscript i
indicates different peak stress levels). The equation may be capable of predicting
the fatigue life for certain sets of rmax and N values, given that the equation indi-
cates that, for example, damage (D) calculated from N1 ¼ 500 cycles and
Nf1 ¼ 1,000 cycles at an applied peak stress level is equal to that from different val-
ues N2 ¼ 50 cycles and Nf2 ¼ 100 cycles at another peak stress level. However, one
problem here is that we do not know if two different damage conditions indicated
by D value (¼ 0.5) provide the same capability of resisting fatigue because the com-
patibility between N and rmax for damage (D) is not known. Another difficulty is
that Nf (which is a system constant in the previous equation, D ¼ N=Nf ) turned
into a variable dependent on applied peak stress (rmax) as rmax is selected as an ad-
ditional independent variable. As a result, compatibility conditions between N and
rmax must be set up. The compatibility is to ensure that no overlaps in damage
quantities develop when damage varies as a function of multiple variables such as
N, Nf, and rmax such that a damage quantity should be a single value for a given
damage severity. If a damage prediction equation lacks compatibility, for example, a
single damage quantity is produced for different damage severities or vice versa.
Also, it is mandatory that the compatibility conditions be developed to comply with
the principles given in the following damage axioms at any point on the S-N plane:
(a) Fatigue damage increases monotonically with an increasing number of load-
ing cycles (N).
(b) Fatigue damage increases monotonically with an increasing applied peak
stress level (rmax).
(c) Residual strength decreases monotonically with increasing fatigue damage
accumulated within a material.
The damage axioms are the primitive principles we propose for developing the
derivative principles in the following sections.
306 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

A Validity Framework Theory for Fatigue


Damage and Damage Function
Two different damage types on the S-N plane generally may be considered. One
is applicable for fiber-reinforced composites or brittle materials and the other
for ductile materials having a yield point distinctively lower than an ultimate
strength. The latter (ductile materials) may tend to have two distinctive damage
mechanisms in terms of deformation depending on the applied peak stress level
while the former may have only one. One of the two mechanisms may be associat-
ed with damage affected by a general yielding (plastic deformation) prevalent over
the whole gauge length of a fatigue specimen at a high range of peak stress levels
or at a low range of cycles. The other is for that affected by the localized yielding
around “small” crack tips or defects without general yielding, which may occur in
brittle or fiber-reinforced composite materials for a full range of peak stress levels.
The current work will focus on the damage of a single mechanism applicable for
brittle or composite materials. Also, an S-N curve and damage are assumed to be
continuous for life prediction purposes, although the number of cycles is
discrete—as will be discussed, and to represent mean quantities of statistical scat-
ter because the fatigue data contain uncertainty from various sources. Stress ratio
(equal to valley stress/peak stress) is also assumed to be zero, which would be a
basis for other stress ratios in the future.
Damage will be considered for three different domains on the S-N plane. The
first domain is for damage accumulated up to the failure points, which are located
on the S-N curve, and simultaneously a set of boundary conditions for given con-
stant peak stress levels. The second domain is for damage at points for N ¼ 0.5 cycle
under the S-N curve, which is simultaneously another set of boundary conditions
for a given set of applied peak stress levels. The third domain is for all other points
between the two sets of boundary conditions under the S-N curve.

BOUNDARY CONDITIONS AND DAMAGE AT HALF CYCLE


It may be conceptually important to differentiate between damage accumulated
before failure (DABF) and failure caused by damage (FCD) for understanding the
nature of the S-N plane and for finding boundary conditions with quantized N.
Also, it may be equally important to find a common property between DABF and
FCD.
The failure in fatigue occurs after damage takes place although the damage can
occur without failure. Thus, DABF and FCD are different from each other and
occur sequentially at different stages. A common property between the two, how-
ever, is residual strength because it indicates part of a condition for either DABF or
FCD such that it can be used as a parametric quantity between DABF and FCD.
Accordingly, the residual strength itself is not necessarily the compatible damage
quantity (as will be discussed here for the “reference damage”), but damage (Df) at
failure can be a function of the residual strength.
ESKANDARI AND KIM, DOI 10.1520/STP159820150099 307

As the number of cycles to failure (Nf) increases, damage and failure may more
and more practically behave in a continuous manner on an S-N curve. However,
when Nf is small, DABF and FCD behaviors tend to be dependent on the quantized
variable N. As illustrated in Fig. 3 for a linear scale S-N curve, when N ¼ 0, no dam-
age is created by loading, and its initial material property (e.g., strength) is neither
known nor recordable on an S-N curve that is practically a locus of FCD points.
The FCD at the highest peak stress level may occur at N ¼ 0.5 (to be precise) as
indicated with Point a on the S-N curve in Fig. 3, given that the failure is not
expected to occur during unloading between N ¼ 0.5 and 1, although it may be
nominally regarded at N ¼ 1. Also, it should be noted that DABF exists for N < 0.5
but does not manifest itself as failure at any peak stress level above the fatigue limit
(see quasi-boundary conditions here). The next higher number of cycles for FCD
corresponding to a peak stress level would occur at N ¼ 1.5 (Point b in Fig. 3a and d),
and then N ¼ 2.5 (Point c in Fig. 3) and so on. Thus, the highest applied stress level
a damaged specimen can resist prior to failure, which is a residual strength or stress
at FCD, is equal to the applied peak stress level (rmax) if we use a number of cycles
N-0.5. According to the damage axiom (a), the relativity in damage quantity on the

FIG. 3 Schematic S-N curve and cyclic loading: (a) failure points (a, b, and c) on S-N
curve on linear scales; (b) point a on cyclic loading curve corresponding to that
on S-N curve; (c) failure stress is lower at N ¼ 1.3 than at N ¼ 1.5; and (d) point b
on cyclic loading curve corresponding to that on S-N curve.

a
sa Da = Df

Db > Da
sb b N = 1.3
smax
c Dc > Db
sc

0 0.5 1.5 2.5


N
(a)
N = 0.5 cycle smax = sa smax < sb smax = sb
N = 1.3 N = 1.5
a

smax
b

(b) (c) (d)


308 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

S-N curve can be found. For example, the damage accumulated (Db) up to N ¼ 1.5 is
higher than Da at N ¼ 0.5, Dc is higher than Db, and so on. Therefore, its damage
behaves as a function of discrete variable N-0.5 where N is an integer N  1. Inciden-
tally, however, the failure stress levels at a relatively small value of N-0.5 other than
integers (N) but greater than N ¼ 1 (e.g., 1.3, 1.4, etc.) would theoretically deviate
from the continuous S-N curve depending on the initially set applied peak stress
because the failure stress (or residual strength) is not equal to the peak stress at N-0.
For example, rmax at N ¼ 1.3 as given in Fig. 3a and 3c is lower than rb at N ¼ 1.5
because damage caused by the first applied peak stress rmax (in Fig. 3c) is higher than
that in the case given in Fig. 3d.
No practical difference between discrete numbers (e.g., 0.5, 1.5, etc.) and
other continuous numbers may be found for a relatively large Nf. Accordingly,
most of the S-N curve may be treated as a continuous curve for analytical pur-
poses. At the same time, such an analytical description for a small and quantized
N is useful for finding boundary conditions. Now, the two damage values at
N ¼ 0 and at N ¼ 0.5 may be considered at certain stress levels for boundary con-
ditions on the S-N plane for damage at FCD and DABF. The first one (N ¼ 0)
has been used in the literature for zero damage (e.g., [13]) but its location is not
clear on the S-N curve without a defined stress level. It, in fact, should not be
used as a boundary condition because zero damage at N ¼ 0 is neither part of
a continuous S-N curve nor part of a continuous S-N plane at any stress level
except a stress level of zero. However, the damage values at Nf ¼ 0.5 is legitimate
for boundary conditions if the S-N curve is assumed to be a continuous line and
only if we know damage values at corresponding stress levels as will be further
analyzed and discussed.

REQUIREMENT OF REFERENCE DAMAGE EQUATION FOR POINTS ON S-N


CURVE
The damage at failure (Df) appearing on an S-N curve may be used as the refer-
ence damage if quantifiable because it is physically known damage such that
damage at all other points can be determined relatively. The magnitude of Df
depends on the location on the S-N curve. Let us consider the damage at Point
A (DfA) and Point B (DfB) in Fig. 4, preferably with a logarithmic scale for N.
A relative relation should be DfA > DfB to comply with the principles given in
the axioms introduced earlier. The reason why DfA is more severe than DfB is
that the residual strength at Point A is lower than that at Point B. Accordingly,
the damage at points on an S-N curve increases with decreasing rmax and its
damage change rate is given by @Df =@rmax ; simultaneously, the corresponding
Df at FCD on the S-N curve changes with a rate of @Df =@ log Nf . However, the
relation between @Df =@ log Nf and @Df =@rmax cannot be known until an S-N
curve from the experiment is obtained.
To find an equation (Df) for describing a set of damage reference points (or
a boundary line), @Df =@rmax is required to be a constant greater than zero.
ESKANDARI AND KIM, DOI 10.1520/STP159820150099 309

FIG. 4 Schematic representation of S-N curve and various points. Inset 1 is for details of
isodamage line slope at point f with respect to slope of line between points
g and f. Inset 2 is for details of point b in relation with points A, B, and C.

Isodamage line slope at point f

Dsmax
B θI
D’
h Dlog N f
G
smax g
f F Inset 1
D A
B
C b

Ddc
Dsmax
θb
A
C b
O Δlog N
E
Log N Δlog Nf
N = 0.5
(or log N = -0.3) Inset 2

The reasons are that: (a) if @Df =@rmax ¼ 0, it violates the principle given in the
aforementioned axiom (b); and (b) if @Df =@rmax is a variable, the theoretical for-
mulation would be convoluted and unnecessarily complex. Therefore, a candidate
reference damage equation (Df) is required to satisfy the following equation:

@Df
¼ A (1)
@rmax

where:
A ¼ constant greater than zero.
The description of derivatives in relation to an S-N curve is presented schemati-
cally in Fig. 5.

A DAMAGE EQUATION AT FAILURE, BOUNDARY CONDITIONS AT N ¼ 0.5,


AND RELATIVITY
The following equation for damage at failure (Df) may be considered as a candidate
reference damage or a damage boundary line:

rmax
Df ¼ B  C for N  0:5 (2)
ra
310 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 5 Schematic relations among derivatives.

smax

dsmax
d(log Nf)

dsmax
dDf

Log Nf

dDf
Df d(log Nf)

where:
B, C ¼ constants, and
ra ¼ breaking stress (see Fig. 3) at N ¼ 0.5 cycle, which is similar to the static
ultimate stress (ru ) i.e.ra  ru .

The damage at failure (Df) depends on constants B and C, and should not be zero
at a peak stress level for N ¼ 0.5 as discussed earlier. Accordingly, it is required that
B > Crmax =ra or B > C. If we choose C ¼ 1 and any value greater than 1 for B,
Eq 2 satisfies Eq 1 and therefore is valid. If we use relatively large values for B for a
boundary condition, however, it will be difficult to determine DABF values at differ-
ent peak stress levels because we do not know yet how DABF varies at N ¼ 0.5
for different peak stress values. Therefore, if we adopt a value approaching unity
for B, Df at N ¼ 0.5 may be used as a value approaching zero, and then damage D at
N ¼ 0.5 at all different stress levels approaches zero as well. Thus, the value Df on
the S-N curve can be found with B ¼ 1. (Note that the way of establishing the value
for B does not violate the relativity of damage quantities.) Therefore, the value of
zero for damage at all different peak stress levels at N ¼ 0.5 may be used as the
quasi-boundary conditions for damage at other points. A reason why it is termed
quasi is that (a) the damage may still vary as a function of N but smaller than 0.5
cycle so that it is a function of a sub-scale variable different from multiples of 0.5
cycle; and (b) damage values at different peak stress levels are not theoretically
known, but zero is practically acceptable. Kim and Zhang [32] already demonstrat-
ed that, using Eq 2 with the same values (i.e., B ¼ C ¼ 1) for constants for materials
complying with @Df =@Nf ¼ armax b, a theoretical S-N curve becomes:
"  #
rb
u rmax 1b
Nf ¼  1 þ N0 (3)
aðb  1Þ ru
ESKANDARI AND KIM, DOI 10.1520/STP159820150099 311

where:
a, b ¼ fitting parameters.

The term N0 in Eq 3 is added to the original equation to define the initial cycle at
N ¼ 0.5 cycle failure point according to the quantized analysis. Eq 3 is useful for
curve fitting, with a and b as fitting parameters, for a wide range of cycles including
low-cycle loadings discussed by Harik, Klinger, and Bogetti [33]. Curve fitting is
required for calculating damage at failure (Df) in conjunction with Eq 2.

DAMAGE CHARACTERISTICS ON S-N PLANE AND COMPATIBILITY


CONDITIONS
Let us consider damage at any point under an S-N curve, which may be repre-
sented by Point C as shown in Fig. 4. Point C is located on the horizontal line DA
and is intersected by a vertical line BE. Points A, B, D, and E are accordingly
dependent on the location of Point C. In other words, if Point C moves toward
Point A, Points A and B become closer to each other on the S-N curve; simulta-
neously, Point E moves along the axis of log N toward the S-N curve. The damage
at Point C (denoted by DC) needs to be quantified relative to the reference damage
at points on the S-N curve. According to the damage axioms, we find that DC is
lower than damage at Point B (DfB) because it is subjected to a lower applied peak
stress at a given number of cycles. It is also lower than damage at Point A (DfA)
because it is subjected to a lower number of cycles (N) at a given applied
peak stress level. Accordingly, the following damage order shown in Eq 4 is
preserved:
DC < DfB < DfA (4)

and, therefore, the differences in damage between points follow Eq 5:

DDBC < DDAC (5)


where:
DDBC ¼ DfB - DC, and
DDAC ¼ DfA - DC.

Accordingly, we find, in Eq 6, that:


 
 DDBC 
 
DD  < 1 (6)
AC

for Point C. Also, as shown in Eqs 7 and 8, damage at any point in hatched area ABC
(DABC) and damage at any point in hatched rectangular area CDEO (DCDEO) (Fig. 4)
are found to be:

DCDEO  DC  DABC (7)


and

DD < DC < DB : (8)


312 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

To find upper and lower bounds for the slope of an isodamage line under an S-N
curve, we may consider damage (D) in hatched area ABC. Given DfG < DfF, we find
damage at a midpoint g (Dg) between Points G and C and damage at another mid-
point f (Df) between Points F and C have the following relation when Points G and
F are close to each other:

Dg < Df (9)

so that the angle (hI) of an isodamage line at Point f (see Inset 1 in Fig. 4) under an
S-N curve is required to satisfy Eq 10:
 
 Drmax 
tan1   < hI < p : (10)
D log N  2

Also, it is found that the slope (¼ j@rmax =@ log N j) of an isodamage line


decreases with increasing logN at a constant peak stress level because hI  p/2 at
N ¼ 0.5, which is the highest slope; but the lowest slope occurs at failure because its
monotonicity is maintained according to the axioms. Accordingly, a lower bound of
 
isodamage line slope is found to be j@rmax =@ log N j > @rmax =@ log Nf  at a con-
stant peak stress level.
Retrospectively, Eq 9 may be explained using a relation between two unknown
functions by introducing a local relative location factor (dr). The unknown damage
functions are shown in Fig. 6 with Points F and G and midpoints f and g for damage
as functions of dr. The local relative location factor, for example, dr ¼ 0, indicates
zero distance from Point C; dr ¼ 1 indicates a full distance, such as Points F and G
on the S-N curve (Fig. 4); and dr ¼ 0.5 indicates a midpoint. The damage monotoni-
cally increases smoothly and similarly due to the proximity of Points F and G, such
that the difference in damage between Points f and g at a given dr also

FIG. 6 Damage versus local relative location factor (dr) for points F and G, and
midpoints f and g shown in Fig. 4.

DfF F
Damage (D)

DfG
f G

g
C

0 0.5 1
Local relative location factor (dr)
ESKANDARI AND KIM, DOI 10.1520/STP159820150099 313

monotonically increases according to the axioms due to the difference in damage of


Points F and G and also due to the differences in logN and rmax. Therefore, it is
found Dg < Df if DfG < DfF, and this relation is true for all other pairs of points at a
given dr.
Also, for isodamage lines in relation with the S-N curve, if damage at Point b
is equal to that at Point B (Fig. 4), the angle (hb) of line connecting Point B and
Point b (see Inset 2 in Fig. 4) is also required to satisfy Eq 11:
 

1  Drmax 
 p
tan  < hb < : (11)
D log Nf  2

Let us define a general relative location factor of Point C as dfC ¼


ðlog NC þ 0:3Þ=ðlog NfC þ 0:3Þ, where NC is the number of cycles at Point C and
NfC is the number of cycles at failure at Point A (e.g., dfC ¼ 0 at Point D and dfC ¼ 1
at Point A). Then, more general relative relationships among the three different
points (i.e., B, b, and C) at a given applied peak stress level above the fatigue endur-
ance limit may be schematically seen in Fig. 7. The upper curve (in Fig. 7) represents
damage (D) for Point B corresponding to df of Point C in Fig. 4, and the lower curve
(in Fig. 7) represents damage (D) for Points b and C in Fig. 4. The upper curve can
be found from an S-N curve (see Eq 3) and Eq 2 with B ¼ C ¼ 1, but the lower curve
is not known until a damage function (D) is found. It is known, however, that
Point b should be on the right side of Point B at a given damage in Fig. 7 and that
Point C should be lower than Point B at a given df. The upper curve would be either
concave or convex depending on the S-N curve shape. If the S-N curve is linear, it
becomes linear. If part of the S-N curve is convex, its corresponding part becomes
concave and vice versa. If two curves vary smoothly (Fig. 7), both DDBC and DdfBb
(¼dfb  dfB) should increase initially and then decrease as Point C moves toward
Point A on the S-N plane (Fig. 4) for any nonlinear monotonic variation of damage

FIG. 7 Damage for points b, B, and C as a function of general relative location factor
(df) of a point on line DA on the S-N plane (see Fig. 4).

ΔdfBb

DfA A

ΔDAC
Damage (D)

DfB B
b
DC
C ΔDBC
0 0.5 1
General relative location factor (df)
314 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

while DDAC always decreases at a given peak stress level. The compatibility between
log N and rmax for damage on the S-N plane can be achieved when a valid set of
points of b for isodamage lines is found. One of the necessary conditions for the
compatibility is that df of Point b should be higher than Point C (Fig. 7) or
the location of Point b on the S-N plane is required to be between Points C and A
(Fig. 4), according to the axioms. Hence, the general relative location factors at
Points b and C are required to be:

dfb > dfC for 1 > dfC > 0; (12a)


dfb  dfC  0 at N ¼ 0:5 cycle; (12b)

and

dfb  dfC  1 at failures on the S-N curve (12c)

to maintain the general damage characteristics on the S-N plane described earlier. The
value of dfb is dependent on the damage mechanism and the S-N curve characteristics
for a given material.

A DAMAGE FUNCTION AND VALIDATION


A damage function (D) is required to be monotonic over an interval between two
points on a scale of logN at a given peak stress level. Accordingly, the following
damage function (D) may be considered:

D ¼ Df dfn (13a)

where:
n ¼ unknown exponent.

Eq 13a satisfies the boundary conditions (i.e., D ¼ 0 at df ¼ 0 or logN ¼ 0.3,


and D ¼ Df at df ¼ 1 or logN ¼ logNf). Also, it is noted that n governs the
change rate of damage (@D=@N ¼ nDf dfn1 ) at a given applied peak stress level
such that, unless a transition of damage mechanism occurs for another mecha-
nism, n should be a single value at various applied peak stress levels. The value of
n is greater than unity and depends on nonlinearity of the S-N curve as described
in Fig. 7.
Let us also consider damage on two horizontal lines DA and D0 B on the S-N
plane (Fig. 4). The upper curve (Fig. 8) represents a damage function (DA) for the
points on line DA (Fig. 4) given by:
n
DA ¼ DfA dfA (13b)

and the lower curve (Fig. 8) represents a damage function (DB) for the points on line
D0 B (Fig. 4) by:
n
DB ¼ DfB dfB : (13c)
ESKANDARI AND KIM, DOI 10.1520/STP159820150099 315

FIG. 8 Two unknown nonlinear functions: The upper curve is to represent a damage
function for the points on line DA and the lower curve is to represent the same
on line D0 B on the S-N plane (see Fig. 4).

DfA
A
DfB b
B
Damage (D) DA
DB

D and D’
0 1 1
⎠ Df B ⎞ n
d b =
Df A ⎠

Relative location factor (df)

The general relative location factor for Point b (dfb) whose damage is equal to
damage at Point B (Fig. 8) is found to be:
 n1
DfB
dfb ¼ (14)
DfA
where:
DfA, DfB ¼ known damage values at two points A and B, respectively, according
to Eq 2 (with constants B ¼ C ¼ 1).

Now, there would be ranges of different values of DdfBb (¼dfb – dfB) affected by the
S-N curve and damage function (D) with a given exponent n (see Fig. 7) at various
applied peak stress levels on the S-N plane. A value of n is now required to satisfy
the conditions given in Eq 12 for validation. A practical technique for determining a
valid n value will be demonstrated, using an example, in the next section.

Theoretical Predictions and Experimental


Results
Experimental results obtained by Broutman and Sahu [6] were adopted to demon-
strate a practical technique for determination of unknown exponent n in damage
function in Eq 13 and for verification of the predictions based on the current
proposed theory. Fig. 9 shows an S-N curve represented by Eq 3 fitted to experi-
mental data for cross-ply glass-fiber-reinforced plastic laminate with an ultimate
strength of 448 MPa and a stress ratio (R) of 0.05 (which is close to zero). Fitting
parameters for Eq 3 were found to be a ¼ 2.8  1041 and b ¼ 14.2. The dashed
lines on both sides of the solid line represent experimental failure probabilities,
316 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 9 The solid line represents an S-N curve represented by Eq 3 fitted to Broutman
and Sahu’s data [6] for cross-ply, glass-fiber-reinforced plastic laminate with an
ultimate strength of 448 MPa and a stress ratio (R) of 0.05. Fitting parameters
for Eq 3 were found to be a ¼ 2.8  1041 and b ¼ 14.2. The dashed lines on both
sides of the solid line represent failure probabilities, 10 % and 90 %.

400

350 P=90 %

300

250
σmax (MPa)

200

150 P=10 %

100

50

0
2 3 4 5 6 7
log N

10 % and 90 %. The experiment was conducted for two sequential blocks of load-
ing, each of which consisted of cycles of constant peak stress level. Each specimen
was cycled at a constant peak stress level 386, 338, 290, or 241 MPa and then
changed to another constant peak stress level until failure occurred. Table 1 lists
the following: peak stress levels (denoted by r1max) first applied in the sequence,
Log10Nf1 (cycles to failure at r1max), the second peak stress level (r2max) applied
after r1max, Log10Nf2 (Nf2 is the number of cycles to failure at r2max), and other
data. In the sequence of loading, N1 is the number of cycles the specimen was sub-
jected to at the first stress level r1max prior to being subjected to r2max with DN2
(the number of cycles to failure at r2max following cycling at r1max). A general rel-
ative location factor (df1) at the end of cycling at r1max is given by (logN1þ0.3)/
(logNf1þ0.3) (see Eq 13b), which is equivalent in damage to another general rela-
tive location factor (df2) (see Eq 13c) at r2max. The predicted and experimental
values for Nf2 are also listed in Table 1.
TABLE 1 Two stress level tests in sequence for comparison with theoretical predictions.

No r1max, MPa Log10Nf1, cycles r2max, MPa Log10Nf2 N_1/N_f1 Log10 DN2, Cycles, Experiment Log10 Nf2, Cycles, Predicted df1 df2 Accu

1 386 2.693 241 5.236 0.507 5.283 5.287 0.897 0.630 0.01
2 386 2.693 241 5.236 0.203 5.286 5.287 0.759 0.533 0.01
3 386 2.693 290 4.168 0.507 3.766 3.813 0.897 0.681 0.11
4 386 2.693 290 4.168 0.203 4.078 4.086 0.759 0.576 0.05
5 386 2.693 338 3.393 0.507 3.097 3.194 0.897 0.758 0.06
6 386 2.693 338 3.393 0.203 3.214 3.243 0.759 0.641 0.04
7 338 3.393 241 5.236 0.405 4.934 4.971 0.894 0.743 0.07
8 338 3.393 241 5.236 0.101 5.211 5.214 0.730 0.607 0.02
9 338 3.393 290 4.168 0.405 3.938 4.045 0.894 0.803 0.07
10 338 3.393 290 4.168 0.101 3.903 3.930 0.730 0.656 0.08
11 290 4.168 241 5.236 0.680 5.236 5.171 0.963 0.891 0.03
12 290 4.168 241 5.236 0.136 5.236 5.076 0.811 0.750 0.05

ESKANDARI AND KIM, DOI 10.1520/STP159820150099


13 241 5.236 290 4.168 0.290 3.572 4.279 0.905 0.978 20.00
14 241 5.236 290 4.168 0.116 3.977 4.212 0.834 0.902 0.02
15 241 5.236 338 3.393 0.290 2.592 3.749 0.905 1.088 0.11
16 241 5.236 338 3.393 0.116 2.905 3.524 0.834 1.033 0.04
17 241 5.236 386 2.693 0.290 N/A N/A 0.905  
18 241 5.236 386 2.693 0.116 2.093 3.152 0.834 1.188 0.23
19 290 4.168 338 3.393 0.68 2.467 3.685 0.963 1.072 0.09
20 290 4.168 338 3.393 0.136 3.111 3.374 0.811 0.902 0.01
21 290 4.168 386 2.693 0.680 N/A N/A 0.963 – –
22 290 4.168 386 2.693 0.136 2.550 2.975 0.811 1.069 0.16
23 338 3.393 386 2.693 0.405 2.473 2.928 0.894 1.058 0.14
24 338 3.393 386 2.693 1.020 2.702 2.817 0.730 0.865 0.10

Note: r1max ¼ the peak stress level first applied in the sequence; Nf1 ¼ cycles to failure at r1max; r2max ¼ the second peak stress level applied after r1max; Nf2 ¼ cycles to failure
at r2max; N1 ¼ the number of cycles the specimen was subjected to at the first stress level r1max; DN2 ¼ the number of cycles to failure at the second stress level r2max;
df1 ¼[(log n1þ0.3)/(log Nf1þ0.3)] ¼ a general relative location factor at the end of cycling at r1max; df2 ¼ a general relative location factor at r2max equivalent to df1 in
damage; Accu ¼ [(log Nf 2 from experimental DN2 and predicted df2) – (log Nf 2 from S-N curve)]/(log Nf 2 from S-N curve).

317
318 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

THEORETICAL VALIDATION AND CALCULATION


The predictions for Log10N2 were made after finding a valid value of exponent n
in Eq 13a. For validation, Eq 14 was used to find a general location factor (dfb)
for Point b whose damage is equal to Point B (Fig. 4), and a general location
factor of Point C (Fig. 4), given by dfC ½¼ ðlog NfB þ 0:3Þ=ðlog NfA þ 0:3Þ or
ðlog Nf 1 þ 0:3Þ=ðlog Nf 2 þ 0:3Þ if r1max > r2max for the current experimental data
from two different applied peak stress levels, was calculated to determine (dfb 
dfC). To find a valid n, a value starting from 1 was incremented with an interval of
0.1 until all the values of (dfb – dfC) at various stress levels become positive to satisfy
the conditions (Eq 12). A pair of two stress levels for respective Points A (or C) and
B (or b) (Fig. 4) with an interval of 1 MPa was used to calculate each value of (dfb 
dfC) for a stress range from 350 MPa ([for Point A in Fig. 4] paired with 351 MPa
[for Point B in Fig. 4]) to 446 MPa (paired with 447 MPa [< ultimate strength]). As
a result, n was found to be 3.4. The reason for the value of 1 MPa chosen for the
stress pair interval is that it is practically small enough—given that the smaller the
interval the higher the resolution of numerical calculations. Also, the range from
350 to 446 MPa was chosen because of a high curvature (or nonlinearity) of the
S-N curve occurs within this range (see also discussion with Fig. 7), but the lower
limit of the range can be extended if the curvature change is not well-defined.
The effect of n value on damage at Points C and b (Fig. 7) calculated using
Eqs 13 and 14 at an applied peak stress level of 200 MPa is shown in Fig. 10. It is
seen (Fig. 10a) that, as n increases, the general location factor (df) increases for all
the curves at a given damage. It is obvious that damage at n ¼ 1 is invalid, damage
at n ¼ 2 and 3 is partially valid (n ¼ 3 is not clearly seen on Fig. 10), and that damage
is finally fully valid at n ¼ 3.4, given that all the points of the lower curve in Fig. 7
should be at the right-hand side of the thick-lined curve. Also, the effect of n on iso-
damage lines on an S-N plane is given in Fig. 10b. It is noted that the slope of the
valid isodamage line slightly increases at a constant stress level, satisfying the com-
patibility condition.
 
The predictions for Nf 2 with DN2 ¼ Nf 2  N2 were conducted with:
 1=n
Df 1
df 2 ¼ df 1 (15)
Df 2

where:
Subscripts 1, 2 ¼ first and second in the loading sequence,
df 1 ¼ ðlog N1 þ 0:3Þ=ðlog Nf 1 þ 0:3Þ, and
df 2 ¼ ðlog N2 þ 0:3Þ=ðlog Nf 2 þ 0:3Þ:
Eq 15 was obtained from Eq 13b and c.
PREDICTION RESULTS AND DISCUSSION
Fig. 11a and b show the predicted number of cycles (Nf 2) for the second stress level
(r2max) with respective high-low and low-high sequence loading paths indicated
by lines between data points, and Fig. 11c shows those from both high-low and
ESKANDARI AND KIM, DOI 10.1520/STP159820150099 319

FIG. 10 Effect of n: (a) on damage at point C of the S-N curve and other points when
point C moves along the horizontal line DA at 200 MPa (see Fig. 4); the curves
with n ¼ 1, 2, and 3 are partially/fully on left-hand side of the thick-lined curve,
representing damage of point B on the S-N curve and (b) on isodamage lines.

0.6

0.5

0.4 Point B on
Damage (D)

S-N curve
0.3
n=1
0.2 n=2
n=3
0.1 n=3.4

0
0 0.2 0.4 0.6 0.8 1 1.2
df

(a)

500

450
n=1

400 n=2
n=3
smax

350 n=3.4

300

250

200
–0.3 0.7 1.7 2.7 3.7 4.7 5.7 6.7
Log N

(b)

low-high sequence loadings without loading paths. The accuracy of prediction


(Accu) was obtained from: Accu ¼ [(log Nf 2 from experimental DN2 with df 2) 
(log Nf 2 from S-N curve)]/(log Nf 2 from S-N curve) and given in Table 1. The dashed
lines in Fig. 11 represent isodamage lines calculated for some chosen points.
320 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 11 Predictions of failure points with experimental results: (a) high-low sequence
loading; (b) low-high sequence loading—two of the data points for premature
failure before the second stress level cycling are denoted by the symbol ‘^’; and
(c) both combined for failure points after second stress level without loading
sequence paths. The dotted lines on both sides of the S-N curve (solid line)
represent failure probabilities, 10 % and 90 %.

500 500

450 450

400 400

350 350
300 300

smax
smax

250 250
200 200
150 150
100 100
50 50
0 0
–0.3 0.7 1.7 2.7 3.7 4.7 5.7 6.7 7.7 –0.3 0.7 1.7 2.7 3.7 4.7 5.7 6.7 7.7
Log N Log N

(a) (b)

High-Low
500
Low-High
450
400
350
300
smax

250
200
150
100
50
0
–0.3 0.7 1.7 2.7 3.7 4.7 5.7 6.7 7.7
Log N
(c)

The load sequence paths closely follow the isodamage lines as calculated. In general,
a majority of experimental data points (Fig. 11c) for Nf 2 appear to be well-predicted.
A close agreement is found between the S-N curve and prediction, particularly for
the high-low loading (Fig. 11a); all the experimental data points at failure predicted
are within, or sufficiently close to, the scatter band represented by dotted lines of
10 % and 90 % probability failures. On the other hand, data from low-high
sequence loading (Fig. 11c) appear scattered a little to one side of the S-N curve with
respect to those of the high-low sequence loading. It may be interesting to under-
stand the difference in statistical behavior between the two types of loading sequen-
ces. Note that an isodamage line between two stress levels for high-low sequence
loading never intersects the S-N curve before failure. In low-high sequence loading,
ESKANDARI AND KIM, DOI 10.1520/STP159820150099 321

however, it may tend to have a higher probability of intersecting the S-N curve be-
fore failure at a higher number of cycles, resulting in a lopsided distribution of fail-
ure points (Nf 2) with some deviation from the mean if the second loading is close
to the S-N curve because of biased sampling. Accordingly, a scatter of Nf 2 (Fig. 11b)
to one side may be expected when individual data points are examined as follows:
(a) many of the data points for the number of cycles (N1) at the first stress level are
already close to the failure points or within the scatter band [indicated by boldfaced
ones in Table 1]; (b) some specimens indicated by a large diamond symbol (^) for
two data points (Fig. 11b) were already broken even before cycling at the second level
stress as also indicated as N/A in Table 1; and (c) there are three other data points
that are already within the scatter band at the first stress level but that survived for
the second stress level cycling. Of course, one might consider, at the same time, the
possibility of load sequence effect for the distribution of Nf 2 points apart from the
scatter as reported in the literature [9,11]. However, such an effect has not been
quantitatively understood yet because the differences found due to the load sequen-
ces in the past were described in terms of cyclic ratio (i.e., N1/Nf1 = N2/Nf 2), which
is not based on the valid damage. Nonetheless, if there would still be an experimen-
tal difference in well-chosen sampling between two different loading sequences,
some effect of load sequence on the prediction of Nf 2 may be possible.

An Evaluative Review on Damage Models for


Fatigue Life Predictions
As demonstrated earlier, life predictions can be conducted by finding equivalent
damage points forming isodamage lines on an S-N plane. Once damage is quantita-
tively defined, equivalent damage points can be found at various numbers of cycles
with corresponding stress levels. Damage can be quantified in different ways to con-
struct the isodamage lines, but this should be a function of (at least) applied stress
level and number of cycles and must be validated. It may be important to note that
the validity of the damage function largely depends on the validity of isodamage
lines. Theoretical validation of isodamage points or damage function, or both, had
not been possible in the past because experimental data verifications for particular
cases had been undertaken in the absence of the compatibility concept. For a gener-
al understanding of why conventional damage models are not reliable, selected
models will be reviewed and theoretically evaluated in the following section using
criteria based on the current validity framework theory we developed.

FATIGUE DAMAGE
The inception of damage modeling by Palmgren [3] dates back to 1924 for the fol-
lowing fatigue damage equation:

DN1 DN2 DN3 DNn X Nn


þ þ þ ::: ¼ 1 or ¼1 (16)
Nf 1 Nf 2 Nf 3 Nfn Nfn
322 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

where:
DNn ¼ Nn  Nn1 , which is the number of cycles added at the nth applied peak
stress rmax;n , and
Nfn ¼ the number of cycles to failure at the nth applied peak stress rmax;n :

Miner [4] attempted to derive Eq 16 using an idea expressed in Eq 17:

w1 DN1 w2 DN2 w3 DN3


¼ ; ¼ ; ¼ ; (17)
W Nf W Nf W Nf

where:
W ¼ “net work absorbed at failure” or “work done to failure,” and
wn ¼ the work done due to DNn cycles.

According to this idea, if W is assumed to be independent of rmax;n and stress ratio


(R), the total work to failure becomes W ¼ w1 þ w2 þ w3    wn ¼ W1 ¼
W2 ¼ Wn and Eq 18:
w1 w2 w3 wn
þ þ þ  ¼1 (18)
W W W W
or Eq 19:
w1 w2 w3 wn
þ þ þ  ¼1 (19)
W1 W2 W3 Wn

at failure although the assumption is not generally true. An isodamage line based on
the Palmgren/Miner model [3,4] is obtained:
Nfi
Ni ¼ Ni1 (20)
Nfi1
where subscript i indicates the sequence in recursion process (i.e., i is the current stage
and i-1 is the previous stage). Isodamage lines at arbitrarily chosen stress levels and
initial numbers of cycles with a constant stress ratio (R) according to Eq 20 are shown
in Fig. 12. The slopes of isodamage line at failure (Fig. 12) are equal to those of the S-N
curve, violating the compatibility between rmax and log N. Also, the boundary condi-
tion (i.e., D ¼ 1 at N ¼ Nf) does not satisfy Eq 1. Another violation is that three differ-
ent points on the three isodamage lines at a constant applied stress have the same
damage value given that damage can be regarded as zero at N ¼ 0.5 cycle. Many
researchers have used or modified Eq 16 without much success as expected.
In 1954, Marco and Starkey [5] revisited the fatigue damage, or fraction of dam-
age (N=Nf ), proposed by Palmgren [3] and addressed the following. They raised the
conceptual problem of “damage” by stating that fatigue damage is a difficult term to
define quantitatively. Also, they paid attention to the linearity of Eq 16 and suggested
an exponential relationship with experimental data obtained at a stress ratio of 1:
 
DNi ai
D¼ (21)
Nfi
ESKANDARI AND KIM, DOI 10.1520/STP159820150099 323

FIG. 12 Isodamage lines based on the Palmgren/Miner model [3,4].

500

450 Palmgren/Miner

400
smax

350

300

250

200
–0.3 0.7 1.7 2.7 3.7 4.7 5.7 6.7
Log N

where exponent ai (>1) is meant to be a function of stress amplitude (ra ) applied, but
a rational method of determination was not proposed. They claimed that the curves in
Fig. 13 representing Eq 21 had been successfully used for predicting test results. It is
noted, however, that damage at N ¼ 0 and N ¼ Nf is still equal to that of Palmgren/
Miner [3,4], the condition (D ¼ 1 at N ¼ Nf) of which does not represent the valid

FIG. 13 Graphical presentation of linear (a ¼ 1) and nonlinear accumulation damage


models.

α=0.3
Damage (D)

α=0.5
α=1
α=2

α=4

0
0 1
Cycle Ratio (ΔN/Nf)
324 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

damage at different stress levels (c.f. Hashin and Rotem’s model [34] which is in the
same general form as Eq 21).
Researchers had not settled on a definition of the fatigue damage concept even
before they quantified it. In 1961, Gatts [35] started using a differential equation
(Eq 22) to describe the damage and to state that damage (D) is a function of load
(F) in relation with residual strength (rR ):

drR
¼ CD (22)
dN

and acknowledged the formidable problem of finding an explicit damage function. In


contrast, Subramanyan [9] simplistically defined the damage as “the slope of an iso-
damage line to the limiting value of the slope of the semi-log S-N curve” to improve
the predictability of the Palmgren/Miner [3,4] isodamage line. Hashin and Rotem [34]
stated that they avoided quantification of damage, replacing “damage” with “residual
life” and then introducing a “damage curve.” They further stated, “A very important
property of damage curves is uniqueness…” and described the damage curve as the
same functional form as the S-N curve. Post, Case, and Lesko [20] described the dam-
age (index) in their review as “a quantity that does not have a bearing on the durabili-
ty properties of a material outside of the model which generated it.” Despite the
conceptual problem of damage, many researchers, including Henry and Ohio [36],
Dubuc et al. [37], Hashin and Rotem [34], and Hashin [38], followed Palmgren/Miner
[3,4] for damage quantities at boundaries (i.e., D ¼ 0 at N ¼ 0 and D ¼ 1 at N ¼ Nf) of
Eq 16. Hashin and Rotem [34] even stated that, for theory development, “The word
‘damage’ should not be taken literally.” Hwang and Han [13] also followed Miner [4]
regarding damage quantities at boundaries and proposed three more damage models.
A damage function involving an endurance limit (r1 ), proposed by Henry and
Ohio [36], is given by:

N
Nf
D¼  : (23)
1 N
1þ 1
Dr  r1 Nf
r1

This equation requires not only an endurance limit (r1 ), but it also adopts an incor-
rect boundary condition at failure (i.e., D ¼ 1). Dubuc et al. [37] proposed a similar
model, claiming a “unified theory” with an additional parameter n assumed to be 8:

N
Nf
D¼    : (24)
rmax rmax ruT n
  
N N r1 r1 r1
þ 1 rmax
Nf Nf 1
r1

The boundary condition at failure of this equation is the same as that of Eq 23 or that
of Palmgren/Miner [3,4]. Fig. 14 shows isodamage lines based on Eq 23 [36] indicating
ESKANDARI AND KIM, DOI 10.1520/STP159820150099 325

FIG. 14 Isodamage lines from Henry and Ohio [36] (see Eq 23) and Hashin and Rotem
[34] (see Eq 25) violating the compatibility. The slopes of isodamage lines
increase at a constant stress level as logN increases and, as a result, three
different damage quantities of three different points at a low number of
cycles (N) simultaneously represent the same damage quantity for one point
of a high N.

500

Hashin and Rotem (1978)


450
Henry and Ohio (1955)

400
smax(MPa)

350

300

250

200
–0.3 0.7 1.7 2.7 3.7 4.7 5.7 6.7
Log N

that they tend to converge on a point due to the nominated endurance limit. Also, not
much difference is found in the isodamage line profiles between Eqs 23 and 24 if they
are plotted.
A two-stage loading prediction model having the same general form as that of
Marco and Starkey’s model [5], which is based on the so-called “damage curve”
proposed by Hashin and Rotem [34], is given by:

 logðNf 2 =N1 Þ
DN1 logðNf 1 =N1 Þ DN2
þ ¼1 (25)
Nf 1 Nf 2

where:
N1 ¼ the number of cycles at a nominated endurance limit, and
DN2 ¼ Nf 2  N2 :
326 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

Hashin and Rotem used a simplified S-N curve, assuming that an endurance
limit exits. Hashin [38] further discussed a remaining (or residual) life theory using
the Palmgren/Miner [3,4] damage boundary condition and derived Eq 25, again
concluding that the damage function “unifies” the residual life and residual strength
methods. Fig. 14 shows isodamage lines based on Eq 25 in comparison with those of
Henry and Ohio [36]. The S-N curve was obtained from experimental data of
Broutman and Sahu [6] with a chosen value of 106 cycles for N1 in Eq 25. The iso-
damage lines based on Hashin and Rotem’s model [34] (Fig. 14) appear to be similar
to those of Henry and Ohio [36] or Dubuc et al. [37] who similarly claimed a
“unified theory.” In fact, all their models violate the compatibility requirements.
The reason for the violation is that three different damage quantities on three dif-
ferent isodamage lines at a low number of cycles (N) tend to simultaneously repre-
sent one damage value, given that the three isodamage lines are originated from the
same zero damage at N ¼ 0.5 cycle. Any damage model with a boundary condition
of 1.0 at failure commonly violates the compatibility conditions in a similar way be-
cause isodamage lines tend to converge on one point or no damage difference is
found on the S-N curve.
In a comparative study of various damage models for experimental specimens with
a hole as the stress raiser, Sarkani et al. [39] defined the damage (D) as given in Eq 26:
ruT  rR
D¼ (26)
ruT  Dr

where:
Dr ¼ stress range,5
and
 C
N
rR ¼ ðruT ÞA  ½ðruT ÞA  Dr : (27)
Nf

It is noted that, when a specimen has a hole, damage is localized due to the stress con-
centration around the hole so that any damage model developed for unnotched speci-
mens is not directly applicable for the fatigue life predictions of specimens with varied
geometry. At the same time, the boundary conditions are the same as those of Palm-
gren/Miner [3,4] (i.e. D ¼ 1 at N ¼ Nf ). Eq 27, however, represents a generalized
form of equation for other damage models, including those proposed by Broutman
and Sahu [6], Hashin [38], and Schaff and Davidson [40].
Broutman and Sahu [6] measured the residual strengths for unnotched fatigue
specimens made of cross-plied E glass fiber/epoxy laminate after being subjected to
fatigue loading at stress ratio (R) ¼ 0.05 as already shown in Fig. 11. For the formula-
tion, they assumed that the residual strength decreases linearly with an increasing

5
It was called “amplitude” in the original paper, but it should be stress range (which equals peak stress
minus valley stress).
ESKANDARI AND KIM, DOI 10.1520/STP159820150099 327

fraction of fatigue life DNðiÞ =Nf ðiÞ , ignoring its nonlinearity, which is the case in
Eq 27 where A ¼ C ¼ 1 for R ¼ 0 or in Eq 28:
NðiÞ
rRðiÞ ¼ ruTði1Þ  ðruT  ri Þ (28)
Nf ðiÞ

where:
rRðiÞ ¼ (residual) strength of the material after cyclic load at ith stress level,
ruTði1Þ ¼ strength before being cycled at ri , and
ruT ¼ ultimate tensile strength of virgin composite.

Thus, when two stress levels are applied in sequence, the residual strength (rRð1Þ ) is
found to be rRð1Þ ¼ ruT  ðruT  r1 ÞNð1Þ =Nf ð1Þ for the first stress level; the residual
strength (rRð2Þ ) at the second stress level (r2 ) is found to be rRð2Þ ¼ r2 ¼
rRð1Þ  ðruT  r2 ÞDNð2Þ =Nf ð2Þ . From these two equations, the remaining fatigue life
(DNð2Þ ) at r2 after being subjected to the first stress r1 is found in Eq 29:
ruT  r1 DNð1Þ DNð2Þ
 ¼ 1: (29)
ruT  r2 Nf ð1Þ Nf ð2Þ

Broutman and Sahu [6] compared these predictions with experimental results
using values calculated from Nð1Þ =Nf ð1Þ þ DNð2Þ =Nf ð2Þ and found that differences
between experimental data and predictions ranged from 0 % to 63 %. A set of iso-
damage points (Ni) from Broutman and Sahu’s model [6] at various stress levels,
with a constant stress ratio determined by equating residual strengths at different
stress levels for an arbitrary initial number of cycles, can be obtained from Eq 30:
ruT  ðruT  ri1 ÞNði1Þ ruT  ðruT  ri ÞNðiÞ
rRði1Þ ¼ ¼ (30)
Nf ði1Þ Nf ðiÞ

yielding (Eq 31):


ðruT  ri1 ÞNði1Þ
NðiÞ ¼ Nf ðiÞ : (31)
ðruT  ri ÞNf ði1Þ

Once initial ri1 and Ni1 are given, an isodamage line (or isoresidual strength
line) can be plotted, as shown in Fig. 15. Isodamage line slopes (j@rmax =@ log N j)
tend to slightly increase with increasing logN, violating the compatibility (i.e., it
should decrease). Another violation is that zero damage at N ¼ 0.5 cycle is equal to
non-zero damage at a low stress level. For example, damage for a point, represented
by an unfilled dot in Fig. 15, at a low cycle is obviously not zero, but simultaneously
it represents zero damage at a point represented by a filled dot. Experimental verifi-
cations do not reveal such general isodamage line characteristics when experiments
are conducted under particular conditions.
A modified version of Broutman and Sahu’s model [6] proposed by Schaff and
Davidson [40] for nonlinearity due to an exponent k is given as:
  !1=k
ðruT  ri1 Þ Nði1Þ k  k
NðiÞ ¼ Nf ðiÞ : (32)
ðruT  ri Þ Nf ði1Þ
328 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 15 Isodamage lines (dash-dot) represent Broutman and Sahu [6], dotted lines are
those modified by Schaff and Davidson [40] with k ¼ 0.7, with solid lines
validated by the current theory.

500

450 Broutman and Sahu


Schaff and Davidson
400 Validated

350
smax

300

250

200

150
–0.3 0.7 1.7 2.7 3.7 4.7 5.7 6.7
Log N

Isodamage lines based on Eq 32 are shown in Fig. 15 in comparison with those from
Broutman and Sahu’s model [6] and also with validated ones (Eq 15). The validity
of Eq 32 with a value of 0.7 for k appears worse than that of Broutman and Sahu’s model
[6] because the isodamage lines are multivalued functions in a region of low values of logN.
Poursartip, Ashby, and Beaumont [41] quantified damage (D) as given in Eq 33
for laminates based on delamination area:
E
D ¼ 2:8571  (33)
E0

where:
E ¼ reduced elastic modulus due to damage and
E0 ¼ elastic modulus of undamaged specimen and used it to find an empirical
damage rate at R ¼ 0.1 given by Eq 34:
 6:393
dD Dr
¼ 9:189  105 : (34)
dN ruT

Subsequently Poursartip and Beaumont [42] determined damage at failure (Df)


given in Eq 35:
rmax
Df ¼ 2:8571  (35)
ruT

and empirically found an S-N curve as a function of stress ratio (R).


ESKANDARI AND KIM, DOI 10.1520/STP159820150099 329

Kim and Zhang [32] derived a damage at failure (Df), Eq 36, with experimental
data for unidirectional fiber-reinforced composites:
rmax
Df ¼ 1  (36)
ruT
and assumed a damage rate at failure given in Eq 37:
dDf
¼ aðrmax Þb (37)
dN
where:
a, b ¼ fitting parameters
and used it to derive an S-N curve given in Eq 3. The quantification of damage
at failure (Df) is regarded as being relatively easy compared to damage (D) prior to
failure. It may be important, however, to note that the damage at failure proposed
by Poursartip, Ashby, and Beaumont [41] or Kim and Zhang [32] is different from
those used by many researchers discussed earlier (i.e., Df ¼ 1 adopted from Palm-
gren/Miner’s model [3,4]), but it is valid for satisfying Eq 1.
Mesmacque et al. [15] proposed a “damage indicator” (Di) that may be distin-
guished from the continuous damage function (D) and is defined as Eq 38:
red1  Dr1
Di ¼ (38)
ruT  Dr1
to indicate damage of Point a in terms of red1 (see Fig. 16). The damage decreases as
the cycling increases, which may not represent the physical damage. Nonetheless, this
may be used to find an equivalent damage (requiv ) at the second stress amplitude Dr2
by equating with Eq 39:
requiv  Dr2
Di ¼ (39)
ruT  Dr2

FIG. 16 Schematic for damage indicator.

σuT

σed1

Δσ1
a

Log N
330 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

giving Eq 40:

red1  Dr1
requiv ¼ ðruT  Dr2 Þ þ Dr2 (40)
ruT  Dr1

for finding the number of cycles at the next Dr2 . Unfortunately, the damage indicator
in this model is constant and is similar to Palmgren/Miner’s model [3,4] and violating
the compatibility.

RESIDUAL STRENGTH
The prediction of residual strength following cyclic loading may be useful as a step
prior to predicting the remaining fatigue lives at different stress levels.
Owen and Howe [43] measured the residual strength of unnotched glass-
reinforced composites. The trend of residual strength as a function of life fraction was shown
to be different from that obtained by Broutman and Sahu [6]. They fitted data for the resid-
ual strength divided by the ultimate strength (ruT ) as a function of N=Nf using Eq 41:
 2
N N
rRð1Þ ð%Þ ¼ 100  30:89 þ 17:60 : (41)
Nf Nf

They stated that the residual strength results were substantially independent of “stress
condition” (which is stress ratio). On the basis of experimental results and this equa-
tion, they defined damage (L0 ) as Eq 42:
100  rRð1Þ
L0 ¼ (42)
13:29

to have L 0 ¼ 1 at N=Nf ¼ 1 and L0 ¼ 0 at N=Nf ¼ 0, which does not satisfy Eq 1. Ac-


cordingly, the damage (L0 ) due to the residual strength becomes Eq 43:
   2
N N
L0 ¼ a b : (43)
Nf Nf

They also measured a total resin cracking length and further described the normalized
total resin cracking length as another version of damage (L) given in Eq 44:
   2
N N
L ¼ a0  b0 (44)
Nf Nf

to show a similarity between L0 and L, which is different from Palmgren/Miner’s [3,4]


linearity. Based on this, for multi-stress levels at failure, they suggested Eq 45:
"    2 #
X Ni Ni
1¼ a b : (45)
Nfi Nfi

Predictability of remaining fatigue life for this equation at multi-stress levels, however,
was not demonstrated by the authors.
ESKANDARI AND KIM, DOI 10.1520/STP159820150099 331

Bond and Farrow [44] adopted a modified version of Owen and Howe’s model
given in Eq 46 [43] for applications of special geometry of specimens with a coun-
tersunk hole subjected to spectrum fatigue loading:
"  d #
X  Ni  Ni
0
L ¼ a b (46)
Nfi Nfi

They found their predictions of fatigue life for angle ply laminates may be reasonable
but overly conservative for quasi-isotropic laminates. If they used unnotched speci-
mens and simple loading cases, the applicability of Owen and Howe’s model [34]
would have been better known.
Yang and Liu [45] derived a residual strength model (Eq 47) using the Weibull
distribution:

rCR ¼ rCuT  KbC ðDrÞb N (47)

where:
C ¼ fitting parameter, and
Dr ¼ stress range,
assuming that an S-N curve as KðDrÞb Nf ¼ 1, which does not fully fit fatigue
data at low cycles. Yang and Jones [46] further developed this to have Eq 48:

rAuT  rAmax
rAR ¼ rAuT    KðDrÞb N (48)
rCuT  rCmax

where A and B are to reflect the rate of degradation of the residual strength requiring
experiments, and C is determined from constant amplitude fatigue life data. Yang and
Jones [47] showed a different version of the residual strength model:
 B
N rAuT  rAmax
rAR ¼ rAuT   B : (49)
Nf rCuT  rCmax

Yang and Du [48] showed Eq 50 as a simplified form for the reason that Eq 49 is “too
cumbersome for practical applications”:
 B
N ruT  rmax
rR ¼ ruT   B : (50)
Nf rCuT  rCmax

They further simplified this with B  1 because of a lack of test data for experimental
verification.
Philippidis and Passipoularidis [19] reviewed and experimentally evaluated re-
sidual strength models including those discussed earlier [6,40,45,46]. They summa-
rized their findings that no model is consistent in accurately predicting residual
strength degradation behavior of various composite laminates. An example for
comparing the predictions from various models and experimental data is given in
Fig. 17. The designations for models in Fig. 17 are: BR for Broutman and Sahu [6],
332 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 17 Comparison of predictions and experimental data conducted by Philippidis and


Passipoularidis [19]. Designations: BR ¼ Broutman and Sahu [6]; INT ¼ Adam
et al. [49]; H ¼ Hahn and Kim [8]; REI and OM ¼ Schaff and Davidson [40] from
two different values of fitting parameter; and W1 and W2 ¼ Sendeckyj [50]
from two different values of fitting parameter, Y1 [45] and Y2 [46].

780 BR
730 INT
Residual Strength (MPa)

680
H
630
OM
580
REI
530
480 W1
430 W2
380 Y1
330 Y2
280
0 50000 100000 150000 200000
N

INT for Adam et al. [49], H for Hahn and Kim [8], REI and OM for Schaff and
Davidson [40] from different values of fitting parameters, W1 and W2 for Sendeckyj
[50] from two different parameters, Y1 for Yang and Liu [45], and Y2 for Yang and
Jones [46]. The nonlinear model [40] or a modified BR model by Schaff and David-
son, labeled as OM, appears to be better than others for residual strengths. However,
as already shown in Fig. 15, the compatibility on the S-N plane for life predictions is
not better than the linear model of Broutman and Sahu [6]. Philippidis and Passipou-
laridis [19] concluded that complicated phenomenological models requiring large
experimental data sets do not necessarily pay back in terms of predictions, and they
regarded Broutman and Sahu’s model [6] favorably for its simplicity.

Conclusion
A theoretical framework for fatigue damage on the S-N plane has been developed,
consisting of damage axioms, boundary conditions, and compatibility conditions,
and requiring the reference damage equation. A damage function satisfying all the
requirements has been derived for a stress ratio of zero. It can be used for theoreti-
cal predictions of remaining fatigue life at any stress level with an S-N curve model
(Kim and Zhang) but without any other experimental fitting parameters. The pre-
dictions based on the developed damage function have been found to be in close
agreement with experimental results from two stress level sequence loading.
ESKANDARI AND KIM, DOI 10.1520/STP159820150099 333

With the benefits of the new compatibility concept and criteria, an evaluative
review on the conventional fatigue damage models has been presented for a general
understanding of defective approaches taken since the early twentieth century.

References

[1] Gerber, W. Z., “Bestimmung der zulassigen Spannungen in Eisen-Constructionen”


[Calculation of the Allowable Stresses in Iron Structures], Z Bayer Archit Ing Ver, Vol. 6,
No. 6, 1874, pp. 101–110.

[2] Schütz, W., “A History of Fatigue,” Eng Fract Mech, Vol. 54, No. 2, 1996, pp. 263–300.

[3] Palmgren, A. G., “Die Lebensdauer von Kugellagern” [Life Length of Roller Bearings],
Zeitschrift des Vereines Deutscher Ingenieure (VDI Zeitschrift), Vol. 68, No. 14, 1924,
pp. 339–341.

[4] Miner, M. A., “Cumulative Damage in Fatigue,” J Appl Mech, Vol. 12, No. 3, 1945,
pp. 159–164.

[5] Marco, S. M. and Starkey, W. L., “A Concept of Fatigue Damage,” Transaction ASME,
Vol. 76, 1954, pp. 627–632.

[6] Broutman, L. J. and Sahu, S., “A New Theory to Predict Cumulative Fatigue Damage in
Fiberglass Reinforced Plastics,” Composite Materials: Testing and Design (Second
Conference), ASTM STP497, ASTM International, West Conshohocken, PA, 1972, pp. 170–188,
http://dx.doi.org/10.1520/STP27746S

[7] Rotem, A., “Residual Strength after Fatigue Loading,” Int J Fatigue, Vol. 10, No. 1, 1988,
pp. 27–31.

[8] Hahn, H. T. and Kim, R. Y., “Fatigue Behavior of Composite Laminates,” J Compos Mater,
Vol. 10, No. 2, 1976, pp. 156–180.

[9] Subramanyan, S. A., “Cumulative Damage Rule Based on the Knee Point of the S-N
Curve,” J Eng Mater-T ASME, Vol. 98, No. 4, 1976, pp. 316–321.

[10] Shokrieh, M. M. and Lessard, L. B., “Multiaxial Fatigue Behavior of Unidirectional Plies
Based on Uniaxial Fatigue Experiments — I. Modelling,” Int J Fatigue, Vol. 19, No. 3, 1997,
pp. 201–207.

[11] Gamstedt, E. K. and Sjögren, B. A., “An Experimental Investigation of the Sequence
Effect in Block Amplitude Loading of Cross-Ply Composite Laminates,” Int J Fatigue,
Vol. 24, Nos. 2–4, 2002, pp. 437–446.

[12] Epaarachchi, J. A. and Clausen, P. D., “An Empirical Model for Fatigue Behavior Predic-
tion of Glass Fiber-Reinforced Plastic Composites for Various Stress Ratios and Test
Frequencies,” Compos Part A-Appl S, Vol. 34, No. 4, 2003, pp. 313–326.

[13] Hwang, W. and Han, K. S., “Cumulative Damage Models and Multi-Stress Fatigue Life
Prediction,” J Compos Mater, Vol. 20, No. 2, 1986, pp. 125–153.

[14] Fatemi, A. and Yang, L., “Cumulative Fatigue Damage and Life Prediction Theories: A
Survey of the State of the Art for Homogeneous Materials,” Int J Fatigue, Vol. 20, No. 1,
1998, pp. 9–34.
334 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

[15] Mesmacque, G., Garcia, S., Amrouche, A., and Rubio-Gonzalez, C., “Sequential Law in Mul-
tiaxial Fatigue, a New Damage Indicator,” Int J Fatigue, Vol. 27, No. 4, 2005, pp. 461–467.

[16] Aghoury, I. E. and Galal, K., “A Fatigue Stress-Life Damage Accumulation Model for Vari-
able Amplitude Fatigue Loading Based on Virtual Target Life,” Eng Struct, Vol. 52, 2013,
pp. 621–628.

[17] Mao, W., Ringsberg, J. W., and Rychlik, I., “Theoretical Development and Validation of a
Fatigue Model for Ship Routing,” Ships Offshore Struc, Vol. 7, No. 4, 2011, pp. 399–415.

[18] Sutherland, H. J. and Mandell, J. F., “The Effect of Mean Stress on Damage Predictions
for Spectral Loading of Fiberglass Composite Coupons,” Wind Energy, Vol. 8, No. 1,
2004, pp. 93–108.

[19] Philippidis, T. P. and Passipoularidis, V. A., “Residual Strength after Fatigue in Compo-
sites: Theory vs. Experiment,” Int J Fatigue, Vol. 29, No. 12, 2007, pp. 2104–2116.

[20] Post, N. L., Case, S. W., and Lesko, J. J., “Modeling the Variable Amplitude Fatigue of
Composite Materials: A Review and Evaluation of the State of the Art for Spectrum
Loading,” Int J Fatigue, Vol. 30, No. 12, 2008, pp. 2064–2086.

[21] D’Amore, A., Caprino, G., Stupak, P., Zhou, J., and Nicolais, L., “Effect of Stress Ratio on
the Flexural Fatigue Behaviour of Continuous Strand Mat Reinforced Plastics,” Sci Eng
Compos Mater, Vol. 5, No. 1, 1996, pp. 1–8.

[22] Epaarachchi, J. A., “Effects of Static–Fatigue (Tension) on the Tension–Tension Fatigue


Life of Glass Fiber Reinforced Plastic Composites,” Compos Struct, Vol. 74, No. 4, 2006,
pp. 419–425.

[23] El Kadi, H. and F., Ellyin, “Effect of Stress Ratio on the Fatigue of Unidirectional Glass
Fiber/Epoxy Composite Laminae,” Composites, Vol. 25, No. 10, 1994, pp. 917–924.

[24] Kawai, M., “A Phenomenological Model for Off-Axis Fatigue Behavior of Unidirectional
Polymer Matrix Composites under Different Stress Ratios,” Compos Part A-Appl S,
Vol. 35, Nos. 7–8, 2004, pp. 955–963.

[25] Kawai, M. and Koizumi, M., “Nonlinear Constant Fatigue Life Diagrams for Carbon/Epoxy
Laminates at Room Temperature,” Compos Part A-Appl S, Vol. 38, No. 11, 2007,
pp. 2342–2353.

[26] Kawai, M. and Itoh, N., “A Failure-Mode Based Anisomorphic Constant Life Diagram for
a Unidirectional Carbon/Epoxy Laminate under Off-Axis Fatigue Loading at Room Tem-
perature,” J Compos Mater, Vol. 48, No. 5, 2014, pp. 571–592.

[27] Kim, H. S., 2013, “Mechanics of Solids and Fracture,” Ventus Publishing ApS, Frederiks-
berg, Denmark, 2013.

[28] Haery, H. A. and Kim, H. S., “Damage in Hybrid Composite Laminates,” JMC, Vol. 1, No. 2,
2013, pp. 127–137.

[29] Kitagawa, H. and Takashima, S., “Application of Fracture Mechanics to Very Small
Fatigue Cracks or the Cracks in the Early Stage,” in Proceedings of the Second Interna-
tional Conference on Mechanical Behavior of Materials, American Society for Metals,
Materials Park, OH, 1976, pp. 627–631.

[30] Bertalanffy, L. V., General System Theory, Braziller, New York, 1968.
ESKANDARI AND KIM, DOI 10.1520/STP159820150099 335

[31] Kim, H. S., “Uncertainty Analysis for Peer Assessment: Oral Presentation Skills for Final
Year Project,” Eur J Eng Ed, Vol. 39, No. 1, 2014, pp. 68–83.

[32] Kim, H. S. and Zhang, J., “Fatigue Damage and Life Prediction of Glass/Vinyl Ester
Composites,” J Reinf Plast Comp, Vol. 20, No. 10, 2001, pp. 834–848.

[33] Harik, V. M., Klinger, J. R., and Bogetti, T. A., “Low-Cycle Fatigue of Unidirectional Com-
posites: Bi-Linear S–N Curves,” Int J Fatigue, Vol. 24, Nos. 2–4, 2002, pp. 455–462.

[34] Hashin, Z. and Rotem, A., “A Cumulative Damage Theory of Fatigue Failure,” Mater Sci
Eng, Vol. 34, No. 2, 1978, pp. 147–160.

[35] Gatts, R. R., “Application of a Cumulative Damage Concept to Fatigue,” J Basic Eng-T
ASME, Vol. 83, No. 4, 1961, pp. 529–534.

[36] Henry, D. L. and Ohio, D., “A Theory of Fatigue Damage Accumulation in Steel,” Trans
ASME, Vol. 77, 1955, pp. 913–918.

[37] Dubuc, J., Thang, B. Q., Bazergui, A., and Biron, A., “Unified Theory of Cumulative Dam-
age in Metal Fatigue,” Weld Res Counc Bull, Vol. 162, 1971, pp. 1–20.

[38] Hashin, Z., “Analysis of Cracked Laminates: A Variational Approach,” Mech Mater, Vol. 4,
No. 2,1985, pp. 121–136.

[39] Sarkani, S., Michaelov, G., Kihl, D. P., and Bonanni, L., “Comparative Study of Nonlinear
Damage Accumulation Models in Stochastic Fatigue of FRP Laminates,” J Struct Eng-
ASCE, Vol. 127, No. 3, 2001, pp. 314–322.

[40] Schaff, J. R. and Davidson, B. D., “Life Prediction Methodology for Composite Structures.
Part I—Constant Amplitude and Two-Stress Level Fatigue,” J Compos Mater, Vol. 31,
No. 2, 1997, pp. 128–156.

[41] Poursartip, A., Ashby, M. F., and Beaumont, P. W. R., “The Fatigue Damage Mechanics of
a Carbon Fiber Composite Laminate: I—Development of the Model,” Compos Sci Tech-
nol, Vol. 25, No. 3, 1986. pp. 193–218.

[42] Poursartip, A., and Beaumont, P. W. R., “The Fatigue Damage Mechanics of a Carbon
Fiber Composite Laminate: II—Life Prediction,” Compos Sci Technol, Vol. 25, No. 4, 1986,
pp. 283–299.

[43] Owen, M. J. and Howe, R. J., “The Accumulation of Damage in a Glass-Reinforced


Plastic under Tensile and Fatigue Loading,” J Phy D Appl Phys, Vol. 5, No. 9, 1972,
pp. 1637–1653.

[44] Bond, I. P. and Farrow, I. R., “Fatigue Life Prediction under Complex Loading for XAS/
914 CFRP Incorporating a Mechanical Fastener,” Int J Fatigue, Vol. 22, No. 8, 2000,
pp. 633–644.

[45] Yang, J. N. and Liu, M. D., “Residual Strength Degradation Model and Theory of Periodic
Proof Tests for Graphite/Epoxy Laminates,” J Compos Mater, Vol. 11, No. 2, 1977,
pp. 176–203.

[46] Yang, J. N. and Jones, D. L. “Load Sequence Effects on the Fatigue of Unnotched Com-
posite Materials,” Fatigue of Fibrous Materials, ASTM STP723, ASTM International West
Conshohocken, PA, 1981, pp. 213–232, http://dx.doi.org/10.1520/STP27622S
336 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

[47] Yang, J. N. and Jones, D. L., “Fatigue of Graphite/Epoxy /0/90/45/-45/S Laminates


under Dual Stress Levels,” J Compos Tech Res, Vol. 4, Vol. 3, 1982, pp. 63–70.

[48] Yang, J. N. and Du, S., “An Exploratory Study into the Fatigue of Composites under
Spectrum Loading,” J Compos Mater, Vol. 17, No. 6, 1983, pp. 511–526.

[49] Adam, T., Dickson, R. F., Jones, C. J., Reiter, H., and Harris, B., “A Power Law Fatigue
Damage Model for Fibre-Reinforced Plastic Laminates,” Proceedings of the Institution
of Mechanical Engineers, Part C: Journal of Mechanical Engineering Science, Vol. 200,
No. 3, 1986, pp. 155–166.

[50] Sendeckyj, G. P., “Fitting Models to Composite Materials Fatigue Data,” Test Methods
and Design Allowables for Fibrous Composites, ASTM STP734, C. C., Chamis, Ed., ASTM
International, West Conshohocken, PA, 1981, pp. 245–260, http://dx.doi.org/10.1520/
STP29314S
FATIGUE AND FRACTURE TEST PLANNING, TEST DATA ACQUISITIONS AND ANALYSIS 337

STP 1598, 2017 / available online at www.astm.org / doi: 10.1520/STP159820160056

Limin Luo,1 Jason Hamilton,1 Zhigang Wei,1 and


Robert Rebandt1

Verification and Validation of


Accelerated Testing Methods for
Fatigue Life Assessment
Citation
Luo, L., Hamilton, J., Wei, Z., and Rebandt, R., “Verification and Validation of Accelerated
Testing Methods for Fatigue Life Assessment,” Fatigue and Fracture Test Planning, Test Data
Acquisitions and Analysis, ASTM STP1598, Z., Wei, K. Nikbin, P. C. McKeighan, and D. G. Harlow,
Eds., ASTM International, West Conshohocken, PA, 2017, pp. 337–358, http://dx.doi.org/
10.1520/STP1598201600562

ABSTRACT
Fatigue-related durability and reliability performance is one of the most
important concerns in many engineering applications, such as for ground
vehicles, aircrafts, and power generation systems. Product designers, validation
engineers, and manufacturers are under continuous and increasing pressure to
reduce the “design cycle” and "time to market" of products while assuring high
levels of durability and reliability of these products. To achieve these goals, a
number of accelerated testing methods have been developed, with some of
them being applied very successfully to engineering applications. Because the
ability to assess the fatigue life of an engineering component or system from the
accelerated testing results is of great importance, the verification and validation
of the accelerated testing methods is becoming critical to properly interpret the
predicted results and decision making. In today’s highly regulated world, there
are many regulations, guidelines, standards, and specifications for conducting
accelerated testing, data analysis, and methods verification. However, methods
are being continuously revised or enhanced by absorbing more advanced

Manuscript received March 1, 2016; accepted for publication October 26, 2016.
1
Tenneco, Inc., 3901 Willis Rd., Grass Lake, MI 49240
2
ASTM Symposium on Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis on May 4–5,
2016 in Grand Hyatt, San Antonio, TX.

Copyright V
C 2017 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.
338 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

developments to address the fundamental issues or specific needs of a given


application, or both. In this paper, two fundamental accelerated testing methods
are reviewed and thoroughly investigated: (1) the test-to-failure method and
(2) degradation testing. Several key issues are identified in this paper with emphasis
on the verification and validation of several new concepts and methodologies. Case
studies are provided to demonstrate the critical issues identified.

Keywords
accelerated testing, linear curve fitting, fatigue, degradation testing

Introduction
Durability and reliability performance are among the most important concerns for
many engineering components and systems, which experience cyclic fatigue load-
ings and may eventually fail over time [1]. Durability and reliability assessment and
associated product validation require effective and robust testing methods [2,3].
Several testing methods are available and, among them, two basic testing methods
are widely used: life testing and degradation testing. There are also two associated
and commonly used accelerated testing methods—accelerated life testing and accel-
erated degradation testing [4–7]. Accelerated testing methods are commonly used
to shorten the development time and reduce the associated cost while not signifi-
cantly sacrificing the accuracy of the assessment. Virtually all laboratory life tests
are accelerated in one manner or another [6]. The life testing (test-to-failure) meth-
od tests a component as to the occurrence of failure under the specified loading.
Degradation testing tests a component to a certain damage level or a specified cycle
so that the final life can be estimated from the information obtained. Degradation
testing can significantly reduce testing time as compared to life testing. To reduce
testing even further, degradation tests can be performed at elevated stresses, called
accelerated degradation testing.
The testing and associated accelerated testing methods should be verified and
validated in order to confidently implement them in practice. To verify a method is
to make sure or demonstrate that the method is reliable, accurate, and justified. To
validate a method is to prove that the method is based on truth, fact, and logic, and
that the final results are acceptable. Generally speaking, for accelerated testing, veri-
fication can be understood as a procedure that follows a standard or specified pro-
cess, and validation can be thought of as a procedure that generates expected and
experimentally correlated results. To achieve the verification and validation goals
for the accelerated testing methods, the following two fundamental aspects must be
properly addressed: (1) test data fitting and (2) data extrapolation. First, accelerated
testing requires the tested data conducted at a higher stress level to be extrapolated
to the lower or service stress level. In order to accurately or reasonably predict the
service life, the accelerated testing must be properly planned, executed, and inter-
preted. Normally, a mean curve or a design curve at a specific reliability or confi-
dence level must be determined with the regression or other data-fitting methods.
LUO ET AL., DOI 10.1520/STP159820160056 339

An inaccurate fit will be amplified with extrapolation and eventually will lead to less
accurate prediction and misleading results. This is particularly true for accelerated
testing with a large ratio of accelerated testing level to service stress level, with
which the extrapolation is unavoidable. Therefore, data fitting is a critically impor-
tant aspect in accelerated testing and it is particularly important for life testing. In
accelerated life testing, the obtained mean curve or design curves obtained from
high stress levels are extrapolated to a lower or service level with a known extrapo-
lation rule, such as linear extrapolation in a log-log plot. Second, how to extrapolate
the tested data obtained at one level to another is not a trivial issue. Normally, the
extrapolation is based on the failure test data, from which a trend can be observed
and an empirical formula can be extrapolated to an untested domain. This
approach usually requires a lot of test data from which the trend is established. In
practice, the number of test samples available are often small in size. Additionally,
simply extrapolating a trend beyond the tested range without understanding the
underlying physical mechanisms is very dangerous. Generally, the farther the accel-
erated stress level is from the normal service stress level, the greater the uncertainty
in the extrapolation. Extrapolation is typically justified on the basis of physically
motivated models; therefore, a fundamental understanding of the relationship
among the extrapolation variables is required [6].
In this paper, these two aspects are critically examined. The first part will focus
on data pattern identification and the associated curve-fitting approaches. The sec-
ond part will introduce a variable transformation technique, which will be the foun-
dation of a degradation testing approach. This paper is organized as follows: The
first section describes the two commonly used accelerated testing methods: acceler-
ated life testing method and the degradation method. The next section describes the
data pattern identification and data fitting for accelerated testing. The following sec-
tion provides a variable transformation technique for degradation (accelerated) test-
ing. Case studies and discussion are then provided. The final section summarizes
the paper and highlights several key observations and recommendations.

Accelerated Testing Methods


ACCELERATED LIFE TESTING
Accelerated fatigue life testing involves the increase of stress/load or temperature.
The increase of temperature could introduce new temperature-related deformation
and failure mechanisms such as creep and oxidation; therefore, it is not as common
as increasing the load. At least two higher stress levels (lower and upper levels) are
introduced to conduct accelerated fatigue life tests. Then, the design parameters at
the service stress level can be estimated from the accelerated testing through simple
extrapolation. When more than one stress level testing is involved, the three-
dimensional load-damage-cycle (S  D  N) diagram, as shown in Fig. 1 [8], can be
used to interpret the accelerated life testing. In Fig. 1, the mutually perpendicular
horizontal and lateral axes represent the applied fatigue cycle N and applied load
340 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 1 The representation of fatigue life population as revealed in the load-damage-


cycle (S  D  N ) diagram.

(or range) level, while the vertical axis represents the damage D. The intersection of
the three axes represents the very beginning of the testing process, at which the
applied cycle N ¼ 0 and damage D ¼ 0. The damage D is always an increasing
function of N because a damage process is often assumed to be an irreversible pro-
cess (i.e., dD=dN  0), so that any possible damage healing processes are ignored.
When the applied cycle N ¼ Nf , the damage D ¼ 1, indicating a complete failure of
the product. The dashed lines shown in Fig. 1 represent the evolution trajectory of
the damage, which can be linear or nonlinear depending on the assumption of the
damage process. Because more than one stress level is involved in an accelerated
life testing process, the load/stress-N curve (S  N), shown in Fig. 2, is often used as
a two-dimensional tool to illustrate the accelerated testing process with the damage
D ¼ 1 at each stress level.

FIG. 2 Schematic of an accelerated life testing procedure in the load/stress-cycle


(S  N) diagram with D ¼ 1.
LUO ET AL., DOI 10.1520/STP159820160056 341

With probabilistic distribution functions, that is, f bNF ðSU Þc and f bNF ðSL Þc, at
the two higher stress levels, SU and SL , the probabilistic distribution f bNF ðSS Þc of the
life at the service stress level SS can be appropriately obtained by extrapolating data
obtained from the higher stress levels, Fig. 2. Probabilistic distribution functions can
be properly used to describe the scatter in fatigue data, which can come from many
sources—such as material uncertainty, loading uncertainty, and other uncertainties.
In many cases, especially for single failure mechanisms and modes, continuous prob-
ability density functions (PDFs) f ðN Þ and cumulative distribution functions (CDFs)
Ð
F ¼ f ðN ÞdN can be effectively used for characterizing the scattered test data. Two-
parameter and three-parameter Weibull or lognormal distribution probabilistic func-
tions are often used to estimate the parameters, that is, mean and standard deviation
for two-parameter lognormal (or normal in a log plot) distribution and shape and
scale parameters for two-parameter Weibull distribution [3]. A threshold or location
parameter is introduced in the three-parameter distributions to model the threshold
physical phenomena, such as crack initiation and propagation.

DEGRADATION (ACCELERATED) TESTING


When life testing takes too long for a product to fail under normal or even acceler-
ated stress conditions, degradation testing method can be used as a complementary
method [4–6]. Degradation data may even yield more accurate life estimates than
life test data, with few failures and yielding better insight into the degradation pro-
cess and how to improve it [6]. A widely cited example of product degradation is
the wear of brake pads, by which the life of the brake pads can be accurately esti-
mated based on the observation by degradation testing made in a relatively short
time. Fatigue crack propagation due to material fatigue is also a widely used exam-
ple [9]. Under a constant load/stress, the three-dimensional load-damage-cycle
(S  D  N) diagram, as shown in Fig. 1, can be reduced to the two-dimensional
damage-cycle (D  N) diagram, Fig. 3. In Fig. 3, two subcases are provided: (1) the

FIG. 3 Schematic of degradation testing procedure in the damage-cycle (D  N)


diagram based on (a) the cycle distribution at a specified damage D and
(b) the damage distribution at a specified cycle Nd.

(a) (b)
342 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

cycle distribution at a specified damage D, Fig. 3a, and (2) the damage distribution
at a specified cycle Nd , Fig. 3b. The lower case d in Fig. 3 stands for degradation. The
scatter and uncertainty of the fatigue behavior can be described by f ½DðN Þ and
F bDðN Þc, which are the PDF and CDF of the damage at a specific applied cycle,
respectively, also shown in Fig. 3.

Data Pattern Identification and Curve Fitting


In a fatigue life testing, the load (stress/strain) is the controlled variable and the life
is the response. In fatigue life data analysis, the load is usually treated as the inde-
pendent variable and fatigue life is treated as the dependent variable. This proce-
dure of fatigue data analysis has been adopted by several standard procedures
[10–12]. However, it is a common practice for an analyst to invert the independent
and dependent variable in data analysis (i.e., doing regression of x on y, assuming
that life is the independent variable and load is the dependent variable). The meth-
od that treats fatigue life as the independent variable and stress/strain as the depen-
dent variable is called the “inverse model,” whereas the opposite method is called
the “non-inverse model” [13]. In the Versailles Project on Advanced Materials and
Standards round-robin test for high cycle fatigue data analysis, six out of seven of
the analyses used stress as the dependent variable [13]. In order to understand how
to choose the independent and dependent variables, an equilibrium-based curve-
fitting method has been developed [14] and it will be briefly described as follows.
The equilibrium mechanism, which is analogous to the equilibrium of force
and angular moment in classic mechanics, has been developed for homoscedastic
[14] and heteroscedastic [15] data. For n given data points ðxi ; yi Þ; i ¼ 1; … n, the
distance of a data point ðxi ; yi Þ to the expected curve y ¼ a þ bx along, say, the ver-
y
tical direction can be expressed as Fi , which is written in terms of “total force.”
Mathematically,
Xn
y
Xn
Fi ¼ ½yi  ða þ bxi Þ ¼ 0: (1)
i¼1 i¼1

The total moment balance, with arms along the x direction of all these data
around a point ðxc ; yc Þ, which could be the centroid of the points system, can be
expressed as
X
n
y
X
n X
n
Mi ¼ f½yi  ða þ bxi Þðxi  xc Þg ¼ f½yi  ða þ bxi Þxi g ¼ 0: (2)
i¼1 i¼1 i¼1

In deriving Eq 2, the terms with Xc are cancelled out because of Eq 1. The final
solution is
P
n P
n P
n P
n P
n P
n P
n
yi xi2  xi xi yi n xi yi  xi yi
i¼1 i¼1 i¼1 i¼1 i¼1 i¼1 i¼1
a¼  n 2 b¼  n 2 : (3)
Pn P Pn P
n xi2  xi n xi2  xi
i¼1 i¼1 i¼1 i¼1
LUO ET AL., DOI 10.1520/STP159820160056 343

Similarly, the equations of the fitted curve can be obtained by applying the
equilibrium concept along another direction, and the solutions are
 n 2
Pn P
n P
n P
n P n P
yi xi yi  xi yi2 n yi2  yi
i¼1 i¼1 i¼1 i¼1 i¼1 i¼1
a¼ b ¼ : (4)
Pn P
n Pn Pn Pn P n
n xi yi  xi yi n xi yi  xi yi
i¼1 i¼1 i¼1 i¼1 i¼1 i¼1

Eq 1, Eq 2, and the final solutions, Eq 3 and Eq 4, are exactly the same as those
obtained from the least squares method [16].
Compared to the existing least squares and maximum likelihood methods [16],
the equilibrium method gives an obvious physical meaning: force equilibrium and
associated data patterns [14,15]. Based on the equilibrium concept, the identifica-
tion of data pattern is treated as an indispensable pre-processing procedure and a
post-processing tool to examine the goodness of fit, especially for data with a large
scatter band. Eventually, two “standard patterns” (i.e., vertical pattern), Fig. 4a, and
horizontal pattern, Fig. 4b, have been identified [14]. Based on the equilibrium
mechanism and for data sets with these patterns, the best-fit curve must be the mid-
dle curve (line) so that the data can be symmetrically distributed around the
expected curve and balanced to make sure that the net “force” and “moment” are
zero as required from the equilibrium principle. Therefore, the best mean curve fit
can be guaranteed for these ideal data patterns. Any deviation from these “standard
patterns” can lead to inaccurate results [14].
It can be demonstrated that, for a given standard data pattern and its corre-
sponding fit curve or surface, any deviation from a standard pattern will result in a
deviated fit curve/surface, which is not accurate and is not desired. Take Fig. 4a, the
“standard vertical pattern,” also shown in Fig. 5a, as an example. If two triangular
data blocks are symmetrically added to the lower and upper bounds of the existing
data set with the standard vertical pattern, Fig. 5b, the fit curve will go down to meet

FIG. 4 Standard patterns for linear data (a) vertical and (b) horizontal [3].

Σ Fi y = 0 Σ Fi x = 0
Σ M iy = 0 Σ M ix = 0

(a) (b)
344 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 5 Equilibrium establishment of (a) vertical pattern, (b) vertical patterned data
with added symmetrical data blocks, and (c) vertical patterned data with added
anti-symmetrical data blocks.

y Σ Fi y = 0 y Σ Fi y = 0 y Σ Fi y = 0
Σ M iy = 0 Σ M iy = 0 Σ M iy = 0

end end F end M

x x x
(a) (b) (c)

the new “force” equilibrium because of the added two blocks of data. The net
“angular moment” can be cancelled out in the case shown in Fig. 5b. If the two
blocks added to the existing standard pattern are antisymmetrical, Fig. 5c, then the
fit curve will rotate around the centroid to a certain degree to establish a new equi-
librium because of the added net “angular moment.” In the case shown in Fig. 5c,
the net “force” contributed from the two blocks is cancelled out. For more general
cases of added data blocks, such as the case with only one triangular block, both
“force” and “moment” will cause the fit curve to make both translation and rotation
movements, which will result in inaccurate fit curves. Therefore, with a certain
equilibrium direction, any deviation in data pattern from a standard pattern will
lead to inaccurate fit curve.
From Fig. 4, it can be seen that, for a given set of data, the vertical and the hori-
zontal offsets methods will result in the same or a similar fit curve as long as the
data patterns are consistent. In other words, the probabilistic distributions along
the vertical and the horizontal directions are related. Because the underlying as-
sumption of the least squares regression method and the equilibrium-based curve-
fitting method is that the probabilistic distribution is a normal distribution [14], it
would be very helpful to reveal the relationship between the vertical offsets fitting
and the horizontal offsets fitting. The following section demonstrates the effort in
this direction.
Generally, if we know a joint probability density function f ðx; yÞ and a distribu-
tion, say, f ð yÞ, then the conditional probability density function for x given y can
be calculated as f ðxj yÞ ¼ f ðx; yÞ=f ð yÞ [17]. The distribution of cycles to failure, say
f ð yÞ ¼ f ½N ðSÞ, can be obtained by fitting test data. However, the joint probability
density function is usually not available. Therefore, the distribution of stress for a
given cycle f ðxÞ ¼ f ½SðN Þ cannot be directly obtained. By employing some engi-
neering mathematics, the pragmatic and heuristic approaches, the relationship
between the two distributions can be indirectly revealed as long as f ½N ðSÞ can be
converted into f ½SðN Þ and in the form of a normal distribution. Now, assume that
LUO ET AL., DOI 10.1520/STP159820160056 345

the distribution of the cycles to failure at a stress level follows a normal distribution
in a log-log plot, Eq 5.
8 h  i2 9
>
< logðNÞ  a  blogðSÞ >
=
1
f ½N ðSÞ ¼ pffiffiffiffiffi exp  (5)
2pr >
: 2r 2 >
;

where the a  blogðSÞ is the mean life logðNÞ. The mean is obtained as
P
logðNÞ ¼ ni¼1 logðNi Þ, and i is the total test numbers.
Based on the data pattern of the linear data, we assume that f ½SðN Þ ¼ Kf ½N ðSÞ
and K is a constant, which is determined by satisfying the condition (Eq 6)
ð
þ1

f ðSÞdS ¼ 1: (6)
1

Rearranging Eq 5, we have
8 h  i 2 9
>
< logðSÞ  alogðN Þ >
=
1 b
f ½SðN Þ ¼ pffiffiffiffiffi exp  (7)
2pðr=bÞ >
: 2ðr=bÞ2 >
;

and K ¼ b.
Clearly, the probabilistic distribution
h as ai function of stress range is still a normal
distribution with the mean of a  logðNÞ =b and the standard deviation of r=b.
The observation is that, no matter how a set of data is viewed, either vertically or hor-
izontally, the values of the parameters obtained from a curve-fitting method should
be exactly the same, as long as the data pattern is consistent with the standard pat-
terns. This point can be clearly
h demonstrated
i. by the fact that logðNÞ ¼ a  blogðSÞ
is equivalent to logðSÞ ¼ a  logðNÞ b. The slope (b) effects can also be demon-
strated by the geometrical relationship shown in Fig. 6b and c, which shows how the
slope changes the scatter or the standard deviation of the distributions.
In summary, by using the pragmatic and heuristic argument, the distribution of
the cycles to failure is transformed to the distribution of stress range, and the rela-
tionship between the two is firmly established. Although the deviation is not mathe-
matically rigorous, it does illustrate the relationship and capture the key factors and
the trends.

A Variable Transformation Technique for


Degradation (Accelerated) Testing Method
The key to making the degradation (accelerated) testing possible is building the
mathematical relationship between the final life distribution and the distribution of
a parameter at an earlier damage stage. The distribution of a parameter could be
the damage distribution at a specified applied cycle or the cycle distribution at a
346 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 6 (a) The relationship between the distribution of cycles to failure at a specified
stress level and the distribution of stress at a specified cycle and (b) a steep S-N
curve and (c) a flat S-N curve.

(a)

(b) (c)

specified damage level or both, Fig. 3. Mathematically, the problem is equivalent to


seeking a target distribution function FY ð yÞ for a given initial distribution function
FX ðxÞ and transformation functions: y ¼ uðxÞ and x ¼ wð yÞ. In fact, closed-form
solutions can be obtained by the following procedure [18], which is well-developed
and described briefly here.
The target distribution function FY ð yÞ can be expressed as

FY ð yÞ ¼ PðY  yÞ ¼ P½uð X Þ  y: (8)

First, consider the case where y ¼ uðxÞ is a strictly monotone increasing func-
tion; x ¼ wð yÞ is then a unique inverse function and
ð yÞ

FY ð y Þ ¼ fX ðxÞdx ¼ FX ½wð yÞ: (9)


1

The target PDF fY ð yÞ of Y is obtained as

dFY ð yÞ dwð yÞ
fY ð yÞ ¼ ¼ fX ½wð yÞ : (10)
dy dy
LUO ET AL., DOI 10.1520/STP159820160056 347

For cases where uðxÞ is a strictly monotone decreasing function,



FY ð y Þ ¼ fX ðxÞdx ¼ 1  FX ½wð yÞ (11)
wð yÞ

dFY ð yÞ dwð yÞ
fY ð yÞ ¼ ¼ fX ½wð yÞ : (12)
dy dy
The two cases in Eq 3 and Eq 5 can be combined as
 
dFY ð yÞ dwð yÞ
fY ð yÞ ¼ ¼ fX ½wð yÞ : (13)
dy dy 

The variable transformation technique shown in Eqs 8–13 is the essential part
of the degradation testing method. With this technique, the distribution of cycles to
failure can be easily calculated from the damage distribution at a given cycle or the
cycle distribution at a given damage. In reverse, if the final life distribution and the
damage evolution equation are known, the distribution of damage at any given cy-
cle and the cycle distribution at any given damage level can be calculated. Clearly,
this variable transformation technique is very different from the curve-fitting meth-
od, in which at least two stress levels and data extrapolation are required.
The distribution of original cycles to failure for normal, lognormal, and Weibull
distributions and their counterparts for the distributions of cycles at a specified
damage level and the distribution of damage at a specific cycle are listed in Table 1

for linear damage rule D ¼ N Nf . Table 2 lists the corresponding three distributions

q
of the Weibull distribution function for nonlinear damage rule D ¼ N Nf [19].
In Weibull distribution, b and g are the shape parameter and scale parameter (char-
acteristic life), respectively. It should be emphasized that, in the normal distribu-
tion, r and l are based on the arithmetic operation of the cycle Nf , whereas in the
lognormal distribution, r and l are based on the logarithmic operation of the cycle

log Nf . The transformations shown in Table 1 and Table 2 demonstrate how to


transform a distribution of cycles to failure to a distribution of cycles at any given
damage D (D < 1) and to a distribution of damage at any given cycles N. The pro-
cess is reversible; that is, from a distribution of cycles at any given damage
D (D < 1) or from a distribution of damage at any given cycles N, the distribution
of cycles to failure can be obtained uniquely. In fact, the reverse processes repre-
sents two degradation testing strategies, as shown in Fig. 3a and b.
It is worth mentioning some properties of the variable transformation for some
distributions. First, the transformed cycle distribution at a given damage as obtained
from the normal, Weibull, and lognormal distributions for the linear damage rule are
still their respective distributions. Second, the transformation of the two-parameter
Weibull distribution
n results
o in the complementary Weibull distribution, that is, the
CDF is exp ½N=ðDgÞb . These properties indicate that the degradation testing
data can be easily fitted using the derived distribution functions, and the estimated fit-
ting parameters can be directly used to construct a life distribution.
348 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

TABLE 1 Variable transformation for several distribution functions with the linear damage rule.

Linear Damage Rule: D 5 u(N, Nf) 5 (N/Nf)q


Nf 5 w(D) 5 N/D dw(D)/dD 5 (N/D2 )
Nf 5 w(N) 5 N/D dw(N)/dN 5 1/D

Function PDF f() CDF F()



"
2 #  
Normal f =F Nf 1 Nf  l 1 Nf  l
pffiffiffiffiffi exp  1 þ erf pffiffiffi
distribution 2pr 2r2 2 r 2
" # 
f =F ðN Þ 1 ðN  DlÞ 2 1 ðN  DlÞ
pffiffiffiffiffi exp  1 þ erf pffiffiffi
2pðDrÞ 2ðDrÞ2 2 ðDrÞ 2
  ( ) 
f =F ðDÞ N 1 ðN  DlÞ2 1 ðN  DlÞ
pffiffiffiffiffi exp  1 1 þ erf pffiffiffi
D 2pðDrÞ 2ðDrÞ 2 2 ð Dr Þ 2

(  2 )

Lognormal f =F Nf 1 logðNf Þ  l 1 log Nf  l
pffiffiffiffiffi exp  1 þ erf pffiffiffi
distribution 2prNf 2r2 2 r 2
2  
f =F ðN Þ 1 ½log N  ðlog D þ lÞ 1 logN  ðlogD  lÞ
pffiffiffiffiffi exp  1 þ erf pffiffiffi
2prN 2r 2 2 r 2
 
f =F ðDÞ 1 ½ log D þ ðlog N þ lÞ2 1 logDþ ðlogN þlÞ
pffiffiffi
pffiffiffiffiffi exp  1 1þerf
2prD 2r2 2 r 2

  ðb1Þ "  # "  #
Weibull f =F Nf b Nf Nf b
Nf b
exp  1  exp 
distribution g g g g
    "  # "  #
f =F ðN Þ b N ðb1 Þ
N b
N b
exp  1  exp 
Dg Dg Dg Dg
   ðb1Þ "  # "   #
f =F ðDÞ N b N N b
N b
exp  exp 
D Dg Dg Dg Dg

Case Studies
Several case studies are provided in this section to demonstrate the importance and
the effectiveness of the data-fitting approach and the variable transformation tech-
nique in accelerated testing and analysis.

TABLE 2 Variable transformation for Weibul distribution function with a nonlinear damage rule.

Nonlinear Damage Rule: D 5 u(N; Nf ) 5 (N/Nf )q


Nf 5 w(D) 5 ND1/q dw(D)/dD 5 (N/q)D(1/qþ1)
Nf 5 w(N) 5 ND1/q dw(N)/dN 5 1/D1/q

Function PDF f() CDF F()



  ðb1Þ "  # "  #
Weibull f =F Nf b Nf Nf b Nf b
exp  1  exp 
Distribution g g g g
  ðb1Þ "  # "  #
f =F ðN Þ b N N b N b
1=q 1=q
exp  1=q 1  exp  1=q
gD D g D g D g
   ðb1Þ "  # "  #
f =F ðDÞ b N N N b N b
exp  exp  1=q
g qD1=qþ1 D1=q g D1=q g D g
LUO ET AL., DOI 10.1520/STP159820160056 349

THE EFFECT OF DATA PATTERN ON THE VALUES OF FIT PARAMETERS


The importance of data pattern identification can be demonstrated by examining
the relationship among the fit parameters obtained with different standard data pat-
terns from a data set of the same source. Fig. 7a is a set of data with a standard

FIG. 7 Fatigue crack growth data with standard data patterns: (a) vertical pattern and
(b) horizontal pattern.

1.E-03
Crack growth rate, mm/SFH

1.E-04 Data

Vertical

Horizontal

Perpendicular

1.E-05
1 10 100
Stress intensity factor range, MPa.m^0.5
(a)
1.E-03
Crack growth rate, mm/SFH

1.E-04 Data

Vertical

Horizontal

Perpendicular

1.E-05
1 10 100
Stress intensity factor range, MPa.m^0.5
(b)
350 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

TABLE 3 Calculated fitting parameters with a ¼ log(C) and b ¼ m for the power law
da/dN ¼ C(DK )m, where N is number of cycles to failure and is DK stress intensity
factor range.

Vertical Offsets Horizontal Offsets

Vertical pattern a 26.168 6.414


b 2.510 2.817
Horizontal pattern a 5.754 26.187
b 1.997 2.535

vertical pattern trimmed from a fatigue crack growth data set from the aircraft in-
dustry [14]. The data shown in Fig. 7b follow the standard horizontal pattern. These
data with the standard patterns are then fitted with the corresponding curve-fitting
methods shown in Fig. 3a and b. The fit curves are also plotted in Fig. 7, and the fit
parameters are listed in Table 3, with the linearized equation y ¼ a þ bx. From the
highlighted data value in Table 3, it is clear that the fit parameters obtained by fitting
the data with the vertical pattern using the vertical offsets method and the data with
the horizontal pattern using the horizontal method are very close to each other,
while the values obtained using other methods vary widely. This clearly demon-
strates the importance of identifying a data pattern before fitting data to a curve;
the predicted best-fit curve should be consistent and accurate as long as the equilib-
rium direction and the corresponding standard data pattern are chosen consistently
and correctly for a given set of data. This observation validates the relationship be-
tween the life distribution in the horizontal direction and the stress distribution in
the vertical direction as established in Eq 5 and Eq 7.

CURVE FIT OF ELASTIC STRAIN-CYCLE DATA FOR LOW CYCLE FATIGUE


In strain-based low cycle fatigue, the total strain range-cycle curve can be constructed
by simply summing each individual elastic and plastic strain ranges at the same cycles
to failure. Low cycle fatigue data of 9Cr-1Mo steel were collected [20]. A total of 139
specimens were tested to failure, and the raw data as well as the elastic, plastic, and
total strain range versus cycles to failure data are shown in Fig. 8a. Only the elastic
data are analyzed here to demonstrate the importance of data pattern in curve fitting
because (1) 9Cr-1Mo steel shows large scatters, and (2) the elastic data is extremely
flat (small value of the slope). The elastic curve is determined with the two curve-
fitting methods: the horizontal offsets and vertical offsets methods. The results are
shown in Fig. 8b for the elastic curve. The fit results are summarized in Table 4.
A striking observation from Fig. 8b is that the mean curve determined from the
horizontal method, which is the ASTM E739-10, Standard Practice for Statistical Anal-
ysis of Linear or Linearized Stress-Life (S  N) and Strain-Life (e  N) Fatigue Data
method [10], is very different from the curves obtained from the vertical method.
Clearly, the fit curve obtained from the horizontal offsets method is not accurate.
LUO ET AL., DOI 10.1520/STP159820160056 351

FIG. 8 The relationship of (a) elastic strain, plastic strain, and total strain ranges-cycles
to failure and (b) elastic strain range-cycles to failure and the fit curves.

(a)

(b)

TABLE 4 Calculated fit parameters for 9Cr-1MO steel [20] with a ¼ log(C ) and b ¼ 1/m for the
power law De ¼ CN1/m.

Vertical Offsets Horizontal Offsets

Elastic a 2.3528 1.6141


b 0.0482 0.2810
Plastic a 0.1550 0.1465
b 0.6215 0.7165
352 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

Similar observations have been reported [14,15]. The main reason for the poor fit is
that the data pattern does not match the standard “horizontal pattern.” Therefore, the
horizontal offsets method should not be used. The large scatter and the flatness of the
data make the fit even worse with the horizontal offsets method. Actually, for this spe-
cific data pattern, the vertical offsets method should be used if data trimming is not
possible or is not convenient. Because the linearity seems to be an intrinsic feature for
the elastic part of the data, the separate treatment of the elastic and the plastic data
could be more representative of the real material properties than other methods. Fur-
thermore, the linearity of the data is more appropriate and reliable than other methods,
such as a simple polynomial fit of the total strain data, for accelerated testing. However,
when accelerated testing is conducted and data extrapolation is allowed, these two
curves obtained from the vertical and the horizontal methods lead to completely differ-
ent life at the service strain. Clearly, the data pattern identification and the selection of
an appropriate curve-fitting method is critically important in accelerated testing.

NONLINEAR CREEP-FATIGUE INTERACTION DATA


Creep-fatigue is a failure mechanism of materials under combined cyclic and sus-
tained loading at high temperatures [21]. In creep-fatigue analysis, creep damage
Dc and fatigue damage Df are usually calculated separately. Failure is conceded
when the sum of the fatigue and creep damages reaches a certain failure criterion.
Fig. 9a shows a set of creep-fatigue data [21], and Fig. 9b plots the trimmed creep-
fatigue data to show how the pattern of the data affects the values of fit parameters
and goodness of fit in nonlinear data analysis. First, the fit curve shown in Fig. 9a,
with a parabolic equation, does not fit the data well because the data pattern does
not follow the standard “vertical pattern” with which the fit is conducted. By follow-
ing the curve-fitting recommendations [14,15], the original data shown in Fig. 9a

FIG. 9 Curve fitting the creep-fatigue data (a) without consideration of data pattern
identification and (b) with data pattern consideration and data trimming.

(a) (b)
LUO ET AL., DOI 10.1520/STP159820160056 353

are trimmed, Fig. 9b, to follow the vertical pattern, and the fit curve now decently
matches the data, Fig. 9b. More details about how to fit the data shown in Fig. 9a
can be found in Wei, Luo, and Lin [22].

DEGRADATION FATIGUE TESTING AND DATA ANALYSIS


Estimate the Life Distribution from the Cycle Distribution at a Damage
Level D
A set of cycle data at D ¼ 0:01 is listed in Table 5. The two-parameter Weibull life

distribution shown in Table 1 and the linearndamage rule oD ¼ N Nf are assumed
for the data. Then the CDF of cycle, 1  exp ½N=ðDgÞb , at a given damage lev-
el D as shown in Table 1 can be used to estimate the values of the Weibull parame-
ters. Because the cycle distribution at any damage level D is still a Weibull
distribution as demonstrated in Table 1, the Weibull function at damage level D can
be directly used to estimate the parameters. The probabilistic plot and the corre-
sponding estimated parameters obtained using Minitab [23] from the data listed in
Table 5 are shown in Fig. 10a. The obtained shape parameter b ¼ 2:37219 and the
scale parameter of g ¼ 750:355 shown in Fig. 10a can be directly used to construct
the Weibull life distribution with the same value of the shape parameter and a pro-
portionally enlarged value of the scale parameter of g=D ¼ 750355. Because the
only difference between the life distribution and the damage distribution is the
factor D, the life distribution can be directly obtained from the cycle data at D by
multiplying a factor 1=D ¼ 100. The probability plot obtained from the trans-
formed data is shown in Fig. 10b, and the values of the shape and scale parameters
obtained are exactly the same as predicted, which confirms the effectiveness of the
extrapolation procedure of the degradation testing for the linear damage assumption.

Estimate the Life Distribution from the Damage Distribution at a Given


Applied Cycle N
Two-parameter Weibull life distribution functions and the linear damage rule

D ¼ N Nf are assumed in this example. In order to demonstrate the effectiveness
of procedure, the following approach is used [8]:
• Step 1: A set of tested life data with six test points is used to estimate the values
of the parameter of the Weibull life distribution.
• Step 2: The damage distribution at a given applied cycle N is obtained with the
variable transformation technique and the linear damage rule.
• Step 3: A set of “virtual” data is randomly generated from the damage distribu-
tion as obtained in Step 2. A Monte Carlo simulation technique can be used in
this step.

TABLE 5 The cycles at D ¼ 0.01.

Load level Cycles at D ¼ 0:01

90 281 492 661 690 710 1113


354 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 10 Probability plots obtained from (a) cycles at D ¼ 0.01 and (b) cycles to failure
(i.e., D ¼ 1.0).

Probability Plot
Weibull - 90% LB
Complete Data - LSXY Estimates
99
Table of Statistics
90 Shape 2.37219
80
70 Scale 750.355
60
50 Mean 665.040
40 StDev 298.266
Percent

30
20
Median 642.935
IQR 417.341
10
Failure 6
5 Censor 0
3 AD* 2.133
2 Correlation 0.972
1
100 1000
Cycles
(a)

Probability Plot
Weibull - 90% LB
Complete Data - LSXY Estimates
99
Table of Statistics
90 Shape 2.37219
80
70 Scale 75035.5
60
50 Mean 66504.0
40 StDev 29826.6
Percent

30
20
Median 64293.5
IQR 41734.1
10
Failure 6
5 Censor 0
3 AD* 2.133
2 Correlation 0.972
1
10000 100000
Cycles to failure
(b)

• Step 4: The values of the parameters of the damage distribution (eventually


the life distribution) are estimated and compared with the values originally
obtained in Step 1.
LUO ET AL., DOI 10.1520/STP159820160056 355

TABLE 6 The cycles to failure data.

Load level Cycles to Failure

120 19,729 33,394 40,963 41,726 43,959 52,873

The values of the parameters in the Weibull life distributions as obtained from
Step 1 and Step 2 should be comparable if the degradation approach (from the
damage distribution to obtain the life distribution) is effective and robust. More
details about the procedure can be found in Wei et al. [8].
To illustrate the approach, a set of test data is selected, which is shown in Table 6
and plotted in Fig. 11. The values of the obtained shape parameter and the scale
parameter are b ¼ 3:23046 and g ¼ 43345, respectively. From the damage distribu-
tion function, a set of randomly distributed data can be generated; subsequently, the
values of the parameters can be estimated from the generated random data. In this
example, 20 random data points are generated, and the values of the parameters are
estimated using the median rank regression method. In this procedure, the median
rank regression methodn is first used
o to linearize the accumulative complementary
Weibull CDF, exp ½N=ðDgÞb , as shown in Table 1. A Monte Carlo method is
subsequently used to estimate the shape parameter and the value is b ¼ 3:153, which
is close to the initially assigned value, b ¼ 3:230. The scale parameter back calculated

FIG. 11 The probability plot of the tested life data shown in Table 6 and the estimated
values of Weibull parameters.

Probability Plot
Weibull - 90% LB
Complete Data - LSXY Estimates
99
Table of Statistics
90 Shape 3.23046
80
70 Scale 43345.0
60
50 Mean 38840.0
40 StDev 13209.4
Percent

30
20
Median 38696.0
IQR 18482.6
10
Failure 6
5 Censor 0
3 AD* 2.134
2 Correlation 0.957
1
10000 100000
Cycles to failure
356 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

is g ¼ 42110:5, which is close to the initially assigned value, g ¼ 43345:0. Based on


many possible uncertainty sources in the estimation process—such as random num-
ber generation, median rank assumption, and so on, the agreement between the ini-
tial parameter values and the back calculated values is satisfactory; therefore, the
approach is effectively validated.

Discussion
Extrapolation operation is the basis for all kinds of accelerated testing [4–6].
Extrapolation is always risky, especially for degradation and accelerated degradation
testing, because so many uncertainties are involved in a degradation process. This is
particularly true for fatigue degradation testing because, in many cases, the definition
of damage is not clearly defined and the damage evolution is essentially unknown.
Identifying which is the most appropriate degradation parameter is a critical step.
The displacement, the temperature rise, the crack length, stiffness reduction, the load
or stress drop, and their combination all could be used as a degradation parameter.
However, none of them is directly related to the linear damage rule (i.e., Miner’s
rule). Therefore, even though the linear damage rule is commonly used in total life
assessment of components under fatigue loading, caution must be exercised when
dealing with damage assessment; otherwise, misleading results could be obtained. It
should be noted that, as damage diagnosis technologies advance, application of degra-
dation testing and accelerated degradation testing is expected to increase.
It should be noted that significant amounts of test data are required to estimate
the distribution of the population. However, in practice, it is difficult and often im-
possible to obtain the distribution of the population because of budget constraints,
timing, or both. In the automotive industry, the number of samples used for prod-
uct design and validation is usually very limited [1]. It should be noted that, in
durability testing of components, the sample size is usually very small, say six;
hence, the parameters estimated from the small sample size contain significant
uncertainty. Therefore, the confidence levels for the estimated parameters should be
provided. Several approaches have been developed to construct the confidence
intervals or limits for commonly used functions such as Weibull distribution [24].

Conclusions
In this paper, the test-to-failure method and degradation test, the two fundamental
accelerated testing methods are reviewed and thoroughly investigated, with emphasis
on the verification and validation of these methods and two critical aspects: (1) data
pattern identification and curve fitting and (2) variable transformation technique for
data extrapolation. The following conclusions can be drawn from this study:
(1) In curve fitting for accelerated testing, the vertical offsets method should be
used for data with a vertical pattern and the horizontal offsets method should
be used for data with a horizontal data pattern. For normally distributed data
(both vertical and horizontal), a pragmatic and heuristic approach demonstrates
LUO ET AL., DOI 10.1520/STP159820160056 357

the relationship between fitting parameters obtained from the vertical and
horizontal fitting approaches. The equilibrium arguments and the exam-
ples of trimmed linear data and nonlinear creep-fatigue data confirm these
conclusions.
(2) The variable transformation technique effectively and rigorously establishes
the mathematical relationship between the probabilistic distribution of one
variable to that of another. With this technique, traditional degradation test-
ing, which is based on data fitting and empirical extrapolation, can simply be
replaced with a few test data points at a stress or cycle level and the fatigue
damage rule.

ACKNOWLEDGMENTS
The authors are grateful to Prof. Gary Harlow for providing the low-cycle fatigue
data. The authors also thank Prof. Kamran Nikbin and Prof. Gary Harlow for their
sustaining contributions to the development of several new concepts and methodolo-
gies on data analysis over the last several years. The authors are grateful to Richard K.
Voltenburg for his helpful comments and suggestions.

References

[1] Lee, Y. L., Pan, J., Hathaway, R., and Barkey, M., Fatigue Testing and Analysis: Theory and
Practice, Elsevier Butterworth-Heinemann, Boston, 2005.

[2] Yang, G., Life Cycle Reliability Engineering, Wiley, New York, 2007.

[3] O’Connor, P. D. T. and Kleyner, A., Practical Reliability Engineering, 5th ed., Wiley, New
York, 2012.

[4] Nelson, W. B., Accelerated Testing: Statistical Models, Test Plans, and Data Analysis,
Wiley, New York, 2004.

[5] Meeker, W. Q., Escobar, L. A., and Lu, C. J., “Accelerated Degradation Tests: Modeling
and Analysis, Technometrics, Vol. 40, No. 2, 1998, pp. 89–99.

[6] Meeker, W. Q. and Escobar, L. A., “A Review of Accelerated Test Models,” Statistical
Science, Vol. 21, No. 4, 2006, pp. 552–577.

[7] Meeker, W. Q. and Escobar, L. A., Statistical Methods for Reliability Data, Wiley, New
York, 1998.

[8] Wei, Z., Start, M., Hamilton, J., and Luo, L., “A Unified Framework for Representing Prod-
uct Validation Testing Methods and Conducting Reliability Analysis,” SAE Int. J. Mater.
Manf., Vol. 9, No. 2, pp. 303–314, http://dx.doi.org/10.4271/2016-01-0269

[9] Lu, C. J. and Meeker, W. Q., “Using Degradation Measures to Estimate a Time-to-Failure
Distribution,” Technometrics, Vol. 34, No. 2, 1993, pp. 161–174.

[10] ASTM E739-10, Standard Practice for Statistical Analysis of Linear or Linearized Stress-
Life (S  N) and Strain-Life (e  N) Fatigue Data, ASTM International, West Consho-
hocken, PA, 2015, www.astm.org
358 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

[11] BS7608, Code of Practice for Fatigue Design and Assessment of Steel Structures, British
Standards Institution, London, 2014.

[12] Det Norske Veritas, “Fatigue Design of Offshore Steel Structures,” DNV-RP-C203, DNV,
Oslo, Norway, 2010.

[13] Shen, C., “The Statistical Analysis of Fatigue Data,” PhD dissertation, University of
Arizona, Tucson, AZ, 1994.

[14] Wei, Z., Yang, F., Cheng, H., Maleki, S, and Nikbin, K., “Engineering Failure Data Analysis:
Revisiting the Standard Linear Approach,” Eng. Fail. Anal., Vol. 30, 2013, pp. 27–42.

[15] Wei, Z., Maleki, S, and Nikbin, K., “Equilibrium Based Curve Fitting Method for Test Data
with Nonuniform Variance,” J. Press. Vessel Technol., Vol. 136, No. 2, 2014, http://
dx.doi.org/10.1115/1.4026020

[16] Neter, J., Wasserman, W., and Kutner, M. H., Applied Linear Statistical Models, Irwin,
Homewood, IL, 1990.

[17] Newland, D. E., An Introduction to Random Vibrations, Spectral and Wavelet Analysis,
3rd ed., Dover, Mineola, NY, 2005.

[18] Elishakoff, I., Probabilistic Theory of Structures, 2nd ed., Dover, Mineola, NY, 1999.

[19] Marco, S. M. and Starkey, W. L., “A Concept of Fatigue Damage,” Trans. ASME, Vol. 76,
No. 4, 1954, pp. 627–632.

[20] Harlow, D. G., “Low Cycle Fatigue: Probability and Statistical Modeling of Fatigue Life,”
Proceedings of the ASME 2014 Pressure Vessels and Piping Conference, Anaheim, CA,
July 20–24, 2014, http://dx.doi.org/10.1115/PVP2014-28114

[21] Spindler, M. W., “Effects of Dwell Location on the Creep-Fatigue Endurance of Cast Type
304L,” Mater. High Temp., Vol. 25, No. 3, 2008, pp. 91–100.

[22] Wei, Z., Luo, L., and Lin, S., “Correlation Measures and Their Applications in Structural
Dynamics and Data Analysis,” SAE Int. J. Mater. Manf., Vol. 8, No. 1, pp. 67–79, 2015,
http://dx.doi.org/10.4271/2014-01-2307

[23] Ryan, B., Joiner, B., and Cryer, J., Minitab Handbook: Update for Release 16, 6th ed.,
Boston, 2012.

[24] Dodson, B., The Weibull Analysis Handbook, 2nd ed., ASQ, Milwaukee, WI, 2006.
FATIGUE AND FRACTURE TEST PLANNING, TEST DATA ACQUISITIONS AND ANALYSIS 359

STP 1598, 2017 / available online at www.astm.org / doi: 10.1520/STP159820160065

Wenjing Wang,1 Jinyi Bai,2 Sichun Li,3 Hongwei Zhao,4


and Weiguang Sun5

Load Spectrum Test and Fatigue


Failure Study of High-Speed Train
Carbody Anti-Yawing Seat
Citation
Wang, W., Bai, J., Li, S., Zhao, H., and Sun, W., “Load Spectrum Test and Fatigue Failure Study
of High-Speed Train Carbody Anti-Yawing Seat,” Fatigue and Fracture Test Planning, Test
Data Acquisitions and Analysis, ASTM STP1598, Z. Wei, K. Nikbin, P. C. McKeighan, and
D. G. Harlow, Eds., ASTM International, West Conshohocken, PA, 2017, pp. 359–370, http://
dx.doi.org/10.1520/STP1598201600656

ABSTRACT
The anti-yawing damper installed between the carbody and bogie is an
important component to improve the hunting motion stability of a railway
vehicle. With the increase of service mileage, fatigue cracks occur in the weld
joint between the carbody anti-yawing seat and underframe side beam, which
seriously threatens the vehicle safety. A finite element simulation, bench test,
and on-track test were carried out to analyze the dynamic stress and load of an
anti-yawing seat. These studies showed that the longitudinal load induced by the
anti-yawing damper is the primary load acted on the seat, which is higher than
the design load and causes the weld joint failure.

Keywords
anti-yawing seat, load spectrum, stress spectrum, fatigue failure

Manuscript received March 11, 2016; accepted for publication November 1, 2016.
1
Beijing Jiaotong University, School of Mechanical, Electronics and Control Engineering, Haidian District,
Beijing 100044, China http://orcid.org/0000-0002-0779-7951
2
Beijing Jiaotong University, School of Mechanical, Electronics and Control Engineering, Haidian District,
Beijing 100044
3
Beijing Municipal Engineering Design & Research Institute Co. Ltd., Xicheng District, Beijing, 100045, China
http://orcid.org/0000-0001-6964-1090
4
CRRC Qingdao Sifang Co., Ltd, Qingdao 266111, China https://orcid.org/0000-0003-1717-5184
5
CRRC Qingdao Sifang Co., Ltd, Qingdao 266111, China
6
ASTM Symposium on Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis on May 4–5,
2016 in Grand Hyatt, San Antonio, TX.

Copyright V
C 2017 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.
360 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

Introduction
The rapid development of high-speed railways has meant increased requirements
for vehicle safety and stability. Although high-speed trains use tight-lock couplers
that have a minor longitudinal impact compared with the ordinary passenger
car and freight car, these longitudinal loads are still among the most important
loads on the carbody—especially when the vehicle is starting, accelerating, and
braking. The anti-yawing damper is the key component for inhibiting the yaw-
ing, rolling, and hunting between the carbody and bogie. After long-term opera-
tion, cracks appear in the connections between the anti-yawing damper seat
and carbody side beam [1–3] (as shown in Fig. 1), which seriously affect the car-
body life.
The load spectrum is an important component of engineering structure
strength in solving life evaluation and life extension problems. After decades of
domestic and foreign load spectrum research and engineering application, much
has been achieved [4–6]. A large amount of load spectrum data has become an
important foundation for fatigue design. In the field of railway vehicle load spec-
trum research, Chapter VII of the Association of American Railroads’ (AAR)
maintenance standard on newly built freight car fatigue design contains several en-
vironmental load spectrums from the track test and lists for specific freight cars as
well as each part of these spectrums [7]. However, research on the high-speed train
anti-yawing damper load spectrum has not been reported before now. Beijing-
Guangzhou’s dedicated high-speed passenger railway line is one part of China’s
"four longitudinal and four horizontal" passenger line networks. The Beijing-
Guangzhou line is longitudinal, and it is also the longest high-speed railway line in
the world, extending 2,118 km with more than 220 tunnels. Bridge tunnels account
for more than 60 % of the line’s length. The maximum design speed is 385 km/h
and, currently, the maximum operation speed is 310 km/h.

FIG. 1 Crack in anti-yawing damper seat.


WANG ET AL., DOI 10.1520/STP159820160065 361

The object of this paper is the anti-yawing damper seat of a 300-km/h high-speed
train that runs on the Beijing-Guangzhou dedicated passenger line. We perform a
long-term track load and dynamic stress test. The longitudinal loads and stress char-
acteristics of the anti-yawing damper seat are studied in the time and frequency do-
main under different line conditions. The results show that the weld fatigue cracks
between the anti-yawing damper seat and carbody side beam are mainly due to the
longitudinal load caused by the anti-yawing damper. The value of the equivalent load
spectrum is more than the design load.

Load Identification
DETERMINATION OF LOAD IDENTIFICATION POINT
The anti-yawing damper seat (as shown in Fig. 2) is an important part of the bogie
suspension system. The load is mainly along the installation direction of the anti-
yawing damper, and it is closely related to the train operation speed, running condi-
tions, and parameters of the damper. The longitudinal load identification point is
the arc part of the seat plate according to the finite element simulation (as shown
in Fig. 3).

CALIBRATION TEST AND RESULT


The calibration test uses a Material Testing Systems (MTS) servohydraulic system.
The anti-yawing damper seat is fixed on the test platform with four bolts, and longi-
tudinal force from 0 to 20 kN (every 5 kN) is applied at the height of the anti-
yawing damper installation (as shown in Fig. 4). The calibration coefficient between
load (kN) and strain (le) fitting the calibration data of each load identification
point was derived, as shown in Table 1.

FIG. 2 Three-dimensional model of anti-yawing damper seat.


362 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 3 Load identification points on anti-yawing damper seat: (a) Measure Points 1 and
2; (b) Measure Points 3 and 4.

(a) (b)

Longitudinal Load Characteristics of


Anti-Yawing Damper Seat
The dynamic strain-time history data for load identification points of the anti-
yawing damper seat were obtained from a high-speed train running on the Beijing-
Guangzhou dedicated passenger line (speed designed to 350 km/h) and on the
Shijiazhuang-Taiyuan dedicated passenger line (speed designed to 250 km/h). The
longitudinal load of the anti-yawing damper seat in operation can be deduced from
the calibration coefficients between load and strain. The longitudinal load charac-
teristics of the anti-yawing damper seat are given in this paper for high-speed
straight line, high-speed curve passing, and into/out of depot or station conditions.
Fig. 5 shows the longitudinal load-time history of the anti-yawing damper seat
in high-speed, straight-line conditions. The sampling time is 40 s. It can be seen
that the fluctuation cycle is relatively stable and the maximum longitudinal load is

FIG. 4 Calibration test site.


WANG ET AL., DOI 10.1520/STP159820160065 363

TABLE 1 Calibration coefficients of load identification points.

Component Point Number Calibration Coefficient (kN/le)

Anti-yawing damper seat 1 0.054


2 0.080
3 0.135
4 0.429

less than 5 kN, which shows that the interference level of track irregularity is minor
when running in a straight line at high speed.
Fig. 6 shows the longitudinal load amplitude-frequency spectra of the anti-
yawing damper seat under high-speed straight line running conditions (at speeds of
300 km/h and 250 km/h). It can be seen that the peak value appears at 10.8 Hz,
26.2 Hz at a running speed of 250 km/h and at 12.8 Hz, 30.8Hz with a running
speed of 300 km/h. Based on the track irregularity spectrum of the Beijing-
Shanghai Line, under the running speeds of 300 km/h and 250 km/h, the vibration
frequencies caused by a track plate with a length of 6.5 m are 10.8 Hz and 12.8 Hz
respectively, which is consistent with the anti-yawing seat vibration frequency. The
peak values of 26.2 Hz and 30.8 Hz are caused by the wheel rotation.
Fig. 7 shows the longitudinal load-time history of the anti-yawing damper seat
under curve-passing running conditions. The carbody and the bogie have relative
rotation, and the anti-yawing damper seat suffers an increasing load in comparison
to running on a straight line. The maximum peak reaches 7.5 kN when train runs
from the straight line to the curved line; subsequently, the peak decreases slowly,
which shows that the anti-yawing damper has a good inhibitory effect on the rela-
tive rotation between the carbody and bogie when the train passes a curve.
Fig. 8 shows the longitudinal load-time history of the anti-yawing damper seat
under into/out of depot conditions with a slow speed—generally below 15 km/h.

FIG. 5 Longitudinal load-time history under high-speed, straight-line conditions.


364 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 6 Longitudinal load frequency spectrum under high-speed, straight-line


conditions (Beijing-Shanghai): (a) at 300 km/h, (b) at 250 km/h.

(a) (b)

It can be found that the maximum longitudinal load is greater than when the train is
running on a straight line, and it reaches about 15 kN. However, the vibration fre-
quency is mainly concentrated in the range of 0–2 Hz, which is shown in Fig. 9. This
verifies that the tracks near a depot or station include multiple switches and small
radius curves; under these conditions, the changes in load amplitude are very obvious.

Fatigue Analysis of Anti-Yawing Damper Seat


CORRELATION COEFFICIENT AND COHERENCE ANALYSIS OF
LONGITUDINAL LOAD AND STRESS FOR ANTI-YAWING DAMPER SEAT
Based on the finite element results and crack location, the stress key points (1007/
1008) lie on the connection between the anti-yawing damper seat and the under-
frame side beam (shown in Fig. 10).

FIG. 7 Longitudinal load-time history under high-speed, curve-passing conditions.


WANG ET AL., DOI 10.1520/STP159820160065 365

FIG. 8 Longitudinal load-time history under into/out of depot conditions.

The longitudinal load and stress time-history waves of the anti-yawing damper
seat are shown in Fig. 11. It can be seen that the load-time history is highly consistent
with stress, which means that the longitudinal load is the main load of the anti-
yawing damper seat. The correlation coefficient between longitudinal load and
stress is more than 0.85 within 40 Hz, which is shown in Fig. 12.

FIG. 9 Longitudinal load frequency spectrum under into/out of depot conditions.


366 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 10 Stress key points of anti-yawing damper seat.

1007
1008

STRESS SPECTRUM AND LONGITUDINAL LOAD SPECTRUM


Based on the stress time-history signals, the rainflow counting method is used to
compile the stress spectrum of fatigue key points and the longitudinal load spectrum
of the anti-yawing damper seat in the into/out of depot, into/out of station, and

FIG. 11 Time-history waves of yaw damper longitudinal load and stress.


WANG ET AL., DOI 10.1520/STP159820160065 367

FIG. 12 Coherent coefficient between longitudinal load and stress.

high-speed straight line running conditions, which are listed in Tables 2 through 4.
Both high load and high stress occur under the into/out of station conditions.

FATIGUE ANALYSIS OF ANTI-YAWING DAMPER SEAT


The eight-level load spectrum is transferred to the constant amplitude load accord-
ing to Eq 1:
 m1
L X m
Faeq ¼ ni ðFi Þ (1)
L1 N

TABLE 2 Load spectrum of anti-yawing damper seat.

Into-Out of Depot Into Station Out of Station High Speed Straight Line

No. Amp Cycles Amp Cycles Amp Cycles Amp Cycles

1 0.46 161,000 0.82 216,000 0.68 211,000 0.40 222,000


2 1.38 831 2.46 2154 2.02 4,539 1.19 47,200
3 2.30 205 4.10 172 3.37 464 1.97 12,700
4 3.22 62 5.74 38 4.72 61 2.76 2,465
5 4.14 23 7.38 18 6.07 23 3.55 315
6 5.06 5 9.02 12 7.41 4 4.34 33
7 5.98 4 10.66 3 8.76 7 5.13 7
8 6.90 1 12.29 2 10.11 4 5.92 1
368 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

TABLE 3 Stress spectrum of anti-yawing damper seat fatigue key point (1007).

Into-Out of depot Into Station Out of Station High Speed Straight Line

No. Amp Cycles Amp Cycles Amp Cycles Amp Cycles

1 0.28 228,000 0.56 411,000 0.58 428,000 0.31 644,000


2 0.84 4,106 1.68 7,957 1.75 8,813 0.92 154,000
3 1.40 379 2.80 293 2.92 416 1.53 28,900
4 1.96 103 3.92 43 4.09 39 2.15 4,322
5 2.51 41 5.05 27 5.25 12 2.76 415
6 3.07 10 6.17 8 6.42 2 3.37 36
7 3.63 6 7.29 2 7.59 1 3.98 8
8 4.19 2 8.41 2 8.75 1 4.60 1

where:
ni is cycle times of level i,
Fi is the amplitude of level I,
m is power of the S-N curve (Sm  N ¼ C), m ¼ 3.5,
Faeq is the equivalent load amplitude, and
N is the cycle number, 2 ¼ 106.

With Eq 1, the equivalent load amplitude was deduced as 12.4 kN, which is greater
than the longitudinal fatigue load (9.8 kN) calculated based on the EN13749 stan-
dard and on damper parameters. Therefore, considering and applying this load am-
plitude in future anti-yawing damper seat fatigue testing is suggested.
Table 5 lists the damage values of key fatigue points calculated by the longitudi-
nal load spectrum and stress spectrum, respectively. It can be seen that damages at
these locations are high and that the damage calculated by the actual load spectrum
is coincident with that of the stress spectrum. Combining this with exposure condi-
tions could result in fatigue failures of some anti-yawing damper seats.

TABLE 4 Stress spectrum of anti-yawing damper seat fatigue key point (1008).

Into-Out of Depot Into Station Out of Station High Speed Straight Line

No. Amp Cycles Amp Cycles Amp Cycles Amp Cycles

1 0.24 259,000 0.54 397,000 0.48 363,000 0.32 692,000


2 0.72 6,425 1.62 6,146 1.6 8,203 0.98 144,700
3 1.2 585 2.7 250 2.72 1,053 1.64 21,683
4 1.66 192 3.76 39 3.84 199 2.28 3,024
5 2.14 96 4.84 19 4.96 30 2.94 84
6 2.62 35 5.92 10 6.08 13 3.6 26
7 3.1 8 7 4 7.2 8 4.24 12
8 3.58 3 8.08 1 8.32 3 4.9 1
WANG ET AL., DOI 10.1520/STP159820160065 369

TABLE 5 Damage of key fatigue points.

Points Fatigue Damage by Stress Spectrum Fatigue Damage by Load Spectrum

1007 0.581 0.572


1008 0.524 0.503

Conclusion
This paper summarized bench and track test research undertaken to make it clear
why fatigue cracks occur in carbody anti-yawing damper seats and to provide a
load spectrum for structure improvement.
(1) The load-stress correlation coefficients were determined through a calibration
test on an MTS bench.
(2) The load characteristics of the anti-yawing seat under different running con-
ditions were discussed. It was shown that the maximum longitudinal load
occurs under in/out of depot or station conditions. The load amplitude
reaches about 15 kN.
(3) Correlation analysis between dynamic stress of the key fatigue point and
anti-yawing longitudinal load was carried out, and the correlation coefficient
was more than 0.85. Based on the load spectra and stress spectra, the struc-
tural damage was calculated separately, and the damages were basically con-
sistent and high. The aforementioned studies suggest that the longitudinal
load induced by the anti-yawing damper is the primary load acting on the
seat and that it is larger than the design load and causes aluminum weld joint
failure under exposure conditions.

ACKNOWLEDGMENTS
The authors are grateful for financial assistance from the National Science and Tech-
nology Support Program of China (Grant No. 2015BAG12B01) and the National Ba-
sic Research Program of China (Grant No. 2015CB654805).

References

[1] Jiling, B. and Youquan, F., “Influence of Anti-Yaw Damper on Locomotive Operation
Quality,” Electric Locomotives & Mass Transit Vehicles, Vol. 27, No. 6, 2004, pp. 6–8.

[2] Mellado, A. C., Gómez, E., and Viñolas, J., “Advances on Railway Yaw Damper Character-
isation Exposed to Small Displacements,” Int J Heavy Veh Syst, Vol. 13, No. 4, 2006,
pp. 263–280.

[3] Liangliang, Y. and Shihui, L., “Influence of Anti-Yaw Damper Mounted Stiffness on Vehi-
cle Dynamics Performance Based on Flexible Frame,” Electric Drive for Locomotives,
Vol. 35, No. 6, 2012, pp. 15–18.
370 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

[4] Klemenc, J. and Fajdiga, M., “Improved Modelling of the Loading Spectra Using a Mixture
Model Approach,” Int J Fatigue, Vol. 30, No. 7, 2008, pp. 1298–1313.

[5] Conghua, H., Shuyun, Q., and Miao, Wu, “Simulating Study of Dynamic Load Spectra
Identification Method of Machinery in Cepstrum Domain,” J China Univ Mining & Tech,
Vol. 16, No. 1, 2006, pp. 22–24.

[6] Epaarachchi, J. A. and Clausen, P. D., “The Development of a Fatigue Loading Spectrum
for Small Wind Turbine Blades,” J Wind Eng Ind Aerod, Vol. 94, No. 4, 2006, pp. 207–223.

[7] Gao, Z., “Compiling of Fatigue Load Spectrum,” AcAAS, Vol. 2, No. 2, 1980, pp. 36–47.
FATIGUE AND FRACTURE TEST PLANNING, TEST DATA ACQUISITIONS AND ANALYSIS 371

STP 1598, 2017 / available online at www.astm.org / doi: 10.1520/STP159820160084

Mark T. Seitz,1 Jason D. Hamilton,1 Richard K. Voltenburg,1


Limin Luo,1 Zhigang Wei,1 and Robert G. Rebandt1

Practical and Technical


Challenges of the Exhaust System
Fatigue Life Assessment Process
at Elevated Temperature
Citation
Seitz, M. T., Hamilton, J. D., Voltenburg, R. K., Luo, L., Wei, Z., and Rebandt, R. G., “Practical and
Technical Challenges of the Exhaust System Fatigue Life Assessment Process at Elevated
Temperature,” Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis, ASTM
STP1598, Z. Wei, K. Nikbin, P. C. McKeighan, and D. G. Harlow, Eds., ASTM International, West
Conshohocken, PA, 2017, pp. 371–401, http://dx.doi.org/10.1520/STP1598201600842

ABSTRACT
Exhaust system durability and reliability assessments are challenging due to the
inherent complexities of mechanical, thermal, and chemical interactions, and
the variability associated with material properties, loading, and environment. The
traditional methodology for predicting exhaust system durability and reliability
was developed to provide a pragmatic, conservative, and economically viable
solution—hence, the predominance of fatigue life approximations in cold or ambient
conditions. Through the definition of target loading using measured field data
collection, and determination of component durability capability through simple
uniaxial or biaxial oscillating load bench tests, system level life assessment is
possible. Exhaust system durability and reliability estimates conducted in cold
conditions require that the life prediction method include a level of conservatism to
account for actual field usage thermal conditions. The ability to bypass the engine
out hot exhaust gas has become increasingly difficult due to the elimination of
standard exhaust manifold designs. Some designs require the converter inlet pipes
to be directly attached to the engine head without any exhaust manifold present.
The natural progression in exhaust system durability assessment is to conduct all

Manuscript received March 22, 2016; accepted for publication October 5, 2016.
1
Tenneco Inc., Clean Air Division, 3901 Willis Rd., Grass Lake, MI 49240
2
ASTM Symposium on Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis on May 4–5,
2016 in Grand Hyatt, San Antonio, TX.

Copyright V
C 2017 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.
372 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

measurements inclusive of engine out elevated temperature conditions and to


conduct component oscillating load bench testing at elevated temperatures. There
are many challenges associated with the development of a robust life assessment
methodology encompassing actual vehicle usage environmental conditions of an
exhaust system. In this paper, work conducted to evaluate new process integration,
encompassing high-temperature life assessment, is introduced. Oscillating load
bench tests are conducted to develop allowable load-life curves, the methods of
strain calibration for hot gages explored, and the method for conversion of load-life
curves to local hot strain life curves conducted. The critical aspects of the
assessment process, such as the determination of the thermal modification factor
and the procedure for transforming a loading profile into a damage equivalent form
at an equivalent temperature for rainflow counting, are explored. The assumptions,
limitations, and viability to integrate these into the standard design and development
process are evaluated.

Keywords
exhaust, durability, elevated temperature, fatigue life

Introduction
Vehicle exhaust components and systems serve both emission control and noise
reduction purposes [1]. Fig. 1 schematically shows the overall structure of a vehicle
exhaust system, which consists of several exhaust components. During vehicle oper-
ation, the high-temperature exhaust gas generated inside the engine flows into
exhaust manifolds, through the exhaust system, and is finally discharged out the
tailpipe. Based on temperature level, an exhaust system can be roughly divided into
two parts: hot end and cold end. At the hot end, the manifolds are directly attached
to the engine and are usually the hottest part in an exhaust system, with peak tem-
peratures as high as 1000 C or even higher [2–5]. The other exhaust components in
the hot end, such as downpipe, catalytic converter, and sometimes flex coupling,
are also subjected to high temperature. At the cold end, the temperature is relatively
low, but the peak temperature at muffler and tailpipe can be 400 C or higher.

FIG. 1 Schematic of a vehicle exhaust system.


SEITZ ET AL., DOI 10.1520/STP159820160084 373

Vehicle exhaust components and systems are almost invariably subjected to me-
chanical cyclic loadings due to vibratory loads from the engine and roads and to the
thermal cycling from the engine heating-up and cooling-down cycles [1,6]. There-
fore, the mechanical and thermal loading characteristics, the temperature effects in
particular, must be properly considered in the durability and reliability performance
assessment of the exhaust components and systems.
Product durability and reliability validation testing and associated life assess-
ment are becoming routine processes for the development of exhaust components
and systems. In product validation, how to handle the temperature effects is still a
controversial issue. Generally, there are two conflicting approaches: (1) Cold testing:
In order to reduce cost and shorten the product development cycle, the hot gas in a
vehicle is often bypassed during the road load data acquisition (RLDA) process;
hence, room or near-room temperature information is collected during RLDA. For
the purpose of consistent performance evaluation, subsequent calibration testing
and component bench testing are also conducted in the same cold conditions [1].
With the cold testing information, the performance of the component or system at
high operating temperatures can be estimated by introducing a temperature factor,
which is used to correct and compensate the temperature effects [1]. With the
introduced temperature factor, the product designed in the cold testing condition
could be reliable if the dominating factors are properly considered in the tempera-
ture factor. (2) Hot testing: As the name implies, all parts of the RLDA, calibration,
and bench testing are conducted in service or equivalent high-temperature condi-
tions. The fatigue life can be assessed in service condition, and no temperature cor-
rection factor is required in the fatigue life assessment.
The advantage of cold testing is that it is cost-effective. The drawback of cold
testing is that the failure mechanisms and modes as observed in real operation can-
not be revealed. Additionally, the cold-testing approach can only be applied to
components outside the proximity of exhaust manifolds. The exhaust manifold
temperature is high enough to make creep and oxidation mechanisms significant
compared to the mechanical fatigue [3,4]. The durability and reliability testing of
these high-temperature components must be conducted in hot conditions. Further-
more, the design based on the cold testing procedure could be either over-
conservative or non-conservative if the uncertainties that affect cold-hot conversion
were not properly considered. Moreover, in conjunction with the elimination of ex-
haust manifolds, the ability to effectively bypass engine out hot exhaust gas has be-
come increasingly difficult. By contrast, hot testing can be more representative of
the thermal conditions a vehicle experiences during operation. Therefore, it should
be more reliable if the testing methods and life assessment methods are properly
developed. However, hot testing poses significant challenges: It is not only more
expensive and time consuming but also technically more difficult to conduct than
cold testing. Uncertainties relevant to hot testing can be introduced and eventually
hinder the data interpretation if the testing procedure is either not properly con-
trolled or the underlying theories are not well-established. Nevertheless, recent
374 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

increasing efforts in hot testing indicate that there seems to be an overall trend to-
ward hot testing and validation across all testing levels. Clearly, with the increased
trend and demands for hot testing, verified, reliable, and robust hot testing and
associated life assessment procedures are in critical demand.
In this paper, the failure mechanisms as experienced by exhaust systems under
various thermal and mechanical loadings are reviewed and discussed first. Eventually,
based on the failure mechanisms and the associated analysis approaches, the loading
cases are categorized into three typical cases: (1) isothermal fatigue, (2) anisothermal
fatigue, and (3) high-temperature thermo-mechanical fatigue (TMF). The current
cold-testing procedure based on isothermal fatigue is then reviewed, and several pos-
sible approaches for cold testing and hot testing are proposed, with emphasis on ani-
sothermal fatigue hot testing and related life assessment approaches. The focus of the
paper will be on the basic concepts involved in the elevated temperature. Finally, sev-
eral case studies are provided to illustrate the procedures of the proposed cold-testing
and hot-testing methods, and the results are discussed.

Categorization of the Thermal and Mechanical


Loadings
Testing and associated engineering models for life assessment must be based on a
complete understanding of the failure mechanisms and modes. Vehicle exhaust
components and systems often experience a wide variety of failure mechanisms
depending on the loading, temperature, and environment. The most commonly
observed failure mechanisms include fatigue, corrosion, oxidation, and creep [3–5].
Fatigue is essentially a cycle-dependent failure mechanism under vibratory environ-
ment caused by engine vibration, thermal cycling due to engine heating-up and
cooling-down cycle, road condition, and so on [7]. Creep, oxidation, and corrosion
are basically time-dependent failure phenomena. Creep and oxidation of metals are
usually issues for components such as vehicle manifolds operated at high tempera-
tures [8]. Creep begins at a temperature of approximately half the absolute tempera-
ture (degrees Kelvin or Rankine) of the metal melting point [8]. Corrosion in
vehicle exhaust systems is usually caused by salt, condensate, urea, and other corro-
sive agents. Corrosion and combined mechanical and corrosion damage/failure
mechanisms, such as corrosion-fatigue, stress-corrosion-cracking, and other
environment-related mechanisms, are very complex [9] and will not be addressed
in this paper.
In a previous paper [6], based on the temperature range, the associated material
damage and failure mechanisms, and related analysis approaches, the thermal and
mechanical loadings are categorized into three groups: (1) isothermal fatigue, (2)
anisothermal fatigue, and (3) high-temperature thermal-mechanical fatigue (TMF).
(1) Isothermal fatigue is where the temperature remains relatively low and con-
stant. In this condition, the high-temperature failure mechanisms such as
creep and heavy oxidation do not appear, or they play a negligible role in the
SEITZ ET AL., DOI 10.1520/STP159820160084 375

damage and failure process. Room-temperature fatigue is a typical example


of isothermal fatigue. The temperature and the mechanical loading profiles of
the isothermal fatigue are schematically shown in Fig. 2a. The fatigue life
assessment of materials in isothermal loading conditions is well-established,
and it is already a standardized procedure in engineering applications [7].
The procedure consists of two parts: (a) obtain the range-cycle matrix using a
rainflow cycle counting method for a given loading input, which can be
stress-time (S-t) data and strain-time (E-t) data. A range-mean-cycle matrix
can also be used if mean stress effect is a concern and (b) fatigue behavior of
the materials at the given temperature, which is often represented by fatigue
stress-cycle (S-N) and strain-cycle (E-N) curves.
(2) Anisothermal fatigue is where the temperature varies and does not have a
single fixed temperature. In real applications, this type of fatigue is typically

FIG. 2 Three typical loading conditions experienced by exhaust components and


systems.

(a) Isothermal loading (b) Anisothermal loading

(c) High-temperature
thermo-mechanical loading
376 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

represented as relatively long thermal cycles superimposed with relatively


high-frequency vibration loadings. Similar to isothermal fatigue, the applied
temperature should be low enough to avoid triggering other failure mecha-
nisms. Failure mechanisms for both anisothermal and isothermal fatigue are
the same, but the analysis of anisothermal fatigue is more difficult because of
the variable temperature. There is no well-accepted method for fatigue life
assessment in the anisothermal fatigue condition. The temperature and the
mechanical loading profiles of the anisothermal fatigue are schematically
shown in Fig. 2b. Most of the exhaust components at cold end and some
components at hot end are subjected to anisothermal fatigue.
(3) High-temperature TMF is where time-dependent mechanisms such as creep
and oxidation are involved and play an important role in failure and life.
High-temperature TMF loading can often be categorized into in-phase and
out-of-phase loadings according to the loading phases. For in-phase loading,
stress and temperature increase and decrease coincidently. The out-of-phase
loading is just opposite. The temperature and the mechanical loading profiles
of the high-temperature TMF are schematically shown in Fig. 2c. Manifolds
are typical components that are subjected to high-temperature TMF. The
high-temperature TMF is complicated because of complex deformation,
damage, failure mechanisms, nonlinearity, and uncertainties in testing and
modeling. TMF has been intensively studied because of its uniqueness and
significant importance in applications [3,4,10–13].
It should be noted that the categorization of thermal and mechanical loadings
into the isothermal, anisothermal, and high-temperature TMF is just a simplified
classification. Although most of the exhaust components can be described by only
one of these three categories, some components may be subjected to their combina-
tions. For example, Halfpenny, Anderson, and Lin [13] indicate that a catalytic con-
verter mounted in the exhaust system experiences significant mechanical vibration
loads arising from road and transmission sources but also experiences transit TMF
during the heat-up and cool-down cycles. Superposition and Miner’s rule are com-
monly used to sum their separate damage contributions [13].
Because most of the exhaust components, excluding the manifolds, are domi-
nated by vibratory road fatigue and enhanced by the superimposed thermal effects,
the thermal-mechanical loading conditions can be roughly approximated as aniso-
thermal fatigue problems. The development of the hot testing and associated life
assessment at anisothermal fatigue is the focus of this paper.

Product Durability and Reliability Assessment


Procedure Based on Cold Testing
To understand the role of the temperature effects in product durability and reliability
performance, it is necessary to review the current cold-test procedure, which provides
essential background and technical basis for the development of the hot-testing proce-
dure. The cold-testing procedures are briefly described here [14].
SEITZ ET AL., DOI 10.1520/STP159820160084 377

THE COLD-TESTING PROCEDURE


Generally, the current cold-testing procedure consists of the following four steps:
(1) RLDA, (2) calibration, (3) component level bench testing, and (4) damage and
life assessment. The process can be best described by the flowchart shown in Fig. 3.
The purpose of conducting component bench tests, RLDA, and dead weight cali-
bration (DWC) is to merge data from each of these tests to perform the safety factor
calculation for each component. The flowchart of the product validation procedure
is schematically shown in Fig. 3a. First, the road load data in terms of nominal strain
(e) are acquired using strain gages at a specific area, which is close to the area of
concern, for example, the weld or notch or other stress concentration sites. Subse-
quently, the calibration testing (e.g., dead weight calibration) is conducted on the
component. In the calibration procedure, the load (P)-nominal strain (e) relationship
is recorded. With the calibration, the load history PðtÞ the component experienced
during RLDA can be estimated. With the proper load finder, the load range (DP)-
cycle (N) relationship at the appropriate cycle range can be obtained by conducting
the fatigue bench test for the component with limited samples. “Load finder” here is
defined as a process to find a load at which an ideal cycle to failure is targeted.
Accordingly, the failure distribution can be estimated with the samples with a certain
confidence. One-load and two-load level bench tests can be conducted. Therefore, the
RLDA can be treated as the driving force for the component to failure, and the bench
test represents the fatigue resistance capability. The calibration testing links the two
parts and the fatigue life of the component can be assessed with PðtÞ, DP-N, and the
linear damage rule (Miner’s rule). This approach is exactly the same as the traditional
S-N approach for high cycle fatigue [7].

Road Load Data Acquisition (RLDA)


The purpose of RLDA is to collect data that an exhaust system and its components
experience. RLDA is usually acquired at a proving ground and is made up of a series
of different events such as a pothole, hill, and gravel road. The data collected can be
in many forms including strain, acceleration, temperature, pressure, and so on. The
acceleration data from the RLDA runs can then be double-integrated to produce a
displacement time signal, which can be used as input for finite element analysis
(FEA). The time signal from each channel for each event is recorded and saved using
computer equipment. The signal for each event is repeated a designated number of
times to make up a full schedule that simulates what an exhaust system can expect to
experience during its required lifetime. When RLDA is collected, the strain gage is
placed in predetermined places where failure is likely to occur. The best-case scenario
is when strain gages are placed near the same failure mode that is seen during the
bench test. This ensures that the load being measured is the same load that causes the
component to fail. However, the local strain cannot be measured because the strain
gage needs to be at the failure location and would need to be very small. Nominal
strain can be measured at the location, which is usually 5 mm to 10 mm from the
notch. A typical RLDA nominal strain time history signal is shown in Fig. 4.
378 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 3 (a) The flowchart and (b) the key steps of the load-based product durability and
reliability performance assessment procedure.

(a)

(b)

Dead Weight Calibration


After RLDA is collected, the exhaust system is taken off the vehicle with all the
strain gages still in place. It is then placed in a fixture to hold it in place while
weights are hung from specific locations up to a predetermined weight, a so-called
dead weight calibration. The load-strain response, usually linear, is measured, and
the slope of the linear relationship is calculated. The factor builds a relationship
between load and strain to link the RLDA strain and the bench test. It is critical for
SEITZ ET AL., DOI 10.1520/STP159820160084 379

FIG. 4 An RLDA strain time history.

the calculation of the damage and safety factor. An example of dead weight calibra-
tion is shown in Fig. 5.

Component Bench Testing


During a bench test, a component is cyclically loaded at a constant load until it
fails. Some of the common parts on an exhaust system that are tested are bush-
ings, hangers, flanges, and so on. Fig. 6 shows a typical bench test set-up for
hanger testing.

FIG. 5 Dead weight calibration.


380 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 6 A typical bench test set-up for hanger testing.

Multiple samples are tested to failure at one, two, and sometimes three load lev-
els. Usually, a certain amount of samples, say six, at a load level are repeated to cap-
ture the scatter characteristics and the probabilistic behavior of the fatigue data.
Tests with two load levels are commonly conducted to determine both the slope
and the intercepts of the linear fatigue curve. The component S-N curves are
assumed to be linear in a log-log plot. The formula is roughly correct as long as
the cycles to failure are not in the very high cycle regime (>106), where fatigue
limit is reached, or in the low cycle regime (<103), where the strain-based non-
linear mechanisms are triggered. Three-load level testing is used to capture the
bilinear fatigue behavior. The test load levels are selected by a load finder pro-
cess, with which the mean lives are targeted at say 50,000 and 500,000, respec-
tively, for the upper load and the lower load of the two-load level tests. With
statistical methods and tools, the geometric mean of the cycles to failure at each
load level is computed to obtain the average cycles to failure at a given load. The
cycle at a specified reliability R (xx %) with a confidence C (yy %) level (e.g.,
R90C90 and R99C50) can be found from the associated probability plot, in
which the uncertainty caused by the sample size is captured using the criterion
RxxCyy [7]. A typical probability plot for two sets (upper load and lower load)
of fatigue life test data generated using Minitab is shown in Fig. 7, in which the
two-parameter Weibull distribution is assumed for the distribution of the data
and the least squares method is used to fit the ranked data.
Based on the obtained test data, a mean S-N curve and a parallel left-side design
curve based on RxxCyy can be generated, and both curves can be expressed in a
power law equation [7]:
SEITZ ET AL., DOI 10.1520/STP159820160084 381

FIG. 7 Probability plots for two-load level test data with sample size of six at each load
level.

Probability Plot
Weibull - 90% LB
Complete Data - LSXY Estimates
99
Variable
90 C1
80 C2
70
60
50 Table of Statistics
40 Shape Scale Corr F C
Percent

30
2.27472 125628 0.991 6 0
20 2.93244 66357 0.950 6 0

10

5
3
2

1
10,000 100,000
Cycles to failure, N

DP ¼ CN b ; b < 0 (1)

where the variable DP represents the load range. Strain range De and stress range
Dr for elastic deformation are often used for high cycle fatigue; C is a constant and
is the intercept of the line in the S-N plot at N ¼ 1; N is cycles to failure and the
exponent b is a constant, which represents the slope of the S-N line in the log-log
plot. In component bench testing, the loads are usually measured in pounds or
Newtons in the case of hanger components and ft-lbs or Newton-Meters in the case
of a bushing or flange components. The value of b can be directly calculated from the
two-load level bench test. For steel, b ¼ 1=3 is often used in some industry stand-
ards [15,16].

Life Assessment
Several other conditions need to be considered to take into account the factors and
their uncertainties that are not accounted for in the cold testing. In practice, to be
conservative, a scale factor is picked and applied to the RLDA in order to shadow
all the uncertainties in the cold-testing method. The most important factors include
the thermal modification factor KC , proving ground factor KPG , and vehicle factor
KVEH . All of these factors should be properly considered [1]. The values of these fac-
tors are usually based on historical data collection, engineering experience, data cor-
relation, and, sometimes, on certain preliminary analyses. For the strain-based
method, which will be addressed in detail in this paper, an amplification factor Kf
382 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

should be considered in addition to the other factors mentioned earlier. Once all of
these factors are identified, quantified, and applied to the RLDA load or strain data,
the rainflow cycle counting can be performed with the help of the linear Miner’s
damage rule. Miner’s rule predicts that failure occurs when damage is greater than
or equal to one. Two safety factors can be used to evaluate the margin of safety: the
load safety factor and the cycle safety factor. Fig. 8 schematically shows the safety
factors as obtained by comparing the design curve (R90C90) as obtained from the
bench test and the damage curve as obtained from the RLDA analysis.

Consideration of Temperature Effects in Product


Validation
The following four fundamental technical issues are investigated in this section:
(1) development of a thermal modification factor for both load-based and strain-
based cold testing, (2) a load transformation technique to convert the variable load-
ing data under variable temperature environment into an equivalent series of
loading data at a reference temperature, (3) an interpretation of the damage mecha-
nism under the combined variable loading and temperature, and (4) the develop-
ment of “equivalent” temperatures for hot testing. These four issues are the
fundamental technical barriers to the development of cold testing and hot testing
for vehicle exhaust products as well as for other products with similar loading and
temperature conditions.

THERMAL MODIFICATION FACTOR FOR COLD TESTING


The purpose of introducing a thermal modification factor is to allow engineers to
evaluate fatigue damage using material properties at room temperature, with RLDA

FIG. 8 Safety factors as obtained by comparing the design curve (R90C90) as


obtained from the bench test and the damage curve as obtained from the RLDA
analysis.
SEITZ ET AL., DOI 10.1520/STP159820160084 383

and DWC being conducted at room temperature, while the system experiences
simultaneously the vibratory and thermal loads in the field [1]. Hence, compensa-
tion for temperature’s effect on the exhaust components is needed. In damage cal-
culations, the measured load or strain history are multiplied by the thermal
modification factor KC for further rainflow counting and damage assessment.
Because fatigue life is very sensitive to the load or strain level, the accuracy in KC
estimation is critically important. Procedures for determining the thermal
modification factors are developed and reported here for both load-based and
strain-based cold-testing methods. In these procedures, the component bench
tests under the same load are conducted and the geometrical mean of the fatigue
cycles to failure at the room temperature and the operating temperature or the
equivalent temperature are compared and used to calculate the thermal modifica-
tion factor. The load-based method is straightforward, whereas the strain-based
method is more convoluted because of the involvement of the strain amplification
effects, which reflects the relationship between the measured strain in the location
and the equivalent strain in terms of the unnotched base material at the failed
notch location.

Load-Based Thermal Modification Factor


The development of the thermal modification factor KCP for the load-based method
can be best reflected in Fig. 9, which basically shows the linear relationship between
the applied load range DP and the cycles to failure Nf at different temperature levels
in a log-log plot. The curves at different temperature levels are parallel to each other
for simplicity purposes and that is usually a good approximation for most engineer-
ing materials. Additionally, the fatigue life is usually a decreasing function of

FIG. 9 Thermal modification factor for the load-based cold testing.


384 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

temperature at the same applied loads, so that the curves go downward as tempera-
ture increases.
The procedure for estimating the thermal modification factor for the load-
based method can be stated as follows:
1. Conduct cold (at room temperature TR ) component bench test at a load level
and record the geometrical mean of the cycles to failure, NfR , which is the cycle
vertically projected onto the logðNÞ axis from Point A shown in Fig. 9.
2. Conduct hot (at an equivalent constant temperature TE or with a high-
temperature profile) component bench test at the same load level and record
the geometrical mean of the cycles to failure, NfE , which is the cycle vertically
projected onto the logðNÞ axis from Point B shown in Fig. 9.
3. Calculate the distance between the two geometrical means of the cycles to
   
failure: jABj ¼ log NfR  log NfE .
4. Project Point B onto the load-life line at the room temperature TR and Point C
is formed; calculate the distance between the two geometrical means of the
loads: kP ¼ jBCj ¼ logðDPE Þ  logðDPR Þ, which can be calculated from the
existing mean load-cycle curve generated from a two-load level test or estimat-
ed when the slope b of the curve is known in advance, and the formula
kP ¼ jBCj ¼ bjABj. b ¼ 1=3 is often assumed for steels.
5. The load-based thermal modification factor is finally calculated as KCP ¼ 10kP .
The following two aspects should be noted. First, Step-1 through Step-5 is
the basic procedure for obtaining a load-based thermal modification factor. In
practice, a thermal modification factor from two different load levels can be esti-
mated based on the same procedure. Then compare the two factors obtained,
and the factor with a larger value can be used for conservative purposes. Second,
the procedure drawn from Step-1 through Step-5 and the geometrical relationship
shown in Fig. 9 indicates that DPE ¼ KCP DPR , but it does not explicitly indicate how
to modify the loading data PðtÞ based on the obtained thermal modification
factor KCP . However, it should be noted that the definitions of the load range at
room temperature and the equivalent temperature are DPR ¼ PRP  PRV and
DPE ¼ PEP  PEV , respectively. Clearly, DPE ¼ KCP DPR can be directly obtained
from PEP ¼ KCP PRP and PEV ¼ KCP PRV . The subscripts P and V represent the
peak and the valley in a cycle.

Strain-Based Thermal Modification Factor


A similar thermal modification factor procedure can be developed for the strain-
based method. This method requires both the measured strain at the location near
the notch (5–10 mm away) and the fatigue strain-cycle (E-N) of the un-notched
material. However, there are two important differences between the load-based
method and the strain-based method.
First, the fatigue E-N curves are usually not linear because both low-cycle
and high-cycle fatigue data are collected. The total strain approach is often used to
unify the low-cycle and the high-cycle data, which can be described using elastic
SEITZ ET AL., DOI 10.1520/STP159820160084 385

strain-life and the plastic strain-life, respectively. The elastic strain and plastic strain
are measured or estimated from a stable cyclic stress-strain hysteresis loop [8]. Tra-
ditionally, the elastic strain-cycle and the plastic strain-cycle are usually described
using the Wöhler curve for high-cycle fatigue and the Manson-Coffin curve for
low-cycle fatigue. The nonlinear behavior of the total strain range cycles to failure
in a log-log plot is schematically shown in Fig. 10. Even though the curves are non-
linear, the strain-based thermal modification factor can still be easily calculated as
KCe ¼ 10ke , which is similar to the formulae used in the load-based method. The
only difference is that Point C is obtained by nonlinear extrapolation rather than
linear extrapolation as is the case with the load-life curve.
Second, a significant difference of the strain-based approach from the load-
based method is that a strain amplification effect has to be introduced to link the
measured strain and the strain at the weld notch, which is supposed to be approxi-
mately represented by the strain-cycle data of the material in terms of fatigue life. It
should be noted that the failure often occurs in either the welds or the heat-affected
zones, which cannot be directly measured. Instead, the strain is often measured at a
location that is about 5 mm to 10 mm away from the welded notch. Fig. 11 schemati-
cally shows the failure location and the location where the nominal strain is
measured.
In order to quantify the strain amplification effect, the strain amplification
factor (SAF) is defined in this paper, and the analysis of elastic-plastic stress-strain
relationship is required to find the solution. Fig. 12 schematically shows the relation-
ship between the cyclic stress and strain. It is assumed that the same stress-strain

FIG. 10 Thermal modification factor for strain-based cold testing.


386 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 11 Schematic of the locations where the nominal strain is measured and where
failure occurs due to the equivalent local strain.

relationship is applicable to the base material and the welds. It should be noted that
welds and their heat-affected zones have different properties than the parent or
base material, but in practice, the latter is often used as far as the linear elastic

FIG. 12 The representation of the SAF Ke in the cyclic stress-strain curve.


SEITZ ET AL., DOI 10.1520/STP159820160084 387

mechanical behavior is concerned. The simplest definition of the SAF can be


expressed in Eq 2:
Kf ¼ e=e: (2)

Again, e is the strain measured using a strain gage, and e is the local strain,
from which the associated strain range De can be estimated by the total strain-
based strain-cycle as expressed in Eq 3 [8]:
ef 0  d  c
De ¼ 2 2Nf þ 2ef 0 2Nf (3)
E
where rf 0 and ef 0 are the fatigue strength coefficient and fatigue ductility coefficient,
and d and c are the fatigue strength component and fatigue ductility exponent. Nf is
the number of cycles to failure, which is obtained from the fatigue bench test.
The factor Kf is often determined from a specific SAF determination test with a
certain load and a certain temperature but applied to the RLDA with various loads
and temperatures. It should be noted that Kf is generally a function of both loading
amplitude and operating temperature. The amplitude effects on the value of the
amplification factor are schematically illustrated in Fig. 12, which shows a cyclic
elastic-plastic stress-strain curve. The cyclic stress-strain curve shows a linear elastic
part and a nonlinear plastic part. Without losing generality, it is assumed here that
the measured strain is elastic in the analysis. Based on the relative strain severity
during real RLDA loading and the SAF test, the following four scenarios can be
created:
• Scenario-I: The equivalent strain at the notch under SAF loading and RLDA
loading are all located in the elastic domain, say Kf ¼ e1 =e1 ¼ Kf 1 , as shown
in Fig. 12.
• Scenario-II: The equivalent strain at the notch under SAF loading is elastic but
that under RLDA loading is plastic, Kf ¼ e2 =e2 ¼ Kf 2 .
• Scenario-III: The equivalent strain at the notch under SAF loading is plastic
but that under RLDA loading is elastic.
• Scenario-IV: The equivalent strains at the notch under both SAF loading and
RLDA loading are plastic.
For Scenario-I, as long as the equivalent strain at the notch is in the elastic
range, the SAF will be a constant. Therefore, to estimate the equivalent strain expe-
rienced at the weld notch, the RLDA load can simply be modified by multiplying
the measured strain history with the single SAF. In this case, Kf ¼ e1 =e1 can also be
called the strain proportionality factor. Clearly, Scenario-I is preferred in terms of
testing and analysis because of its linearity. In Scenario-II, the strain at the notch
under the RLDA is underestimated because Kf is calibrated in the elastic domain.
In the plastic part, the plastic strain increases more drastically than the stress or
load. Clearly, the damage estimated from Scenario-II is non-conservative. Scenario-
III is just the opposite of Scenario-II, and the predicted damage from RLDA would
be over-conservative. Scenario-IV is more complex, and the conservativeness
depends on the details of the RLDA loading and the SAF calibration.
388 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

It should be noted that the conversion between stress range and amplitude
should be performed correctly. Finally, it is noted that the SAF is different from the
notch factor as defined in Neuber’s rule and other similar definitions, which deals
with the relationship between the notched specimen and the separate un-notched
specimen [17,18].
Overall, for the strain-based method, the stain amplification factor and the
thermal modification factor can be combined to form a new factor, K ¼ KCe Kf ,
which can be used for RLDA data modification and eventual damage and life
assessment.

LOADING TRANSFORMATION USING DAMAGE EQUIVALENCY PRINCIPLE


If the temperature variation is not significant and the material fatigue behavior is
not very sensitive to temperature, a specific temperature level can be approximately
used for both temperature response characterization and rainflow counting. The
hanger rods at bends and the brackets on body sides in a vehicle exhaust system are
such examples. However, for some of the exhaust applications with variable temper-
atures, the use of the isothermal assumption can lead to significant errors—in
particular for the pipes that have direct contact with the gas flow.
For anisothermal fatigue, the preferred approach in the automotive industry is
to use the fatigue curves obtained from isothermal tests at several different tempera-
ture levels. Then, the temperature in question is interpolated from the tested fatigue
curves. However, a critical issue in life assessment of anisothermal applications is
how to count the cycles for the loads with a superimposed variable temperature
profile. In order to take the temperature effect into account, several methods have
been proposed to handle the temperature variation [19,20]. In these methods, the
maximum temperature over the duration of the fatigue cycle is determined for a
given cycle to make the life assessment conservative; hence, the worst-case scenario
is actually considered.
An effective cycle-counting procedure has been recently proposed [21] for load-
ing data under variable amplitude temperature and mechanical loadings. The pro-
cedure first converts the loading data to an equivalent form at a reference
temperature using the damage equivalency principle. This can be done by simply
multiplying a temperature-dependent factor to account for the temperature effect.
Subsequently, the standard rainflow counting method can be used to count the
cycles. The key to the procedure is briefly described as follows.
For a fatigue event under constant stress range Dri and constant temperature
Ti , fatigue damage D accumulated in a cycle can be expressed as Eq 4:

D½DrðTi Þ ¼ 1 Nf ¼ ðDri Þm f ðTi Þ (4)

where Ti is a temperature-dependent constant and f ðTi Þ is a temperature-related


function; m ¼ 1=b and b is defined in Eq 1. The subscript i represents the tempera-
ture level i; f ðT Þ function in power law and Arrhenius forms are often
 used [21]. The
1=m
S-N curve corresponding to this is Nf ¼ ðDrÞm f ðT Þ1 or Dr ¼ Nf f ðT Þ1=m .
SEITZ ET AL., DOI 10.1520/STP159820160084 389

By applying the damage equivalency principle and under the isothermal assumption,
constant stress range Dri at temperature Ti can be transformed into the temperature
level i associated stress range DrRi at the reference (R) temperature TR . Then, the
damage equivalency D½DrðTi Þ ¼ D½DrðTRi Þ can be expressed in Eq 5:

ðDri Þm f ðTi Þ ¼ ðDrRi Þm f ðTR Þ (5)

or Eq 6:

DrRi ¼ Dri ½f ðTi Þ=f ðTR Þ1=m : (6)

With the help of DrRi ¼ rRiP  rRiV and Dri ¼ riP  riV , Eq 7 can be
derived:

rRi ¼ ri ½f ðTi Þ=f ðTR Þ1=m : (7)

Subscripts P and V represent the peak and valley in a cycle, respectively. Therefore, a
loading profile including the peaks, valleys, and any data points in between can be
converted to an equivalent form with Eq 7 based on the temperature profile and the
reference temperature.
Similarly, for another fatigue event under constant stress range Drj and con-
stant temperature Tj , based on the equivalency damage principle, the stress range
time history can be transformed into an equivalent form at the reference tempera-
   1=m
ture TR : rRj ¼ rj f Tj =f ðTR Þ . From, rRi in Eq 7 and rRj , it is clear that the
framework of stress transformation at different temperature T are the same. There-
fore, for a general stress loading profile with variable temperature, the general form
of the transformation can be expressed in Eq 8:

rR ðt Þ ¼ rðt Þg ½T ðt Þ; TR  ¼ rðt Þff ½T ðt Þ=f ðTR Þg1=m (8)

where g ½T ðt Þ; TR  ¼ ff ½T ðt Þ=f ðTR Þg1=m . With the transformation technique, the orig-
inal stress history rðt Þ with variable temperature history T ðt Þ can be transformed to
stress history rR ðt Þ at a constant reference temperature TR . In damage assessment, the
rainflow counting can be directly conducted on the transformed loading profile, and
the damage can then be calculated from the fatigue S-N curve at the reference
temperature.
Fig. 13a schematically shows that transformation of a cyclic loading time history
rðt Þ with a constant loading amplitude and stress range Dr ¼ riP  riV at a con-
stant temperature Ti into a damage equivalent cyclic loading time with a constant
amplitude loading amplitude stress range DrRi ¼ rRiP  rRiV at a constant tem-
perature TR . This simple and specific example is the one from which Eqs 5–7 are
derived. The application of the procedure for variable loading profile and a constant
temperature is schematically shown in Fig. 13b. The procedure for both variable
loading and temperature profiles is shown in Fig. 13c. It should be noted that for the
r  t data at temperature TR , there are infinite sets of DrR that generate the same
390 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

FIG. 13 Transformation of stress histories into their equivalent form at a reference


temperature TR : (a) constant amplitude stress at a constant temperature Ti ,
(b) variable amplitude stress at a constant temperature Ti , (c) variable stress
and temperature histories, and (d) the linear relationship between the original
stress input and the resulting equivalent stress output.

(a) (b)

(c) (d)

equivalent damage, such as the loading waves as highlighted on the insert shown on
the right middle of Fig. 13a. However, there is only one set of loading waves satisfy-
ing the transformation, Eq 7, with the proportional factor g ½T ðt Þ; TR . The unique
linear relationship among the original input stress and the output stress at two tem-
peratures Ti and Tj is schematically shown in Fig. 13d.

DAMAGE MECHANISMS UNDER COMBINED VARIABLE LOADING AND


TEMPERATURE
The damage accumulation under variable load (stress) and temperature can be
understood by studying the stress-temperature diagram shown in Fig. 14a, in which
SEITZ ET AL., DOI 10.1520/STP159820160084 391

FIG. 14 (a) The loading path represented in a stress-temperature diagram, (b) stress
variation caused by temperature change, and (c) the equivalent stress due to
the temperature variation.

(a) (b)

(c)

a half-cycle is presented with several possible loading paths. Without losing general-
ity, we assume that r2 > r1 and Tj > Ti . In practice, the loading-path dependency
of the damage is essentially unknown. Theoretically, as well as being shown as fol-
lows, the path dependency seems to have intrinsic characteristics of the variable
load and temperature. Among infinite loading paths, the following three typical
loading paths are the most significant in both mechanisms’ description and mathe-
matical operation: (1) Path-I: the loading path ABD, (2) Path-II: the loading path
ACD, and (3) Path-III: the straight path AD. Path-I undergoes a temperature
increase (i.e., the path AB) at a constant stress before it undergoes a stress increase
Dr ¼ r2  r1 at the temperature Tj (i.e., the path BD). The Path-II has a similar
loading path but in an opposite order: stress increase Dr ¼ r2  r1 (path AC) at
the temperature Ti , followed by a temperature increase DT ¼ Tj  Ti (path AB).
For both Path-I and Path-II, the stress change part at a constant temperature can
be well-described by the isothermal fatigue theory, in which a fatigue S-N curve
is sufficient to describe the fatigue behavior. However, the fatigue damage caused
by temperature change at a given stress is different from the deformation-
based mechanism and requires further data collection to build up a new database.
392 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

Fig. 14b illustrates the transformed stress profile as obtained from the transforma-
tion technique and the variable temperature profile. Path-III shows a combined and
simultaneous temperature and stress increase, and it is more representative of the
cases as experienced in real applications. Scenario-III will be used here to demon-
strate the effects of the combined variable stress and temperature on the damage
accumulation.
Based on Eq 5 and Eq 6, the damage per cycle at temperature Ti and Tj is
   
DðTi Þ ¼ ðDrÞm f ðTi Þ and D Tj ¼ ðDrÞm f Tj , respectively. The accumulated
damage per cycle D ¼ ðDrÞm f ðT Þ can be considered as a result of the integration of
a damage growth rate over the domain [0, Dr]. Similar treatment has been studied
for fatigue crack growth [22,23]. The differential form of the damage growth rate at
a constant temperature T can be written as:

dD
¼ mðDr Þm1 f ðT Þ: (9)
dr
Integrating Eq 9, with D ¼ 0 at r ¼ 0, leads to D ¼ ðDrÞm f ðT Þ at r ¼ Dr. To
investigate the loading path effect, the power law f ðT Þ ¼ T c and c ¼ 1 is assumed in
this study. From f ðT Þ ¼ T c , it can be seen that c > 0 indicates an increased damage
rate with temperature and that is usually the case for fatigue at elevated temperature.
From Fig. 14a, it is found that the temperature and the stress increment are not inde-
pendent, rather there is a linear relationship that can be expressed as in Eq 10:
 
Tj  Ti 
T¼ r þ Ti (10)
Dr
Clearly, T ¼ Ti when r ¼ 0 and T ¼ Tj when r ¼ Dr; therefore, the new
crack growth equation under the combined stress and temperature loading can be
expressed in Eq 11:
 
dD m Tj  Ti
¼ ðDr Þm þ mT1 ðDr Þm (11)
dr Dr

Integrating Eq 11 with D ¼ 0 at r ¼ 0 leads to:


mTj þ Ti
D ¼ ðDrÞm (12)
mþ1

at r ¼ Dr. When Ti ¼ Tj , Eq 12 is reduced to the damage per cycle at a constant


m m
temperature;
  when mTi < Tj ,  D  DðT ¼ Ti Þ ¼ mþ1 ðDrÞ Tj  Ti > 0 and
D  D T ¼ Tj ¼ ðDrÞ Ti  Tj < 0. Therefore, the damage caused by the linear
path is between that caused by the same stress range at the highest temperature and
that caused by the same stress range at the lowest temperature.
The damage mechanism under the combined variable load and temperature can
be schematically shown in Fig. 14c. According to Miner’s.linear damage rule,
. the dam-
j
age caused by a stress range Dr at Ti and Tj are Di ¼ 1 Nfi and Dj ¼ 1 Nf , respec-
j
tively; Dj > Di because Nf < Nfi , as shown in the logðDrÞ logðNÞ curve; see Point A
SEITZ ET AL., DOI 10.1520/STP159820160084 393

and Point B in Fig. 14c. According to Eq 12 and Fig. 14c, the variation in temperature
. at
ij ij
the same stress range results in a new cycles to failure Nf , damage Dij ¼ 1 Nf , and
Dj > Dij > Di . Based on the loading transformation technique shown earlier, the
equivalent stress ranges will be DrRi and DrRj , at the reference temperatures Ti and Tj ,
respectively. Both transformed stress ranges result in the same damage at their respec-
tive reference temperatures because they are equivalent according to the load transfor-
mation technique as described in the earlier section on the thermal modification factor
for cold testing.

DEVIATION OF “EQUIVALENT” TEMPERATURE FOR HOT TESTING


In hot testing, the ideal scenario would be that the tests are conducted at a wide
range of temperature levels to accurately interpret the temperature effects. However,
hot RLDA, hot calibration, and hot bench testing are expensive and time consum-
ing. In order to reduce the cost, the amount of tests needs to be reduced. The tem-
perature profile from RLDA is predetermined, and it is unique for a run for a given
vehicle, operating condition, and road condition. Definitely, different runs will
result in different loading profiles, which are random and can be described in a sta-
tistical manner. Hot calibration at different temperatures is relatively easy. Hot
component bench testing is the area where the testing cost could increase. It would
be costly and time consuming to test at two and more than two temperature levels
because 13 component tests need to be conducted at each temperature level. There-
fore, testing at a single temperature level is preferred. Then, which temperature
should be used? Room temperature definitely would not be considered because it
does not reflect the hot RLDA characteristics. The peak temperature can reflect the
hot RLDA behavior but it would be over-conservative. Therefore, an “equivalent”
temperature concept would be helpful. Clearly, the equivalent temperature should
be derived from the damage equivalency principle. An equivalent temperature can
be described by equalizing the damage values made by the load data at variable tem-
peratures for a given period of time and the damage by the same loading at a con-
stant “equivalent” temperature for the same given period of time.
Because the damage accumulation for a load with variable temperature has
been determined from the equivalent damage principle, as shown earlier, the dam-
age can be calculated using the Miner’s linear damage rule for the calculated cycles
for different ranges using the rainflow counting and the fatigue S-N database at var-
ious temperatures. Generally speaking, the equivalent temperature is not necessarily
a constant. It could be a variable temperature as long as the accumulated damage is
the same. However, a constant temperature is preferred in component bench test-
ing. The high-level mathematical formula to determine the equivalent temperature
using the damage principle can be expressed as:
D½rðt Þ; T ðt Þ ¼ D½rR ðt Þ; TR ðt Þ ¼ D½rðt Þ; Te  (13)
The principle is simple and the numerical calculation is trivial and straight-
forward, but in some simplified cases, some analytical solutions can be obtained.
394 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

Generally, the temperature and the loading are coupled, and the procedure to find
an equivalent temperature has to be determined numerically using Eq 13. A much-
simplified formula can be derived for the special case: the temperature effects are
symmetrical with respect to the statistical mean of the load, so that the overall con-
tribution from the load variation at variable temperature conditions can be can-
celled out. Usually, this assumption can be reinforced with the independent and
random load, which means the occurrence of one does not affect the other. This is
reasonable for some applications—for example, the vehicle vibration is caused by
the road condition and the temperature is caused by the engine condition. Listed
next is a procedure to demonstrate how to derive the equivalent temperature
approach for these simplified cases.
Without losing generality, a constant load is assumed here to simplify the analy-
sis and the load-temperature independency assumption. From Eq 4, the damage
made during a unit time can be written as dD ¼ ðDPÞm f ½T ðt Þndt, with n ¼ N=t0
representing the cycles N per unit time t0 . Therefore, the total damage during the
Ð Ðt
time period of t0 can be integrated as D ¼ dD ¼ 00 ðDPÞm f ½T ðt Þndt. It should be
noted that DP is also a function of temperature T. An alternative equivalent tempera-
ture profile Te ðt Þ that can make the same amount of damage is expressed in Eq 14:
ð ð t0
De ¼ dD ¼ ðDPÞm f ½Te ðt Þndt (14)
0

Assuming that the load and the temperature are weakly dependent on each
other and the load range variation is not significant over time, then Eq 14 can be
further simplified and it becomes Eq 15:
ð t0 ð t0
f ½T ðt Þdt ¼ f ½Te ðt Þdt (15)
0 0

Further assuming that the equivalent temperature Te is a constant, the formula


can be expressed as:
ð t0
1
Te ¼ f 1 f ½T ðt Þdt (16)
t0 0

where f 1 is the inverse function of f ½T ðt Þ.


It should be noted that load-dependent temperature is often used in some
applications. For example, in low-cycle thermal-mechanical fatigue analysis, Taira
[24] recommends the peak temperature as an equivalent temperature if the operat-
ing temperature is high enough to cause creep while the mean temperature as an
equivalent temperature when the operating temperature is relatively low.

Examples and Results


The following three case studies are used to demonstrate the procedures devel-
oped in the previous text section on consideration of temperature effects in
SEITZ ET AL., DOI 10.1520/STP159820160084 395

product validation, with the emphasis on handling of the temperature effects:


(1) the estimation of the value of the thermal modification factor for a welded
exhaust component at a constant temperature environment, (2) the estimation of
the value of a non-welded exhaust component at a variable temperature environ-
ment, and (3) the damage calculations from the transformed “equivalent” loading
profile at two different reference temperatures and the calculation of load-temperature
independent equivalent temperature and load-temperature-dependent equivalent
temperature.

CASE-1: CONSTANT SKIN TEMPERATURE FOR A WELDED GASOLINE


EMISSION CONTROL DEVICE
Fig. 15a shows the test setup of a component bench test for gasoline emission
control. The component design is the cone-to-pipe structure with a 360 weld
connection. Test samples are made of 409 stainless steel welded with 409 weld
wire. The nominal material thicknesses at the crack area are 1.6 mm (cone)
and 1.75mm (pipe). The room-temperature component bench test uses a nominal
strain of 501 l in/in magnitude. Multiple samples are tested and their bench lives
result in a mean of 175,035 cycles. Fig. 15b shows the fatigue failure location at the
360 weld. A second batch of samples of the same design are bench tested with
the same oscillating load but with a constant skin temperature of 450 C. The
bench lives in the second test result in a mean of 31,463 cycles. Based on the pro-
cedure described earlier on the thermal modification factor for cold testing and
b ¼ 1=3, the estimated SAF KCP is 1.772. It indicates that, to compensate for the
temperature effects, the cold RLDA load data should be multiplied by 1.772 to cal-
culate the damage as would be experienced by the vehicle exhaust during hot
operation.

FIG. 15 (a) Component bench test setup of a gasoline emission control device and
(b) fatigue crack in the component during the bench test.

(a) (b)
396 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

CASE-2: VARIABLE TEMPERATURE FOR NON-WELDED SILENCING DEVICE


Silencing devices are used to reduce exhaust noise, and many silencing devices have
internal design features that cause stress concentration. Fig. 16a shows a test setup of
the component bench test for a silencing device. The test samples are made of 409
stainless steel. The nominal material thickness at the crack area is 1.4 mm. The
room-temperature component bench test uses a nominal strain of 444 l in/in (mag-
nitude). Multiple samples are tested, and their bench lives result in a mean of
237,591 cycles. Fig. 16b shows a fatigue failure location at the tuning dimple on the
inner tube, which is close to the head face.
A second batch of samples of the same design are bench tested with the same
oscillating load but under specified temperature cycles (5 min at 425 C, 7 min at
525 C, and 15 min at 600 C). The bench lives in the second test result in a mean of
23,063 cycles. Based on the procedure described earlier and b ¼ 1=3, the estimat-
ed SAF KCP is 2.176. This indicates that to compensate for the temperature effects,
the cold RLDA load data should be multiplied by 2.176 to calculate the damage as
would be experienced by the vehicle exhaust during hot operation.

CASE-3: LOAD TRANSFORMATION TECHNIQUE AND EQUIVALENT


TEMPERATURE ESTIMATION
A pair of strain time data and temperature time data are shown in Fig. 17a and b,
respectively. The unit of the temperature shown in Fig. 17b is Kelvin. The trans-
formed strain profiles for two reference temperatures, that is, peak temperature (c)
and room temperature (d), by taking the temperature-dependent fatigue into account.
In the calculation, b ¼ 1=3 for the slope of the fatigue S-N curve, and c ¼ 1 for the
exponent in the power law f ðT Þ ¼ T c are assumed. Additionally, the strain-cycle E-N
curve at peak temperature is used for the transformed strain-time data with the peak
temperature as the reference temperature, and strain-cycle E-N curve at room
temperature is used for the transformed strain time data with the room temperature as

FIG. 16 (a) Component bench test setup of a silencing device, and (b) fatigue crack in
the silencing device component bench test.

(a) (b)
SEITZ ET AL., DOI 10.1520/STP159820160084 397

FIG. 17 (a) The original strain-time data profile, (b) the temperature-time data profile,
(c) the transformed strain-time profile using the reference of the peak
temperature, and (d) the transformed strain-time profile using the room
temperature.

the reference temperature. Clearly, as compared to the original strain profile, the
transformed strain profile shows lower strain amplitude when the peak temperature
is used as the reference temperature and higher amplitude when the room tempera-
ture is used as the reference temperature. This is understandable because higher
temperature results in higher damage at the same strain; therefore, to keep the same
damage, the strain must be lowered. Furthermore, the relative difference in damages,
2ðDPeak  Droom Þ=ðDPeak þ Droom Þ, estimated from the two transformed profiles,
Fig. 17c and d, at the two respective reference temperatures (i.e., the peak temperature
and the room temperature) is only 2:975  104 %, which is mainly caused by digital
rounding-out and can be negligible in most practical applications. This clearly demon-
strates the correctness and effectiveness of the loading transformation technique.
Even though the damage equivalency procedure is directly derived from simple
constant amplitude load and two constant temperatures, the results show a negligi-
ble difference in damage in the specific example shown. This indicates the sound-
ness of the approach in more general applications. It should be noted that the
difference in two reference temperatures is large (i.e., peak temperature and room
temperature). Therefore, the procedure can be stated as reference temperature inde-
pendent or reference temperature insensitive, which is a desired feature of any good
theoretical framework.
The equivalent temperature Te as calculated from Eq 13 for the strain and tem-
perature profiles shown in Fig. 17 is 664.18K. In the calculation, the equivalent
398 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

temperature is found from trial and error and an iteration process. In this case, the
coupling effect of the strain and temperature is fully accounted for. Assuming that
the strain and the temperature are decoupled and completely independent of each
other, Eq 16 can be used, and the estimated equivalent temperature Te is 661.86 K.
The mean temperature simply calculated by averaging the highest (peak) and the
lowest temperature is 664.67 K. The three equivalent temperatures estimated with
the three methods are similar.
The three definitions of the equivalent temperatures for the example result in
three different values of the temperatures. However, the difference is relatively small
for this specific case. The difference is also negligible as compared with the mea-
surement error in testing. Clearly, the damage equivalency and load-temperature-
dependent approach, Eq 13, is more general and easy to implement; therefore, it is
strongly recommended. However, in some cases, when the damage contribution
caused by loading variation respective to the given temperature profile is cancelled
out, the simplified formula, Eq 16, can be used.

Discussion
First, although the complete application of the procedures developed in this paper
has not been fully performed at this moment, the fundamental techniques and
issues related to achieving reliable and effective product durability and reliability
performance have been identified and subsequently solved. The major points made
in this paper are essentially theoretical framework, principle, and procedures, but
these developments can shed light on the cold-testing and hot-testing planning,
data acquisition, and data analysis.
Second, load, stress, and strain are different parameters measuring the degree
and severity in loading. They are related through the geometry and the constitutive
laws of materials. However, in this paper, when procedures of fatigue damage and
life assessment are mentioned, these three parameters are interchangeably used. In
other words, either load, stress, or strain can be used for the description of a dam-
age event as long as these measures are consistently used (i.e., load-cycle fatigue
curve is used for load-time history, stress-cycle curve is used for stress-time data,
and strain-cycle curve is used for strain-time data).
Third, both amplitude and range are commonly used in fatigue modeling and
testing. The same situation also happens to cycle and reversal. However, mixed use
of these definitions is often the common source of errors. For example, in RLDA,
calibration testing strain amplitude is directly measured while strain or stress range
is often used. For a correct damage and life assessment, a proper conversion must
be conducted. Additionally, the relationship between the parameters in a monoton-
ic stress-strain curve and the corresponding cyclic stress-strain curve is another
area where errors can be made. These delicate definitions are not explicitly distin-
guished in the text of this paper by assuming that the readers are aware of their
differences and importance.
SEITZ ET AL., DOI 10.1520/STP159820160084 399

Finally, several assumptions are made in this paper to make the theoretical
framework complete. Examples of these assumptions include:
• The exponents b in Eq 1 and m in Eq 4 are assumed to be constants at all the
temperature levels. In reality, these values might vary. An average value could
be used if their variations are not significant.
• No mean stress correction is accounted for in the fatigue models.
• Sample size effect is not considered in the thermal modification factor deter-
mination in the first two examples shown in the previous section.

Conclusions
Durability and reliability assessment of vehicle exhaust systems is a challenging task
because of the complex mechanical and thermal loading during operation. Cold
testing in terms of RLDA, calibration, and component bench tests is the most com-
monly used cost-effective product validation method. However, the temperature
effects must be properly considered to compensate for the information loss during
cold testing. Hot testing is a relatively new research direction, and significant tech-
nical barriers as well as cost issues exist on the road toward practical applications.
How to develop a thermal modification factor to compensate for the cold testing
and how to lay out a theoretical framework for hot testing remain one among the
most critical technical challenges. In this paper, the fundamental issues of both cold
testing and hot testing are addressed, and the following conclusions can be drawn
from this research.
1. Based on the temperature range, the material failure mechanism, and analysis
approach, the thermal and mechanical loadings are categorized into three gen-
eral groups: isothermal fatigue, anisothermal fatigue, and high-temperature
TMF. Anisothermal fatigue, which usually consists of high-frequency mechan-
ical loading and low-frequency thermal cycling, is found to be the dominant
mechanism for the vehicle exhaust system being studied.
2. The current cold-testing-based product durability and reliability validation
procedure is reviewed, and a simple formula for a thermal modification factor
is developed to account for the temperature effects, which can be reflected and
estimated from the mean life comparison between the cold and hot bench tests
of the same design under the same loading condition.
3. Based on a damage equivalency principle, a procedure is developed to trans-
form the load (stress, strain) histories into an equivalent form at a reference
temperature by taking the temperature history into account—real temperature
data linked from point to point with strain data. With the transformed load
data at a reference temperature, the rainflow algorithm can be used for cycle
counting. Two “equivalent” temperature measures are also developed for com-
ponent bench testing, which could provide minimum sample size and the best
cost-saving strategy.
4. Several case studies have demonstrated the effectiveness of the thermal modifica-
tion factor in cold-testing applications and in the damage equivalency-based data
transformation technique in life assessment under anisothermal fatigue loading.
400 STP 1598 On Fatigue and Fracture Test Planning, Test Data Acquisitions and Analysis

References

[1] Lin, S., “Temperature Effect in Exhaust System Fatigue Prediction,” SAE Technical
Paper: 2011-01-0783, SAE International, Warrendale, PA, 2011, http://dx.doi.org/10.4271/
2011-01-0783

[2] Greuter, E. and Zima, S., Engine Failure Analysis: Internal Combustion Engine Failures
and Their Causes, SAE International, Warrendale, PA, 2012.

[3] Santacreu, P. O., Faivre, L., and Acher, A., “Life Prediction Approach for Stainless Steel
Exhaust Manifold,” SAE Int. J. Passeng. Cars-Mech. Syst., Vol. 5, No. 2, 2012, http://
dx.doi.org/10.4271/2012-01-0732

[4] Chinouilh, G., Santacreu, P. O., and Herbelin, J. M., “Thermal Fatigue Design of Stainless
Steel Exhaust Manifolds,” SAE Technical Paper 2007-01-0564, SAE International,
Warrendale, PA, 2007.

[5] Wei, Z., Luo, L., Lin, B., Yang, F., Konson, D., Ellinghaus, K., Pieszkalla, M., Avery, K., Pan,
J., and Engler-Pinto, C., “Hold-Time Effect on Thermal-Mechanical Fatigue Life and Its
Implications in Durability Analysis of Components and Systems,” Mater. Perform.
Charact., Vol. 4, No. 2, 2014, pp. 198–217, http://dx.doi.org/10.1520/MPC20140032

[6] Wei, Z., Qu, Y., Jiang, D., Luo, L., Hamilton, J., Ellinghaus, K., and Pieszkalla, M.,
“Probabilistic Isothermal, Anisothermal, and High-Temperature Thermo-Mechanical
Fatigue Life Assessment and CAE Implementation,” SAE Technical Paper, 2016-01-0370,
SAE International, Warrendale, PA, 2016.

[7] Lee, Y. L., Pan, J., Hathaway, R., and Barkey, M., Fatigue Testing and Analysis: Theory and
Practice, Elsevier Butterworth-Heinemann, Boston, 2005.

[8] Manson, S. S. and Halford, G. R., Fatigue and Durability of Metals at High Temperatures,
ASM International, Materials Park, OH, 2009.

[9] Wei, Z., “Characterization of Materials for Exhaust Systems under Combined Mechanical
and Corrosive Environment,” SAE Technical Paper, 2013-0102420, SAE International,
Warrendale, PA, 2013, http://dx.doi.org/10.4271/2013-01-2420

[10] Gallerneau, F., Nouailhas, D., and Chaboche, J. L., “A Fatigue Damage Model Including
Interaction Effects With Oxidation and Creep Damages,” Fatigue ’96: Proceedings of
The Sixth International Fatigue Congress, G. Lütjering, and H. Nowack, Eds., Berlin,
Germany, May 6–10, 1996, pp. 861–866.

[11] Neu, R. W. and Sehitoglu, H., “Thermomechanical Fatigue, Oxidation and Creep: Part I.
Damage Mechanisms,” Metall. Trans. A, Vol. 20, No. 9, 1989, pp. 1755–1767.

[12] Neu, R. W. and Sehitoglu, H., “Thermomechanical Fatigue, Oxidation and Creep: Part II.
Life Prediction,” Metall. Trans. A, Vol. 20, No. 9, 1989, pp. 1769–1783.

[13] Halfpenny, A., Anderson, R., and Lin, X., “Isothermal and Thermo-Mechanical Fatigue
of Automotive Components,” SAE Technical Paper: 2015-01-0548, SAE International,
Warrendale, PA, 2015, http://dx.doi.org/10.4271/2015-01-0548

[14] Voltenburg, R. K., “Alternative Component Validation Methodology Study,” Thesis, Bach-
elor of Science in Mechanical Engineering, Kettering University, Flint, MI, 2010.
SEITZ ET AL., DOI 10.1520/STP159820160084 401

[15] BS 7608, Code of Practice for Fatigue Design and Assessment of Steel Structures, British
Standards Institution, London, 1993.

[16] DNV-RP-C203, Recommended Practice for Fatigue Design of Offshore Steel Structures,
Det Norske Veritas (DNV), Høvik, Norway, 2010.

[17] Neuber, R. E., “Theory of Stress Concentration for Shear-Strained Prismatical Bodies With
Arbitrary Nonlinear Stress-Strain Law,” J. Appl. Mech., Vol. 28, No. 4, 1961, pp. 554–550.

[18] Topper, T. H., Wetzel, R. M., and Morrow, J., “Neuber’s Rule Applied to Fatigue of
Notched Specimens,” J. Mater., Vol. 4, No. 1, 1969, pp. 200–209.

[19] Nagode, M. and Hack, M., “An Online Algorithm for Temperature Influenced Fatigue Life
Estimation: Stress-Life Approach,” Int. J. Fatigue, Vol. 26, No. 2, 2004, pp. 163–171.

[20] Kang, H. T., Lee, Y. L., Chen, J., and Fan, D., “A Thermo-Mechanical Fatigue Damage
Model for Variable Temperature and Loading Amplitude Conditions,” Int. J. Fatigue,
Vol. 29, Nos. 9–11, 2007, pp. 1797–1802.

[21] Wei, Z., Lin, S., Luo, L., Yang, F., and Konson, D., “A Thermal-Fatigue Life Assessment
Procedure for Components under Combined Temperature and Load Cycling,” SAE Tech-
nical Paper 2013-01-0998, SAE International, Warrendale, PA, 2013, http://dx.doi.org/
10.4271/2013-01-0998

[22] Wei, Z. and Dong, P., “Fatigue Life Assessment of Welded Structures,” Eng. Fract. Mech.,
Vol. 77, No. 15, 2010, pp. 3011–3021.

[23] Wei, Z. and Dong, P., “A Generalized Cycle Counting Criterion for Arbitrary Multi-Axial
Fatigue Loading Conditions,” J. Strain Anal. Eng., Vol. 49, 2014, pp. 325–341.

[24] Taira, S., “Relationship between Thermal Fatigue and Low-Cycle Fatigue at Elevated Tem-
perature,” Fatigue at Elevated Temperatures, ASTM STP520, A. E. Carden, A. J. McEvily,
and C. H. Wells, Eds., ASTM International, West Conshohocken, PA, 1973, pp. 88–101, http://
dx.doi.org/10.1520/STP38831S
Wei | Nikbin | McKeighan | Harlow
ASTM INTERNATIONAL
Selected Technical Papers

Fatigue and Fracture


Test Planning, Test

Acquisitions and Analysis


Fatigue and Fracture Test Planning, Test Data
Data Acquisitions
and Analysis

STP 1598
Editors:
Zhigang Wei
Kamran Nikbin
Peter C. McKeighan
Gary D. Harlow

ASTM INTERNATIONAL
Helping our world work better
ASTM International

ISBN: 978-0-8031-7639-3
Stock #: STP1598
www.astm.org

You might also like