Fatigue and Fracture of Metals and Alloys Numerical and Experimental Study

Download as pdf or txt
Download as pdf or txt
You are on page 1of 370

Special Issue Reprint

Fatigue and Fracture


of Metals and Alloys
Numerical and Experimental Study

Edited by
Jaroslaw Galkiewicz, Lucjan Śnieżek and Sebastian Lipiec

www.mdpi.com/journal/materials
Fatigue and Fracture of Metals and
Alloys: Numerical and Experimental
Study
Fatigue and Fracture of Metals and
Alloys: Numerical and Experimental
Study

Editors
Jaroslaw Galkiewicz
Lucjan Śnieżek
Sebastian Lipiec

MDPI • Basel • Beijing • Wuhan • Barcelona • Belgrade • Manchester • Tokyo • Cluj • Tianjin
Editors
Jaroslaw Galkiewicz Lucjan Śnieżek Sebastian Lipiec
Kielce University of Military University and Kielce University of
Technology Technology Technology
Poland Poland Poland

Editorial Office
MDPI
St. Alban-Anlage 66
4052 Basel, Switzerland

This is a reprint of articles from the Special Issue published online in the open access journal Materials
(ISSN 1996-1944) (available at: https://www.mdpi.com/journal/materials/special_issues/Fatigue_
Fracture).

For citation purposes, cite each article independently as indicated on the article page online and as
indicated below:

LastName, A.A.; LastName, B.B.; LastName, C.C. Article Title. Journal Name Year, Volume Number,
Page Range.

ISBN 978-3-0365-7600-8 (Hbk)


ISBN 978-3-0365-7601-5 (PDF)

Cover image courtesy of Lucjan Sniezek

© 2023 by the authors. Articles in this book are Open Access and distributed under the Creative
Commons Attribution (CC BY) license, which allows users to download, copy and build upon
published articles, as long as the author and publisher are properly credited, which ensures maximum
dissemination and a wider impact of our publications.
The book as a whole is distributed by MDPI under the terms and conditions of the Creative Commons
license CC BY-NC-ND.
Contents

Preface to "Fatigue and Fracture of Metals and Alloys: Numerical and Experimental Study" . vii

Michal Dziendzikowski, Marcin Kurdelski, Piotr Reymer, Sylwester Klysz and


Krzysztof Dragan
Application of Operational Load Monitoring System for Fatigue Estimation of Main Landing
Gear Attachment Frame of an Aircraft
Reprinted from: Materials 2021, 14, 6564, doi:10.3390/ma14216564 . . . . . . . . . . . . . . . . . . 1

Bohan Wang, Li Cheng and Dongchun Li


Experimental Study on Forged TC4 Titanium Alloy Fatigue Properties under Three-Point
Bending and Life Prediction
Reprinted from: Materials 2021, 14, 5329, doi:10.3390/ma14185329 . . . . . . . . . . . . . . . . . . 21

Sangwon Cho, Geon-Il Kim, Sang-Jin Ko, Jin-Seok Yoo, Yeon-Seung Jung, Yun-Ha Yoo and
Jung-Gu Kim
Comparison of Hydrogen Embrittlement Susceptibility of Different Types of Advanced
High-Strength Steels
Reprinted from: Materials 2022, 15, 3406, doi:10.3390/ma15093406 . . . . . . . . . . . . . . . . . . 37

Nagaraj Ekabote, Krishnaraja G. Kodancha, T. M. Yunus Khan and Irfan Anjum Badruddin
Effect of Strain Rate and Temperature on Tensile and Fracture Performance of AA2050-T84 Alloy
Reprinted from: Materials 2022, 15, 1590, doi:10.3390/ma15041590 . . . . . . . . . . . . . . . . . . 55

Wiktor Wciślik and Sebastian Lipiec


Void-Induced Ductile Fracture of Metals: Experimental Observations
Reprinted from: Materials 2022, 15, 6473, doi:10.3390/ma15186473 . . . . . . . . . . . . . . . . . . 71

Ireneusz Szachogłuchowicz, Lucjan Śnieżek and Tomasz Śl˛ezak


Mechanical Properties Analysis of Explosive Welded Sheet of AA2519-Ti6Al4V with Interlayer
of AA1050 Subjected to Heat-Treatment
Reprinted from: Materials 2022, 15, 4023, doi:10.3390/ma15114023 . . . . . . . . . . . . . . . . . . 99

Sławomir Rowiński
Effect of Steel-Cutting Technology on Fatigue Strength of Steel Structures: Tests and Analyses
Reprinted from: Materials 2021, 14, 6097, doi:10.3390/ma14206097 . . . . . . . . . . . . . . . . . . 113

Marta Orłowska, Ewa Ura-Bińczyk, Lucjan Śnieżek, Paweł Skudniewski, Mariusz Kulczyk,
Bogusława Adamczyk-Cieślak and Kamil Majchrowicz
The Influence of Heat Treatment on the Mechanical Properties and Corrosion Resistance of the
Ultrafine-Grained AA7075 Obtained by Hydrostatic Extrusion
Reprinted from: Materials 2022, 15, 4343, doi:10.3390/ma15124343 . . . . . . . . . . . . . . . . . . 129

Liyong Ma, Chi Liu, Minglei Ma, Zhanying Wang, Donghao Wu, Lijuan Liu and
Mingxing Song
Fatigue Fracture Analysis on 2524 Aluminum Alloy with the Influence of Creep-Aging Forming
Processes
Reprinted from: Materials 2022, 15, 3244, doi:10.3390/ma15093244 . . . . . . . . . . . . . . . . . . 149

Xin Li, Wen Shao, Jinyuan Tang, Han Ding and Weihua Zhou
An Investigation of the Contact Fatigue Characteristics of an RV Reducer Crankshaft,
Considering the Hardness Gradients and Initial Residual Stress
Reprinted from: Materials 2022, 15, 7850, doi:10.3390/ma15217850 . . . . . . . . . . . . . . . . . . 165

v
Urszula Janus-Galkiewicz and Jaroslaw Galkiewicz
Analysis of the Failure Process of Elements Subjected to Monotonic and Cyclic Loading Using
the Wierzbicki–Bai Model
Reprinted from: Materials 2021, 14, 6265, doi:10.3390/ma14216265 . . . . . . . . . . . . . . . . . . 187

Josef Arthur Schönherr, Larissa Duarte, Mauro Madia, Uwe Zerbst, Max Benedikt Geilen,
Marcus Klein and Matthias Oechsner
Robust Determination of Fatigue Crack Propagation Thresholds from Crack Growth Data
Reprinted from: Materials 2022, 15, 4737, doi:10.3390/ma15144737 . . . . . . . . . . . . . . . . . . 205

Stanisław Mroziński, Zbigniew Lis and Halina Egner


Influence of Creep Damage on the Fatigue Life of P91 Steel
Reprinted from: Materials 2022, 15, 4917, doi:10.3390/ma15144917 . . . . . . . . . . . . . . . . . . 227

Yun Zeng, Meiqiu Li, Han Wu, Ning Li and Yang Zhou
Experiment and Theoretical Investigation on Fatigue Life Prediction of Fracturing Pumpheads
Based on a Novel Stress-Field Intensity Approach
Reprinted from: Materials 2022, 15, 4413, doi:10.3390/ma15134413 . . . . . . . . . . . . . . . . . . 239

Muhammad Usman Abdullah and Zulfiqar Ahmad Khan


Further Investigations and Parametric Analysis of Microstructural Alterations under Rolling
Contact Fatigue
Reprinted from: Materials 2022, 15, 8072, doi:10.3390/ma15228072 . . . . . . . . . . . . . . . . . . 263

Kun Liu, Shuai Wang, Lei Pan and X.-Grant Chen


Thermo-Mechanical Fatigue Behavior and Resultant Microstructure Evolution in Al-Si 319 and
356 Cast Alloys
Reprinted from: Materials 2023, 16, 829, doi:10.3390/ma16020829 . . . . . . . . . . . . . . . . . . . 279

Abhishek Biswas, Dzhem Kurtulan, Timothy Ngeru, Abril Azócar Guzmán, Stefanie Hanke
and Alexander Hartmaier
Mechanical Behavior of Austenitic Steel under Multi-Axial Cyclic Loading
Reprinted from: Materials 2023, 16, 1367, doi:10.3390/ma16041367 . . . . . . . . . . . . . . . . . . 297

Chih-Hang Su, Tai-Cheng Chen, Yi-Shiun Ding, Guan-Xun Lu and Leu-Wen Tsay
Effects of Micro-Shot Peening on the Fatigue Strength of Anodized 7075-T6 Alloy
Reprinted from: Materials 2023, 16, 1160, doi:10.3390/ma16031160 . . . . . . . . . . . . . . . . . . 321

Žilvinas Bazaras and Vaidas Lukoševičius


Statistical Characterization of Strain-Controlled Low-Cycle Fatigue Behavior of Structural Steels
and Aluminium Material
Reprinted from: Materials 2022, 15, 8808, doi:10.3390/ma15248808 . . . . . . . . . . . . . . . . . . 335

vi
Preface to "Fatigue and Fracture of Metals and Alloys:
Numerical and Experimental Study"
Dear Readers,

We aimed to present a new book that includes a set of 19 works. We have selected articles
from all over the world that show modern research directions in fatigue and fracture mechanics. The
papers are addressed to scientists who are involved in the assessment of fatigue strength and fracture
toughness. Creating any book requires the cooperation of authors, reviewers, and editors. We would
like to use this opportunity to warmly thank all involved in the editorial process for this book.
Have an excellent read.

Jaroslaw Galkiewicz, Lucjan Śnieżek, and Sebastian Lipiec


Editors

vii
materials
Article
Application of Operational Load Monitoring System for Fatigue
Estimation of Main Landing Gear Attachment Frame of
an Aircraft
Michal Dziendzikowski 1, *, Artur Kurnyta 1 , Piotr Reymer 1,2 , Marcin Kurdelski 1 , Sylwester Klysz 1,3 , Andrzej
Leski 2,4 and Krzysztof Dragan 1

1 Airworthiness Division, Air Force Institute of Technology, ul. Ks. Boleslawa 6, 01-494 Warszawa, Poland;
[email protected] (A.K.); [email protected] (P.R.); [email protected] (M.K.);
[email protected] (S.K.); [email protected] (K.D.)
2 Faculty of Mechanical Engineering, Military University of Technology, ul. gen. S. Kaliskiego 2,
00-908 Warszawa, Poland; [email protected]
3 Faculty of Technical Sciences, University of Warmia and Mazury in Olsztyn, ul. M. Oczapowskiego 2,
10-719 Olsztyn, Poland
4 Institute of Aviation, Lukasiewicz Research Network, al. Krakowska 110/114, 02-256 Warszawa, Poland
* Correspondence: [email protected]

Abstract: In this paper, we present an approach to fatigue estimation of a Main Landing Gear (MLG)
attachment frame due to vertical landing forces based on Operational Loads Monitoring (OLM)
system records. In particular, the impact of different phases of landing and on ground operations and
Citation: Dziendzikowski, M.; fatigue wear of the MLG frame is analyzed. The main functionality of the developed OLM system
Kurnyta, A.; Reymer, P.; Kurdelski,
is the individual assessment of fatigue of the main landing gear node structure for Su-22UM3K
M.; Klysz, S.; Leski, A.; Dragan, K.
aircraft due to standard and Touch-And-Go (T&G) landings. Furthermore, the system allows for
Application of Operational Load
assessment of stress cumulation in the main landing gear node structure during touchdown and
Monitoring System for Fatigue
allows for detection of hard landings. Determination of selected stages of flight, classification of
Estimation of Main Landing Gear
Attachment Frame of an Aircraft.
different types of load cycles of the structure recorded by strain gauge sensors during standard full
Materials 2021, 14, 6564. stop landings and taxiing are also implemented in the developed system. Based on those capabilities,
https://doi.org/10.3390/ma14216564 it is possible to monitor and compare equivalents of landing fatigue wear between airplanes and
landing fatigue wear across all flights of a given airplane, which can be incorporated into fleet
Academic Editors: management paradigms for the purpose of optimal maintenance of aircraft. In this article, a detailed
Jaroslaw Galkiewicz description of the system and algorithms used for landing gear node fatigue assessment is provided,
and Lucjan Śnieżek and the results obtained during the 3-year period of system operation for the fleet of six aircraft are
delivered and discussed.
Received: 30 July 2021
Accepted: 25 October 2021
Keywords: aircraft load monitoring; fatigue estimation; structural health monitoring; landing gear
Published: 1 November 2021
monitoring

Publisher’s Note: MDPI stays neutral


with regard to jurisdictional claims in
published maps and institutional affil-
1. Introduction
iations.
Developments of the industrial revolution, in particular rapid growth of rail transport,
led to the discovery that failure of the material can occur under a stress level much lower
than static tensile strength [1]. Limited durability of structural elements was the main driver
Copyright: © 2021 by the authors.
for the pioneer experiments carried out among others by Wöhler in the field of S–N curves
Licensee MDPI, Basel, Switzerland.
or Palmgren and Miner who developed the linear damage cumulation—which provided
This article is an open access article
the basis for estimating safe-life of structures [2]. In aerospace, the beginnings of system
distributed under the terms and solutions focused on a high level of safety date back to the early post-war years, when the
conditions of the Creative Commons safe-life design concept was introduced, in order to preserve human health and life as well
Attribution (CC BY) license (https:// as mitigating potential material losses [2,3]. Due to the advancement of aircraft structures,
creativecommons.org/licenses/by/ complexity of different types of connections and variable geometry of structural elements,
4.0/). the most reliable method for determining the service life is the subcomponent or full scale

Materials 2021, 14, 6564. https://doi.org/10.3390/ma14216564 1 https://www.mdpi.com/journal/materials


Materials 2021, 14, 6564

fatigue tests of aircraft structures [4–6]. After the F-111 aircraft crash in 1969 [7,8], a new
design and maintenance concept was introduced—the damage tolerance approach [9,10].
In the new methodology, a mandatory requirement for assuring structural integrity is the
introduction of Non-Destructive Testing (NDT) procedures and definition of inspection
intervals of aircraft critical structural elements [11]. Non-destructive inspections were
proposed in order to assure possibility of damage detection, before their development
could jeopardize the safety of aircraft operation. Introduction of the damage tolerance
concept resulted in significant decrease of the risk of air accidents due to fatigue damage.
Despite remarkable advances in the field of fracture mechanics and the development of
numerical modeling methods for complex physical processes, it is impossible to predict all
factors that may increase the risk of an accident. In particular, the design load spectrum
of the structure determines the intervals between subsequent NDT inspections as well
as safety margins for aircraft operation. However, the way in which a particular aircraft
is operated after its introduction into service does not necessarily fit to its pre-assumed
profile, especially for combat aircraft, therefore the load and fatigue spectra can change
during long-term operational use due to numerous factors [12]. Thus, it is necessary
to introduce an Individual Aircraft Tracking (IAT) program in order to monitor actual
load spectrum of every aircraft in operation [10,13,14]. A modern approach to meet this
requirement is to implement Operational Loads Monitoring (OLM) systems [15] as a
necessary component of aircraft avionics. The approach for this purpose has evolved
from g-counters to modern on-board parametric systems incorporating several aspects,
e.g., material fatigue, measurement techniques, signal analysis, modeling, etc., and utilizing
a network of sensors, e.g., strain gauges or Fiber Bragg Gratings (FBG) [16–19] which are
permanently mounted in the aircraft structure and measure local strain in predefined
aircraft locations. Regarding the above, introduction of an Individual Aircraft Tracking
(IAT) program with use of proper OLM system is essential to ensure safety and extend
aircraft service life [13,14,20,21]. Development of OLM systems can also significantly
increase potential use of ageing aircrafts [12], as after performing the necessary overhaul,
such technology may allow their further operation beyond initially designed service life,
if fatigue wear or condition of critical structures can be precisely monitored based on
indications from network of integrated sensors.
In this article, an approach to fatigue estimation of the Main Landing Gear (MLG)
attachment frame due to vertical landing forces based on Operational Loads Monitoring
(OLM) system developed for Su-22UM3K aircraft is presented. In Section 2, the approach to
OLM system design is presented and algorithms for signal analysis and fatigue assessment
are defined. In Section 3 the results obtained during the 3-year period of system operation
for the fleet of six aircraft are delivered and discussed, and in the last section the article
is concluded.

2. Methodology for Load Analysis of Landing Gear Attachment Node


2.1. Full Scale Fatigue Test and Load Monitoring System of Su-22 Aircraft
Su-22 is a variable wing sweep angle fighter–bomber aircraft which was introduced
into operation in Polish Armed Forces (PLAF) in mid-1980’s. There are two versions of the
aircraft: single-seated combat version M4 and a two-seated trainer version UM3K. Based
on the PLAF decision, a service life extension program was launched in 2014 in order to
prolong the designed operation period guaranteed by the manufacturer. As some of the
aircraft were close to the original limits, the actual service life, adequate to the flight profile
in Polish Air Force, had to be validated. Full-Scale Fatigue Tests (FSFTs) have been carried
out in order to verify that the required service life is available with respect to adequate
load spectrum [22]. Furthermore, for two-seated Su22UM3K aircraft, the development of
an OLM system providing reliable and detailed data for Individual Aircraft Tracking (IAT)
program was required, due to their increased usage compared to M4 version, in particular
an increased number of standard and Touch-And-Go landings [14].

2
Materials 2021, 14, 6564

FSFT was designed to be a four-stage test due to the variable sweep angle of the aircraft
wings [22]. This was decided in order to minimize the laborious and time consuming wing
sweep changes and actuator reconfigurations. Different wing sweep angles are used for
takeoff and landing (30◦ ), subsonic flights (45◦ ) and supersonic flights (63◦ ). The final goal
of the FSFT was determination of the total durability of the structure in order to reach
3200 flight hours and 6000 landings. Based on analysis of flight profiles of Polish Su-22,
four aircraft configurations for subsequent stages were determined (Table 1). Landing
loads and loads for flight with extended flaps were represented in separate stages due
to different hydraulic actuators arrangement under wings. The final stage was designed
to represent flights with the highest vertical overloads from Stage 2 until appearance of
critical damage of the structure. In order to represent the load distribution in each stage,
an array of hydraulic actuators was designed for each stage (Figure 1) taking into account
load and displacement range for each considered load node as well as wing sweep angle
and aircraft configuration.

Table 1. Configurations of aircraft for Full-Scale Fatigue Test.

Stage Loads Wing Sweep Angle


I Landing 30°
II Flight 45°
III Flap 30°
IV Flight with high overloads 45°

Figure 1. Full-Scale Fatigue Test rig: (a) schematic view and (b) actual view.

This article is devoted to the development of a methodology for fatigue estimation of


Main Landing Gear (MLG) attachment frame due to aircraft landing; therefore, we will
focus on Stage I of the test while more detailed description of the approach to FSFT of Su-22
aircraft and other test Stages can be found in [22,23]. For Stage I, the wing sweep was set to
30° and the structure was fixed with clamps on the fourth and thirty-fourth fuselage frame
as well as by two rods mounted in the engine bay (highlighted in blue in Figure 2). Loads
were exerted on the structure by means of 16 hydraulic actuators. Six actuators (highlighted
in red in Figure 2) were used in order to represent front (1 actuator acting in the vertical
direction) and main landing gear loads (2 actuator on each side representing vertical and
longitudinal landing forces and 1 actuator in the middle representing transverse taxing
forces), whereas 4 actuators were acting on the variable sweep wing part (2 each side)
corresponding to lift and inertia forces during landing. The remaining 6 actuators were
distributed along the fuselage in order to represent inertial forces. In this study, fatigue
wear of the attachment frame due to it being carried by MLG is considered as one of the
key factors determining remaining service life of an aircraft. Those forces were represented
by actuators denoted as no. 6 in Figure 2 (one actuator on each side). Different types of full
stop landings were represented during Stage I with the same landing mass of the aircraft

3
Materials 2021, 14, 6564

but different vertical load levels during touchdown. Simulated vertical load sequences
on the actuators in terms of the equivalent weight carried by main landing gear node is
shown in Figure 3, and the number of different type of landings represented in FSFT load
spectrum is provided in Table 2 below. Each full stop landing was represented by 14 load
levels exerted on the frame of the aircraft in linear sequences (Figure 3).

Figure 2. Configuration of actuators for Stage I of Full-Scale Fatigue Test execution.

Figure 3. Simulated load sequences exerted on landing gear attachment node during the Full-Scale
Fatigue Test representing different types of full stop landings.

Load envelope and landing statistics were estimated based on dedicated flight tests
program and historical flight data. The information about stress level occurring in various
areas of interest during flight was collected with use of a network of strain gauges installed
on a test aircraft. After sensor installation, a dedicated flight test program was performed
in order to acquire sufficient data about load distribution during flight for the purpose
of FSFT load spectra preparation for different test stages. In total, 40 strain gauges were
installed on the test aircraft selected for in flight measurements (Figure 4):

4
Materials 2021, 14, 6564

• 15 on the left wing (and symmetrically on the right wing):


– 4 on the wing fuselage joint,
– 2 on the main pivot joint,
– 9 on main landing gear and in the compartment and
• 10 on the fuselage in 5 selected Sections (2 each).
The same installation was used for test structure monitoring during FSFT.

Table 2. Number of different landings type represented during Full-Scale Fatigue Test.

Landing Type Landing Count in Load Spectrum


Type 1 1926
Type 2 5603
Type 3 1849
Type 4 589
Type 5 157
Type 6 36
Type 7 9
Type 8 1

Figure 4. General view of the strain gauge network on the test aircraft (due to the symmetry of the
sensor network only one side of the aircraft is shown).

The strain gauge measurement array was designed to monitor the following loads:
• bending momentum in the wing fuselage joint (one section each side);
• bending momentum in the main pivot joint (one section each side);
• bending momentum in the fuselage (5 sections);
• main landing gear loads:
– vertical force along z axis;

5
Materials 2021, 14, 6564

– bending momentum along y axis;


– bending momentum along x axis.
All of the strain gauges were in the form of tee rosette configuration (Figure 5b), which
allowed for strain measurements in the primary direction with temperature compensation
due to the secondary perpendicular strain gauge. In Figure 5a, localization of strain gauges
installed in the main landing gear compartment is presented.

Figure 5. Localization and type of strain gauges installed on test aircraft: (a) localization of sensors in MLG compartment
and (b) tee rosette strain gauges used (250UT manufactured by Vishay).

In addition, a load monitoring system with a reduced number of strain gauges was
installed on six selected Su-22UM3K aircraft, in order to track their individual usage and
remaining service life. In particular, as aircraft selected to be equipped with a Operational
Load Monitoring (OLM) system are used for pilot training and certification, it was nec-
essary to develop a methodology for main landing gear attachment node fatigue wear
monitoring, as those aircraft perform more landings (especially Touch-And-Go landings)
than combat one-seat aircraft. There exist many approaches to Individual Aircraft Track-
ing [10] depending on the type of data available for load estimation in critical locations.
A through revision of different methods for fatigue assessment and comparison of different
monitoring techniques is provided in [24]. One of the most common approaches is to
use the vertical overload parameter recorded by Flight Data Recorder (FDR) in order to
determine stress levels at critical locations. This parameter, in particular in combination
with some other flight parameters, e.g., angle of attack and aircraft mass distribution, is es-
pecially efficient for fatigue assessment of wing spars or wing to fuselage attachment nodes,
as vertical acceleration is the key parameter determining fatigue wear of such structural
elements. Based on acceleration measurements, it is moreover possible to define many use-
ful damage sensitive signal features for the purpose of Structural Health Monitoring [25].
However, the bandwidth and sensitivity of accelerometers used in FDR are sometimes not
sufficient for proper determination of fatigue due to landings, in particular it can be hard to
detect and properly assess touchdown of an aircraft based on that parameter. Furthermore,
accelerometers cannot be used for determination of takeoff weight which in the case of
landing gear attachment node can significantly contribute to Ground-Air-Ground cycle,
therefore strain gauges were used for OLM purposes.
The sensor network of the system developed for the IAT program is reduced with
respect to the system installed on the test aircraft (Figure 4). It contains eight strain gauges
which are installed symmetrically in the most relevant structural elements of the aircraft
on both wings and main landing gear frame. In Figure 6 general location of strain gauges
of the reduced system is presented. The limitation to eight strain gauges was due to the

6
Materials 2021, 14, 6564

hardware requirements of the FDR used. Four sensors are placed on the main wing spar,
on lower (sensors denoted as T1 and T5) and upper (sensors denoted as T2 and T6) flanges.
For main landing gear monitoring, the sensor denoted as GGPZ1 in Figure 5a was selected
to measure backward bending moment of the landing gear (denoted as T3 and T7 in the
reduced system), and sensor GGPS2 (Figure 5a) is used for estimation of vertical force
during landing (denoted as T4 and T8 in the reduced system). Strain gauge selection
was determined based on signal-to-noise ratio measured during flight tests as well as the
results of sensors calibration on test specimen during FSFT. All the selected strain gauges
depended linearly on the corresponding relevant forces. In particular, sensor T4 (Figure 6b)
was linearly dependent on vertical force applied to main landing gear in full range of the
load spectrum while it was barely sensitive to force applied in perpendicular direction.
Therefore, for the purpose of fatigue wear estimation of the MLG attachment node due to
vertical landing forces, the strain values ε measured by sensors T4 (left node) and T8 (right
node) were used.

Figure 6. Reduced system for load monitoring: (a) general location of strain gauges of the system (left MLG bay) and
(b) view of T4 sensor with indication of measurement direction.

2.2. Methodology of Signal Analysis for Landing Operations Extraction


In order to determine fatigue of landing gear attachment node, it is necessary to
extract signals corresponding to on-ground operations from full record of a given flight.
For the detection of on-ground segments recorded by the system, it was assumed that the
threshold ε T for detection of significant deformations originating from landing gear loads
corresponding to landings is 20% of the base deformation, which is determined before the
takeoff of the plane, using the loads at full stop. In Figure 7, an example of signal acquired
during landing with multiple touchdowns preceding final touchdown and deceleration is
presented. As can be seen, signal can exceed the estimated threshold during before final
deceleration and load transfer to landing gear node; therefore, for a proper determination
of signal corresponding to landings, an additional algorithm is required.
In the presented algorithm, the first step of on-ground operations detection and
classification is to select all the signal samples ε(t) below the determined threshold ε T ,
i.e., satisfying the condition
ε(t) < ε T , (1)
where t is time of a given sample acquisition. Then, starting from the first signal sample
satisfying this condition, consecutive disjoints sets of data are determined:

τ1 = {ε(ti,1 ), . . . ε(t f ,1 )}, . . . , τl = {ε(ti,l ), . . . ε(t f ,l )}, (2)

7
Materials 2021, 14, 6564

such that the difference between the acquisition time ti,k of the initial signal value ε(ti,k )
from a given set τk and acquisition time t f ,k−1 of the final signal value ε(t f ,k−1 ) from the
preceding set τk−1 is not less than 3 s. For proper on-ground connected signal segments
detection, datasets τj lasting no longer than 2 s, i.e., for which

t f ,j − ti,j ≤ 2 (3)

are disregarded, in order to remove eventual artificial events due to natural signal distur-
bances, e.g., due to short power outages or drops of pressure in hydraulic blocks during
flight. In the next step, for every segment τj , all the signal samples ε(t) satisfying

ti,j − 3 ≤ t ≤ t f ,j + 3 (4)

are added to a given set τj in order to analyze full data records corresponding to subsequent
on-ground operations, i.e., each part of the signal corresponding to such operation is
extended by additional 3 s time offset of signal before and after the operation, in order
to track all load cycles exerted on landing gear node. The first of such extended sets, τ1 ,
contains signal acquired during aircraft takeoff and taxiing before takeoff, and the last set
τl corresponds to full stop landing and taxiing after landing, whereas all sets in between
are classified as Touch-And-Go (T&G) landings.
In Figure 8, the outcome of the proposed algorithm for landing presented in Figure 7
is delivered. The signal corresponding to full stop landing was extracted from raw signal
with use of the proposed algorithm. In particular, all the aircraft touchdowns are included
in the extracted signal, and an additional 3 s time offset provides data corresponding to no
load condition on landing gear node; therefore, based on such data, it is possible to capture
all relevant load cycles exerted on the node due to landing.

Figure 7. An example of signal acquired by strain gauge located at main gear attachment node
during full stop landing with indication of signal threshold for landing detection.

All the parameters, e.g., threshold ε T and offset levels, were decided based on algo-
rithm results on database of reference signals, where signals corresponding to on-ground
operations were manually determined, so the number and duration of landings was com-
pared between expert and automated analysis for different adjustments of the parameters.
Furthermore, the performance of the algorithm is periodically verified, i.e., number of
landings is compared with maintenance data, but also validity and performance of the algo-
rithm is evaluated by the experts based on randomly selected flights from a given period.

8
Materials 2021, 14, 6564

Figure 8. An example of automated extraction of signal corresponding to full stop landing.

In Figure 9a, an example of an OLM system record with indication of subsequent


on-ground operations is presented. The approach is efficient in detection of on-ground
operations, in particular for Touch-And-Go landings (Figure 9b). Detection of such events
based on flight parameters acquired by a standard flight data recorder was very inaccurate,
as many simulated landing approaches, but without touchdown, are performed during
pilot training process. In Figure 10, an example of system records for a flight with three
simulated attempts to landing is presented. For every such maneuver, low pass flight over
runway was performed (Figure 10b) with released landing gear, which caused a slight
change of strain values recorded on the node (Figure 10a); however, no actual touchdown
occurred. This is correctly recognized by the proposed algorithm, as only takeoff and
full stop landing were detected in that flight (Figure 10a). As both records—the altitude
parameter as well as indicator of landing gear release command—were similar as for proper
Touch-And-Go landings, such events were often misclassified with use of algorithms based
on only flight parameters records, as sensitivity and signal sampling rate of the standard
g-load sensor used on this type of aircraft are not sufficient for such events detection.
An interesting example of algorithm output is presented in Figure 11. The recorded
strain signal shows a clear example of a bounced landing. In this case, after the first
touchdown, the plane took off from the ground for about two seconds, and full stop
landing was performed afterwards. The algorithm detected a Touch-And-Go landing event
and subsequent proper landing. Such events, when detected by the system with high
confidence, can provide an automated tool for human error assessment, which could be
beneficial for pilot training programs.
In Figure 12, an example of a strain gauge signal recorded during full stop landing
is shown, with indication of the characteristic stages of this process. One of the param-
eters used to assess the stress level of the main landing gear node structure during full
stop and Touch-And-Go landings is the amplitude of the maximum deformation cycle
recorded during the touchdown (Figure 12). A characteristic feature of the touchdown
is the rapid deformation change related to the slowing down of the descent speed of the
aircraft and the dissipation of energy on the elements of the landing gear node [26]. Two
warning levels were defined for the strain signals characterizing hard landings, which are
presented in Table 3. In addition to structural load monitoring purposes, the distribution
of landings corresponding to different warning levels can also be used for the assessment
of pilot training advancements. The warning levels were determined based on landing
statistics determined for a certain period of time. Another approach could be based on
relevant material data and depend on the stress values at the critical point for the main
landing gear attachment node, yet the adopted statistical approach is better suited for pilot
training purposes.

9
Materials 2021, 14, 6564

Figure 9. An example of load monitoring system records: (a) full record with indication of all detected on-ground operations;
(b) signal acquired during first detected Touch-And-Go landing (marked in green).

Figure 10. An example of flight records with simulated landing approaches: (a) strain on landing gear attachment node
with indication of take-off and full stop landing; (b) altitude (barometric).

Table 3. Warning levels for maximum strain amplitude [μStr] during aircraft touchdown.

Min. Touchdown Amplitude Max. Touchdown Amplitude


standard landing 0 1100
first warning level >1100 1500
second warning level >1500 -

10
Materials 2021, 14, 6564

Figure 11. An example of system record with subsequent Touch-And-Go and full stop landing.

Figure 12. An example of strain record for full stop landing with indication of subsequent stages of
this maneuver.

2.3. Methodology of Landing Gear Node Fatigue Estimation


In this paragraph, a method for fatigue estimation of Main Landing Gear (MLG)
attachment frame due to vertical landing forces based on OLM system records is delivered.
The classic approach to fatigue estimation is based on the S–N fatigue curve of a given
material and linear cumulative Palmgren–Miner hypothesis. The S–N curve determines
the relationship between the equivalent stress amplitude σeq of the zero-to-tension load
cycle and the number of cycles N needed to fracture of an element, subjected to load cycles
of this amplitude. The S–N curve is usually described by the relationship [27]

log σeq = A + B log N. (5)

11
Materials 2021, 14, 6564

eq eq
Cumulative fatigue D of a set of load cycles {σ1 , . . . , σM } is given by the expression

M M  1
1 eq − B
D= ∑ Ni = C ∑ σi , (6)
i =1 i =1

where C is a constant dependent on material constants A and B.


Based on classic S–N and cumulative fatigue equations presented above, an approach
to fatigue wear estimation of main landing gear attachment frame can be defined as
follows. In order to determine the fatigue wear of a given structural element, a linear
relation between stress level σc at critical point of the element and some physical parameter
p can be assumed:
σc = αp. (7)
As the parameter p, records of strain gauge ε installed in a point of the structure
where stress level is linearly proportional to σc can be used, or p can be a function of flight
parameters determining load values of a given structural element, e.g., weight and g-load
factor in the case of fuselage wing attachment. Equation (6) can be rewritten as follows:

M  1 M
eq − B −1
D = C ∑ σc,i = C̃ ∑ pi B , (8)
i =1 i =1

where C̃ = Cα− B and pi is value of physical parameter corresponding to equivalent


1

zero-to-tension stress amplitude of the i-th load cycle.


The material constant C and proportionality parameter α can be omitted if reference
fatigue for a given set of cycles is known, for instance, if data from Full Scale Fatigue Test
are available. In that case,
MFSFT
− B1
1 = DFSFT = C̃ ∑ pj , (9)
j =1

where summation is carried over load cycles p j exerted on the structure during Full-Scale
Fatigue Test (FSFT) and MFSFT is number of load cycles during FSFT. Therefore,

− B1
D ∑iM
=1 p i
D= = , (10)
DFSFT M − B1
∑ j=FSFT
1 pj

can be calculated if material constant B and reference data from FSFT is known.
There exist several conditions which need to be satisfied in order to apply the presented
approach based on Miner’s law. First, if only laboratory material data are available with no
reference fatigue exerted on a real test structure, then the correspondence between stress
level in the critical location of the structure and the parameter used for fatigue evaluation
must be known in exact form and Miner’s Equation (6) needs to be used directly instead of
the Equation (10). Another requirement is a linear relation between the physical parameter
used for fatigue assessment and stress level σc in the critical location. As mentioned in
Section 2.1, strain gauge readings used in the study are linearly dependent on vertical
landing force in full range of admissible loads of the aircraft, as confirmed during FSFT rig
calibration. Furthermore, in the case of Su-22 aircraft, it is assumed in load spectra design
that the aircraft will be operated within the linear elastic regime of the materials, as if high
overloads during flight or very harsh landings occur, special procedures are introduced
in order to evaluate aircraft condition (e.g., plastic deformations) and its airworthiness.
In addition, no looseness in highly loaded joints is allowed, therefore linearity between
vertical force and stress level in critical location is legitimate assumption in our case. Finally,
fatigue limit [1] of the material needs to be considered for fatigue estimation based on
Equation (10), as load cycles below the fatigue threshold in a given point can be above
fatigue limit in critical location. In our case, all strain data corresponding to physical load

12
Materials 2021, 14, 6564

cycles exerted on main landing gear attachment are taken into account for the purpose of
fatigue assessment, therefore all the load cycles in critical location are considered as well.
For the purpose of this study, the equivalent weight w carried by landing gear node
was adopted as a physical parameter needed for fatigue wear calculation in accordance
with the Equation (8). Equivalent weight was assessed by strain gauge reading and results
of linear physical scaling, i.e., values of strain for lifted aircraft with released landing gear
were related to no load condition and values of strain readings for aircraft on-ground
were related to its measured weight carried by a given landing gear node. Furthermore,
data from actuators used for the Full-Scale Fatigue Test were rescaled in those units and
used in the denominator of the Equation (10) for fatigue calculation. The parameter B was
estimated based on laboratory fatigue tests of material used for landing gear attachment
node manufacturing.
For easier interpretation of the obtained results, a notion of Landing Fatigue Equivalent
(LFE) can be introduced. LFE represents the relative fatigue of landing gear node due to
loads exerted during a given landing Dl with respect to mean fatigue due to simulated
landings during FSFT:
M − B1
∑i=l1 wi
LFE = , (11)
1 M − B1
N ∑ j=FSFT
1 wj

where:
• wi denotes amplitude of equivalent zero-to-tension load cycles exerted on landing gear
node during landing (in terms of equivalent weight carried by the node as measured
by strain gauge),
• Ml denotes number of load cycles recorded during landing,
• w j denotes amplitude of equivalent zero-to-tension load cycles exerted on landing
gear node during FSFT (in terms of equivalent weight carried by the node) and
• N denotes number of simulated landings during FSFT.
In order to account properly Ground–Air–Ground cycle when calculating LFE for
full stop landing, the records of strain gauge obtained for takeoff and landing are joined
(Figure 13) prior to determination of load cycles with use of Range-Pair Counting algo-
rithm [28]. Furthermore, the noise level of the recorded signal during on-ground operations
was estimated, and for LFE calculation, only relevant recorded cycles, i.e., corresponding
to physical load of the structure and higher than the level of noise, were considered. LFE
provides a direct measure to estimate the fatigue wear of the landing gear attachment node
for a given flight if:
• LFE < 1 then loads exerted on landing gear attachment node during a given landing
was less severe than mean landing profile during FSFT; or
• LFE > 1 then loads exerted on landing gear attachment node during a given landing
was more severe than mean landing profile during FSFT.
Cumulative LFE obtained for a given aircraft can be considered as a limiting condition
for possibility of further aircraft operation instead of total estimated fatigue for landing
gear attachment.
For Touch-And-Go landings, all the load cycles are due to aircraft touchdown. For full
stop landing, additional information about this process can be obtained by distinction of
load cycles occurring during different stages of landing (Figure 12) and calculation of the
corresponding fatigue. An algorithm for the classification of structure load cycles recorded
by strain gauges during the normal landings and other ground loads during take-off and
full stop landing was developed. It was assumed that the cycles are classified into the
following categories:
• cycles recorded during touchdown—P;
• Ground–Air–Ground cycle—GAG;
• cycles recorded during braking—W;
• cycles recorded during taxing before take-off—KSR;

13
Materials 2021, 14, 6564

• cycles recorded during taxing after landing—DKL;


• other cycles recorded during takeoff—OS;
• other cycles (not classified elsewhere)—O.
In Figure 13, the general concept of division of a single take-off and landing operation
into above defined cycles is shown—the height of the shown color boxes correspond the
assumed minimum and maximum values of certain types of cycles for a given flight.

Figure 13. An example of strain record for full stop landing with indication of subsequent stages of
this maneuver.

Since:
Ml
− B1 −1 − B1 − B1 − B1 − B1
∑ wi = wGAG
B
+ ∑ wc + ∑ wc + ∑ wc + ∑ wc + (12)
i =1 c∈ P c ∈W c∈KSR c∈ DKL
− B1 − B1
+ ∑ wc + ∑ wc ,
c∈OS c ∈O

where c denotes cycles corresponding to different types, also fatigue equivalent obtained
for full stop landing can be factored accordingly:

LFE = LFEGAG + LFEP + LFEW + LFEKSR + LFEDKL + LFEDKL + (13)


+ LFEOS + LFEO .

Based on investigation of the defined contributions to fatigue equivalent, additional


conclusions can be drawn, e.g., with respect to pilot training program or with respect to
condition of the runway, as will be shown further in the text.

3. Results
In the period 2018–2021, a total number of 1563 full stop landings and 350 Touch-And-
Go landings of six Su-22 aircraft equipped with Operational Load Monitoring system were
recorded. In Figure 14, boxplots of LFE for full stop landing and Touch-And-Go landings
are presented.

14
Materials 2021, 14, 6564

Figure 14. Boxplot of Landing Fatigue Equivalent for full stop (FS) and Touch-And-Go (T&G) landings.

It is demonstrated that the fatigue wear of the landing gear attachment node is signifi-
cantly higher for full stop landings compared to Touch-And-Go landings. In the former
case, the median value of LFE is 1.24 compared to 0.13 obtained for the latter landings
type. This result is of particular importance for estimation of remaining fatigue life of
landing gear attachment node for aircraft not equipped with a load monitoring system,
as usually service life is evaluated with respect to number of landings without distinction
between full stop and Touch-And-Go landings. Median value of fatigue corresponding
to full stop landing is about 24% higher mean fatigue value of simulated landings repre-
sented during Full-Scale Fatigue Test. Partially, it can be related to number of touchdown
cycles represented for simulated landings. For simulated landings, only two load cycles
were represented for touchdown as shown in Figure 3, in order to meet time constraint
requirements for the test completion. In reality, the number of load cycles during aircraft
touchdown is significantly higher as can be seen in Figure 9b or Figure 12.
In Table 4, the distribution of LFE to a different type of fatigue cycle is presented.
Higher fatigue wear of the landing gear node obtained for full stop landing is mostly
due to the Ground–Air–Ground cycle. The median value of the LFE for the GAG cycle is
1.04, which is approximately 84% of total median fatigue equivalent. This indicates that
increased simulated take-off weight of aircraft during FSFT (Figure 3) would represent
better true distribution of operational take-off aircraft weight. Considerable contribution to
LFE is also due to touchdown loads, i.e., median LFEP = 0.083.

Table 4. Contribution of different types of cycles to Landing Fatigue Equivalent (LFE).

Cycles Type GAG P D KL KS Other Cycles


median LFE 1.040 0.083 0.004 0.030 0.003 0.016

Typically, higher LFE is due to increased takeoff weight or due to hard touchdown
during landing (Figure 15). In Figure 15a, takeoff and full stop landing for a flight
with LFE = 3.27 is shown. In this flight, the performed touchdown was smooth with
LFEP = 0.25; nevertheless, take-off weight of the aircraft was about 2000 kg higher than
assumed for simulated landings during FSFT. This caused significant growth of the GAG
cycle contribution to the fatigue of landing gear attachment node. The obtained LFEGAG
for this flight was 2.84.

15
Materials 2021, 14, 6564

Figure 15. An example of system records for flights with high values of Landing Fatigue Equivalent obtained for full stop
landing: (a) flight with increased take-off aircraft weight; (b) flight with hard touchdown.

The second example is a flight with hard touchdown during landing with very high
LFE = 8.37. Take-off weight in this case was comparable, as assumed for FSFT, which
resulted in LFEGAG = 1.30. However, very high load cycle was exerted on the structure
during touchdown with the amplitude about 10,000 kg of equivalent carried weight. Thus,
LFEP obtained for the flight was 6.95. Such events can be due to a sudden rapid gust
of wind or due to the premature release of the drogue parachute (Figure 16), which is
sometimes trained as simulation of landing on an airfield with a very short runway.
Furthermore, interesting cases of increased LFE were observed for aircraft operations
performed from a particular air base. In Table 5, the total mean LFE with contributions
of selected type of cycles is presented. Significant increase of fatigue of landing gear
attachment node was obtained for aircraft operations from one of the airfields (Figure 17).

Table 5. Mean Landing Fatigue Equivalent values corresponding to different types of cycles with
respect to operations performed from different air bases.

Localization GAG P KL KS Total


air base No. 1 0.95 0.17 0.04 0.07 1.17
air base No. 2 0.99 0.20 0.03 0.06 1.24
air base No. 3 2.41 0.86 0.18 1.21 4.30

Total mean LFE obtained for landings on that airfield was 4.30 compared to fatigue
equivalent 1.17 obtained for operations from the home air base. Significant increase of
contributions to LFE due to different type of cycles was observed for Base 3, in particular
LFE of cycles due to taxing before take-off, was more than 18 times higher than in regular
conditions. In fact, abnormal variation of strain during aircraft take-off was observed as
shown in Figure 18, which can be caused by particularly bad condition of runway.

16
Materials 2021, 14, 6564

Figure 16. Examples of landings with prematurely opened drogue parachute [29–31].

Figure 17. Landing Fatigue Equivalent for subsequent flights of an aircraft in a given period.

17
Materials 2021, 14, 6564

Figure 18. Records of strain during take-off for selected flights from the home base and air base no. 3.

4. Conclusions
In this paper, application of operational load monitoring system of an aircraft to
fatigue equivalent estimation of landing gear attachment node was presented. In particular,
system description algorithms for signal processing and fatigue estimation were provided.
The notion of Landing Fatigue Equivalent (LFE) was introduced as a measure representing
the relative fatigue of landing gear node due to loads exerted during a given landing with
respect to mean fatigue of simulated landings during Full-Scale Fatigue Test of the structure.
It was demonstrated that the LFE obtained for full stop landings is significantly higher
than LFE for Touch-And-Go landings. The contribution of different types of load cycles
to LFE was also investigated. Predominant contribution to LFE, i.e., over 80%, is due the
Ground–Air–Ground (GAG) cycle; furthermore, loads exerted on landing gear attachment
node during aircraft touchdown have considerable effect on fatigue wear of the structure.
Particular examples of flights with high values of LFE were presented and discussed.

Author Contributions: Conceptualization, M.D.; methodology, M.D., M.K. and P.R.; software, M.D.;
validation, M.K. and A.K.; formal analysis, M.K. and K.D.; investigation, M.D. and M.K.; resources,
M.D. and A.K.; data curation, A.K. and M.D.; writing—original draft preparation, M.D. and P.R.;
writing—review and editing, S.K., A.L. and K.D.; visualization, M.D. and P.R.; project administration,
M.K., A.L. and K.D.; scientific consultation, S.K., A.L. and K.D.; supervision, S.K. and K.D. All
authors have read and agreed to the published version of the manuscript.
Funding: This research received no external funding.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Data used in this study are available on-demand from the correspond-
ing author.
Acknowledgments: In the memory of Colonel Andrzej Lesniczak.
Conflicts of Interest: The authors declare no conflict of interest.

18
Materials 2021, 14, 6564

Abbreviations
The following abbreviations are used in this manuscript:

FDR Flight Data Recorder


FSFT Full Scale Fatigue Test
GAG Ground–Air–Ground (cycle)
IAT Individual Aircraft Tracking
LFE Landing Fatigue Equivalent
MLG Main Landing Gear
OLM Operational Load Monitoring
T&G Touch-And-Go (landing)

References
1. Milella, P.P. Fatigue and Corrosion in Metals; Springer: Milano, Italy, 2013.
2. Schütz, W. A history of fatigue. Eng. Fract. Mech. 1996, 54, 263–300. [CrossRef]
3. Riddick, H. Safe-Life and Damage-Tolerant Design Approach for Helicopter Structures Applied Technology Laboratory; US Army Research
and Technology Laboratories (AVRADCOM): Newport News, VA, USA, 1984.
4. Schijve, J. Fatigue of aircraft materials and structures. Int. J. Fatigue 1994, 16, 21–32. [CrossRef]
5. Leski, A.; Kurdelski, M.; Reymer, P.; Dragan, K.; Sałaciński, M. Fatigue life assessment of PZL-130 Orlik structure–final analysis
and results. In Proceedings of the 28th ICAF Symposium, Helsinki, Finland, 1–5 June 2015; pp. 294–303.
6. Daverschot, D.; Mattheij, P.; Renner, M.; Ardianto, Y.; De Araujo, M.; Graham, K. Full-Scale Fatigue Testing from a Structural
Analysis Perspective. In ICAF 2019—Structural Integrity in the Age of Additive Manufacturing; Springer: Cham, Switzerland, 2019;
pp. 788–800.
7. Schijve, J. Fatigue damage in aircraft structures, not wanted, but tolerated? Int. J. Fatigue 2009, 31, 998–1011. [CrossRef]
8. Wanhill, R.; Molent, L.; Barter, S. Milestone Case Histories in Aircraft Structural Integrity. In Reference Module in Materials Science
and Materials Engineering; Elsevier Science: Amsterdam, The Netherlands, 2016.
9. Kim, Y.; Sheehy, S.; Lenhardt, D. A Survey of Aircraft Structural-Life Management Programs in the US Navy, the Canadian Forces, and
the US Air Force; Rand Corporation: Santa Monica, CA, USA, 2006; Volume 370.
10. MIL-STD-1530D, Department of Defense Standard Practice Aircraft Structural Integrity Program (ASIP); US Department of Defense:
Washington, DC, USA, 2016.
11. Grandt, A.F., Jr. Fundamentals of Structural Integrity: Damage Tolerant Design and Nondestructive Evaluation; John Wiley & Sons:
Hoboken, NJ, USA, 2003.
12. Dilger, R.; Hickethier, H.; Greenhalgh, M.D. Eurofighter a safe life aircraft in the age of damage tolerance. Int. J. Fatigue 2009,
31, 1017–1023. [CrossRef]
13. Rui, J.; Xiaofan, H.; Yuhai, L. Individual aircraft life monitoring: An engineering approach for fatigue damage evaluation. Chin. J.
Aeronaut. 2018, 31, 727–739.
14. Reymer, P.; Kurdelski, M.; Leski, A.; Leśniczak, A.; Dziendzikowski, M. Introduction of an Individual Aircraft Tracking Program
for the Polish Su-22. Fatigue Aircr. Struct. 2017, 9, 101–108. [CrossRef]
15. Boller, C.; Staszewski, W.J. Aircraft Structural Health and Usage Monitoring. In Health Monitoring of Aerospace Structures; John
Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2003; Chapter 2, pp. 29–73.
16. Kressel, I.; Dorfman, B.; Botsev, Y.; Handelman, A.; Balter, J.; Pillai, A.; Prasad, M.; Gupta, N.; Joseph, A.; Sundaram, R.; et al.
Flight validation of an embedded structural health monitoring system for an unmanned aerial vehicle. Smart Mater. Struct. 2015,
24, 075022. [CrossRef]
17. Nicolas, M.J.; Sullivan, R.W.; Richards, W.L. Large scale applications using FBG sensors: Determination of in-flight loads and
shape of a composite aircraft wing. Aerospace 2016, 3, 18. [CrossRef]
18. Iele, A.; Leone, M.; Consales, M.; Persiano, G.; Brindisi, A.; Ameduri, S.; Concilio, A.; Ciminello, M.; Apicella, A.; Bocchetto, F.;
et al. Load monitoring of aircraft landing gears using fiber optic sensors. Sens. Actuators A Phys. 2018, 281, 31–41. [CrossRef]
19. Ma, Z.; Chen, X. Fiber Bragg gratings sensors for aircraft wing shape measurement: Recent applications and technical analysis.
Sensors 2019, 19, 55. [CrossRef] [PubMed]
20. Molent, L. A review of a strain and flight parameter data based aircraft fatigue usage monitoring system. In Proceedings of the
USAF ASIP Conference, San Antonio, TX, USA, 3–5 December 1996.
21. Molent, L.; Barter, S.; Foster, W. Verification of an individual aircraft fatigue monitoring system. Int. J. Fatigue 2012, 43, 128–133.
[CrossRef]
22. Leski, A.; Reymer, P.; Kurdelski, M.; Zieliński, W.; Jankowski, K. Full scale fatigue test of the Su-22 aircraft–Assumptions, process
and preliminary conclusions. AIP Conf. Proc. 2016, 1780, 020002.
23. Reymer, P.; Kurdelski, M.; Leski, A.; Jankowski, K. The Definition of the Load Spectrum for SU-22 Fighter-Bomber Full Scale
Fatigue Test. Fatigue Aircr. Struct. 2015, 1, 28–33. [CrossRef]
24. Molent, L.; Aktepe, B. Review of fatigue monitoring of agile military aircraft. Fatigue Fract. Eng. Mater. Struct. 2000, 23, 767–785.
[CrossRef]

19
Materials 2021, 14, 6564

25. Civera, M.; Ferraris, M.; Ceravolo, R.; Surace, C.; Betti, R. The Teager-Kaiser Energy Cepstral Coefficients as an Effective Structural
Health Monitoring Tool. Appl. Sci. 2019, 9, 5064. [CrossRef]
26. Kurdelski, M.; Leski, A.; Krzysztof, D. Fatigue life analysis of main landing gear pull–rod of the fighter jet aircraft’. In Proceedings
of the 28th International Congress of the Aeronautical Sciences, Brisbane, Australia, 23–28 September 2012.
27. DOT/FAA/AR-MMPDS-01: Metallic Materials Properties Development and Standardization (MMPDS); Office of Aviation Research:
Washington, DC, USA, 2003.
28. ASTM. E1049–Standard Practices for Cycle Counting in Fatigue Analysis; ASTM International: West Conshohocken, PA, USA, 2017.
29. Su-22—Hamowanie Spadochronem Przed Przyziemieniem. Available online: https://www.youtube.com/watch?v=g4
jMkJiCTLE (accessed on 7 October 2021).
30. RIAT 2015|Su-22 Arrives. Available online: https://www.youtube.com/watch?v=iyLowoSilgY (accessed on 7 October 2021).
31. Su-22 Drogue-Chute Landing Gone Wrong. Available online: https://www.youtube.com/watch?v=c3DlE_GJFMw (accessed on
7 October 2021).

20
materials
Article
Experimental Study on Forged TC4 Titanium Alloy Fatigue
Properties under Three-Point Bending and Life Prediction
Bohan Wang *, Li Cheng and Dongchun Li

Aeronautics Engineering College, Air Force Engineering University, Xi’an 710038, China;
[email protected] (L.C.); [email protected] (D.L.)
* Correspondence: [email protected]; Tel.: +86-029-8478-7225

Abstract: Ultrasonic fatigue tests of TC4 titanium alloy equiaxed I, II and bimodal I, II obtained by
different forging processes were carried out in the range from 105 to 109 cycles using 20 kHz three-
point bending. The results showed that the S-N curves had different shapes, there was no traditional
fatigue limit, and the bimodal I had the best comprehensive fatigue performance. The fracture mor-
phology was analyzed by SEM, and it was found that the fatigue cracks originated from the surface
or subsurface facets, showing a transgranular quasi-cleavage fracture mechanism. EDS analysis
showed that the facets were formed by the cleavage of primary α grains, and the fatigue cracks
originated from the primary α grain preferred textures, rather than the primary α grain clusters.
From the microstructure perspective, the reasons for better equiaxed high-cycle-fatigue properties
and better bimodal ultra-high-cycle-fatigue properties were analyzed. The bimodal I fatigue life
prediction based on energy was also completed, and the prediction curve was basically consistent
with the experimental data.

Keywords: forging; TC4 titanium alloy; three-point bending; very high cycle fatigue; microstructure;
Citation: Wang, B.; Cheng, L.; Li, D. life prediction
Experimental Study on Forged TC4
Titanium Alloy Fatigue Properties
under Three-Point Bending and Life
Prediction. Materials 2021, 14, 5329. 1. Introduction
https://doi.org/10.3390/ma14185329
With the continuous improvement of aviation equipment reliability and life index,
the ultra-high-cycle-fatigue (107 ~1012 cycles) problem has attracted increasingly more atten-
Academic Editors: Jaroslaw Galkiewicz
tion, and has become the focus of fatigue research [1–5]. Ultra-high-cycle fatigue belongs to
and Lucjan Śnieżek
micro-scale fatigue, which differs from traditional high-cycle fatigue (105 ~107 cycles) not
only in the life length, but also in the more complex crack initiation and initial propagation
Received: 24 August 2021
Accepted: 13 September 2021
mechanism, and also in the great difference due to different materials [6–8]. Titanium alloys
Published: 15 September 2021
have become the most widely used metal materials in the modern aviation industry because
of their high specific strength, low density, excellent corrosion resistance, and good heat
Publisher’s Note: MDPI stays neutral
resistance, and the ultra-high-cycle fatigue problems are also the most prominent. Among
with regard to jurisdictional claims in
them, TC4 titanium alloy is widely used in the manufacture of aero-engine compressor
published maps and institutional affil- blades. The engine rotating speed is as high as tens of thousands of revolutions per minute,
iations. and it is very easy for rotating parts such as blades and discs to cause high-frequency
flutter under the internal flow disturbance and rotor imbalance action, and ultra-high-
cycle-fatigue fracture problems are easy to occur when accumulating for a long time [9,10].
This shows that the safety of aero-engine compressor design cannot be ensured according
Copyright: © 2021 by the authors.
to the traditional 107 fatigue limit design theory.
Licensee MDPI, Basel, Switzerland.
Due to the observation method limitations and the theoretical analysis complexity,
This article is an open access article
ultra-high-cycle-fatigue studies basically adopt phenomenological analysis methods based
distributed under the terms and on experimental results, including macroscopic analysis of the S-N curve and Goodman
conditions of the Creative Commons diagram based on experimental results, and microscopic analysis of crack initiation and
Attribution (CC BY) license (https:// initial propagation based on fracture morphology. The proportion of crack initiation life
creativecommons.org/licenses/by/ is more than 95% in the high-cycle-fatigue and ultra-high-cycle-fatigue regimes [11,12];
4.0/). therefore, the crack initiation mechanism and its relationship with fatigue life are very

Materials 2021, 14, 5329. https://doi.org/10.3390/ma14185329 21 https://www.mdpi.com/journal/materials


Materials 2021, 14, 5329

important. A large number of studies have shown that [13–15] the remarkable crack
initiation zone sign for titanium alloys is the smooth facet morphology, especially for
failure on the subsurface. The reasons for the facet formation involve three different
mechanisms: cleavage [16,17], slip [18,19], and twins [20], which are highly controversial at
present. Researchers have also made in-depth analyses based on the test results, constructed
many ultra-high-cycle-fatigue fracture analysis models of titanium alloys, and predicted
the life [21,22]. However, most studies have focused on the axial tension-compression
loading mode, which is not consistent with the bending vibration mode of aero-engine
blades, so whether these analytical models are suitable for three-point bending loading
remains to be further studied.
Forging is the most widely used method for the plastic forming of titanium alloys,
and it is also a key link in the blade manufacturing process. The forging performance
is mainly determined by the forging process. Once a bad microstructure is formed in
the forging process, the subsequent heat treatment is difficult to improve. In addition,
titanium alloys are very sensitive to the forging process, in which temperature determines
the solid phase transformation behavior of titanium alloys, and the deformation degree
and deformation rate also affect the morphology, size, proportion, and distribution of α
and β phases [23,24]. Tanaka et al. [25] studied the quantitative relationship between the
Ti-17 titanium alloy fatigue properties and microstructure at different forging temperatures.
It was found that the fatigue strength is closely related to the microstructure factors such as
the volume fraction of the equiaxed α phase, and it is one of the crack initiation sites during
low-temperature solid solution aging. The strength difference between the acicular α phase
and fine α + β phase is the main reason for crack initiation after high-temperature solid
solution aging. Nikitin et al. [26,27] studied the ultra-high-cycle-fatigue crack initiation
mechanism of forged VT3-1 titanium alloy under tension and torsion. It was found that
the traditional fatigue limit does not exist and the crack initiation position changes from
surface to subsurface. When the stress ratio is R = −1, the subsurface cracks initiate at the
strong defects, “macro region” boundaries, quasi-facets, and facets, while when the stress
ratio is R = 0.1, the cracks mainly initiate at the “macro region” boundaries. Sinha et al. [28]
studied the relationship between the fatigue life scatter and fracture mechanism of the
forged Ti-6242Si alloy, and clarified the microcrack initiation and propagation process by
the quantitative fatigue fracture characterization. Combined with the crystallographic
characteristics analysis of the crack initiation facet and adjacent facet, it was concluded
that the difference in cracking degrees along the primary α grains base plane leads to great
specimen fatigue life variability.
In this paper, the forged TC4 titanium alloy three-point bending fatigue properties
were studied. Equiaxed I, II and bimodal I, II were obtained by four forging processes.
In addition, 20 kHz room-temperature ultrasonic fatigue tests were carried out, and the
corresponding S-N curves and fatigue fractures were obtained. The crack initiation mech-
anism was revealed by fracture SEM + EDS analysis, and the microstructure effect on
fatigue properties was analyzed by the changing trend of S-N curves. The bimodal I life
prediction based on the energy method was also carried out. The work performed was
helpful to determine the forging process with the best anti-fatigue ability of aero-engine
blade materials.

2. Materials and Methods


2.1. Materials
The raw material was aviation-grade TC4 bar with a diameter of 28 mm, and its
chemical composition is shown in Table 1. The TC4 titanium alloy phase transition tem-
perature measured by the continuous temperature-increasing metallographic method was
998 ◦ C. On this basis, four process combinations of two temperatures (950 ◦ C and 985 ◦ C)
and two deformations (39.3% and 69.6%) were selected for plate die forging. The single
piece blanking size was 150 mm, and the bars were heated to the specified temperature

22
Materials 2021, 14, 5329

for 15~20 min and then flattened along the radial direction. After forging, the bars were
annealed at 720 ◦ C × 1 h + AC.

Table 1. Chemical composition of TC4 titanium alloy (wt%).

Al V Fe O C H N Ti
6.27 4.08 0.048 0.021 0.020 0.004 0.031 Bal.

Figure 1 shows the microstructures obtained by four forging processes. It was found
that the forging temperature has a great influence on the microstructure morphology, while
the effect of deformation is relatively small. The equiaxed primary α phase can be clearly
observed, and the remaining phases are the β transformation microstructure, including
the fine acicular secondary α phase and residual β phase. No obvious defects such as
crack, inclusion, segregation, and folding were found. The primary α content, grain size,
and standard deviation were measured, and the results are shown in Table 2. The equiaxed
I, II and bimodal I, II were obtained.

Figure 1. Microstructures of different forged TC4 titanium alloys. (a) Equiaxed I; (b) equiaxed II;
(c) bimodal I; (d) bimodal II.

Table 2. Measurement results of primary α phase content, grain size, and standard deviation.

Forging Forging Content Grain Size Standard


Microstructure
Process Type (%) (μm) Deviation (μm)
950 ◦ C/39.3% α+β 45.24 42.54 13.75 equiaxed I
950 ◦ C/69.6% α+β 48.74 34.68 16.80 equiaxed II
985 ◦ C/39.3% near β 27.54 23.62 19.92 bimodal I
985 ◦ C/69.6% near β 23.78 24.68 10.43 bimodal II

The room-temperature tensile properties of four kinds of forgings were obtained by


a WDW-300 universal testing machine, and the test results were averaged for 3 times.
The stress–strain curves were drawn according to the relationship between tension and
displacement during loading, as shown in Figure 2. Finally, the tensile strength σR , yield
strength σ0.2 , elongation A, section shrinkage Z, elastic modulus E, and Poisson’s ratio v
were obtained, as shown in Table 3.

23
Materials 2021, 14, 5329

Figure 2. Stress–strain curves of different forged TC4 titanium alloys.

Table 3. Room-temperature tensile properties of different forged TC4 titanium alloys.

Microstructure σR /MPa σ0.2 /MPa A/% Z/% E/GPa v


equiaxed I 967 898 16.2 46.8 115.85 0.317
equiaxed II 1003 951 15.8 47.6 116.83 0.329
bimodal I 986 907 14.2 51.1 120.29 0.332
bimodal II 998 924 15.2 49.5 118.91 0.324

2.2. Methods
Finite element modal analysis was used to carry out the specimen design, the linear
perturbation analysis step was selected after the three-dimensional model was introduced,
the first 10 free vibration modes were extracted, the stress and displacement were set as
output variables, and the structured hexahedral grid was divided. Finally, the modal
analysis results of the 7th-order three-point bending vibration were obtained, as shown in
Figure 3. The modal frequency was 19,954 Hz, which is close to the design frequency of
20 kHz. The middle area of the specimen bottom bore the maximum tensile stress, which is
the expected fracture position, and the two displacement standing points were symmetrical
with respect to the middle section, which are the fulcrum positions. By modifying the
specimen length to optimize the resonant frequency, it was determined that the four forged
specimen lengths L were 31.5, 31.6, 31.8, and 31.7 mm respectively, as shown in Figure 4.
The tests were carried out by using the HC-DF2020GD-K2 multi-function ultrasonic
fatigue testing machine (HC SONIC, Hangzhou, China). The loading frequency was
18.5~20.5 kHz, the static loading range was 0~10 kN, and the force sensor control accuracy
was 0.5%. The horn output amplitude was 10~150 μm, and the amplitude control precision
was 1 μm. The three-point bending loading process is shown in Figure 5. The middle
position of the specimen bottom bore both bending stress σm and vibration stress σa , and the
stress ratio R = (σm − σa )/(σm + σa ). The relationship between σm and static pressure P was
calibrated by a strain gauge. σa is proportional to the horn output amplitude A0 , and the
ratio was obtained from the stress–displacement nephogram in Figure 3. The fatigue
test started from the high-stress area, and the stress amplitude was gradually reduced
by 15 MPa until the cycle reached 1 × 109 , and each stress level was subjected to 1 to
2 samples. The output amplitudes under different ultrasonic powers were measured by

24
Materials 2021, 14, 5329

an LV-S01 laser vibrometer (SOPTOP, Ningbo, China) and used to calibrate the control
software. In order to restrain the temperature rise caused by high frequency and ensure
the stability of the fatigue test, the specimens were cooled by water mist.

Figure 3. Modal analysis nephogram of equiaxial I. (a) Stress distribution; (b) displacement distribution.

Figure 4. Design drawing of three-point bending specimen.

Figure 5. Schematic diagram of three-point bending loading process.

The Apreo S SEM (FEI, Hillsboro, OR, USA) was used to observe the fracture mor-
phology, and through the energy spectrum module installed on it, EDS analysis of the
ultra-high-cycle fatigue crack origin zones of the four microstructures was carried out.

25
Materials 2021, 14, 5329

3. Results
3.1. S-N Curves
As the number of loading cycles increases, the crack initiation changes from surface
to subsurface, which shows that the difference between ultra-high-cycle-fatigue and high-
cycle-fatigue lies not only in lower stress and longer life, but also in the fatigue failure
mechanism. The S-N curves of the four microstructures include the single-step decline
type, double-step decline type, and straight-step decline type, and there is no traditional
fatigue limit. The equiaxed high-cycle-fatigue properties are obviously better than those of
bimodal microstructures, while the equiaxed I ultra-high-cycle-fatigue property is worst.
The equiaxed II ultra-high-cycle-fatigue property is similar to that of bimodal I, but the
fatigue life scatter is larger, so it is considered that the bimodal ultra-high-cycle-fatigue
properties are better than those of equiaxed microstructures. From the analysis in Figure 6,
it seems that the high-cycle-fatigue and ultra-high-cycle-fatigue performances cannot
be obtained at the same time. In the range of 105 to 109 cycles, bimodal I has better
comprehensive fatigue performance, and 985 ◦ C with a 39.3% deformation degree is
determined to be the best forging process.

Figure 6. S-N curves of different forged TC4 titanium alloys.

3.2. Fatigue Fracture Morphology


Figure 7a–d show the high-cycle-fatigue crack origin zone SEM morphology of dif-
ferent forged TC4 titanium alloys at low magnification. It is observed that the river-like
crack propagation patterns converge on the specimen surface, in which the bimodal I has a
bright and continuous main crack, and the propagation direction is almost vertical. There
are no obvious inclusion traces in the origin zones, which is significantly different from the
fisheye morphology in high-strength steel. Figure 7e–h show the high-cycle-fatigue crack
origin zone SEM morphology at high magnification, and many smooth facet features are
observed, especially on the surface, which are considered to be the crack origins. There are
a large number of cleavage steps and microplastic tear (MPT) morphologies around the
facets. Overall, there is no obvious difference in the crack origin zone SEM morphology for
the four kinds of microstructures, and all of them are mainly transgranular brittle fracture.
Combined with the crack propagation path, it is determined that the crack first initiates
from the surface facet, initially propagates with the cleavage steps and MPT characteristics,
and then converges with the subsurface facet cracks to form the main crack, which finally
leads to fatigue fracture, which is shown as the transgranular quasi-cleavage fracture
mechanism.

26
Materials 2021, 14, 5329

Figure 7. SEM morphology of high-cycle-fatigue crack origin zone. (a,e) Equiaxed I, σa = 330 MPa,
Nf = 1.26 × 106 ; (b,f) equiaxed II, σa = 300 MPa, Nf = 4.25 × 106 ; (c,g) bimodal I, σa = 285 MPa,
Nf = 5.08 × 106 ; (d,h) bimodal II, σa = 270 MPa, Nf = 6.39 × 106 .

Figure 8a–d show the ultra-high-cycle-fatigue crack origin zone SEM morphology
of different forged TC4 titanium alloys at low magnification. The crack still propagates
in a river shape, but converges on the specimen subsurface, in which the equiaxed II
main crack patterns are the most obvious, and the propagation direction is oblique 45◦ .
Figure 8e–h show the ultra-high-cycle-fatigue crack origin zone SEM morphology at high
magnification, and the origin zone center is the facet cluster morphology, not the inclusion
characteristics. There are a large number of cleavage steps and MPT features around the
facets. The analysis shows that the near small facets converge into large facets through
the cleavage steps, while the distant large facets are connected by MPT characteristics,
and the whole is still dominated by transgranular brittle fracture. The distance between
the subsurface facet and the surface of the four kinds of microstructures is about 48, 39, 45,
and 43 μm. Combined with the crack propagation path, it is concluded that the crack first
originates from the subsurface facets, and then the faceted cracks with different heights

27
Materials 2021, 14, 5329

and orientations connect and propagate each other through the cleavage steps and MPT
features, and finally merge with the main crack to lead to fatigue fracture, which is still the
transgranular quasi-cleavage fracture mechanism.

Figure 8. SEM morphology of ultra-high-cycle-fatigue crack origin zone. (a,e) Equiaxed I,


σa = 218 MPa, Nf = 6.06 × 108 ; (b,d) equiaxed II, σa = 233 MPa, Nf = 3.16 × 108 ; (c,f) bimodal
I, σa = 225 MPa, Nf = 4.15 × 108 ; (g,h) bimodal II, σa = 233 MPa, Nf = 3.25 × 108 .

The fatigue crack initial propagation depth is about between 71 μm~273 μm, and then
it will enter the slow and stable propagation stage. The crack propagation region is
relatively flat, and the slow propagation zone shows a mixed shape of dense shear tear
edges and secondary cracks. Due to the low crack propagation rate, the fatigue band is
not obvious, but there are a large number of secondary cracks, indicating that the crack
direction changes after it encounters the second phase or grains with different orientations
in the crack propagation process. The stable propagation region is composed of many small
fault blocks with different sizes and heights, on which there exist thin and short fatigue
striations, as shown in Figure 9. A series of basically parallel and slightly curved fatigue

28
Materials 2021, 14, 5329

striations are clearly visible. Each fatigue striation represents the crack tip position under
the cycle, and the number of fatigue striations is roughly equal to the number of cycles
and perpendicular to the local crack propagation direction. The high-cycle-fatigue striation
spacings of the four microstructures are 0.097, 0.139, 0.219, and 0.140 μm, respectively,
and the ultra-high-cycle-fatigue striation spacings are 0.095, 0.087, 0.118, and 0.086 μm,
respectively. For the same microstructure, the distance between the fatigue striation
observation position and the origin zone is basically the same, so the fatigue striation
spacing in the local microzone basically reflects the crack propagation rate. The ultra-
high-cycle-fatigue striation spacing is smaller than that of high-cycle-fatigue, so the crack
propagation life is longer.

Figure 9. SEM morphology of fatigue crack stable propagation zone. (a) Equiaxed I, σa = 330 MPa,
Nf = 1.26 × 106 ; (b) equiaxed I, σa = 218 MPa, Nf = 6.06 × 108 ; (c) equiaxed II, σa = 300 MPa,
Nf = 4.25 × 106 ; (d) equiaxed II, σa = 233 MPa, Nf = 3.16 × 108 ; (e) bimodal I, σa = 285 MPa,
Nf = 5.08 × 106 ; (f) bimodal I, σa = 225 MPa, Nf = 4.15 × 108 ; (g) bimodal II, σa = 270 MPa,
Nf = 6.39 × 106 ; (h) bimodal II, σa = 233 MPa, Nf = 3.25 × 108 .

29
Materials 2021, 14, 5329

4. Discussion
4.1. Analysis of Crack Initiation Mechanism
For TC4 titanium alloy, Al and V are the stable elements of the α and β phase, re-
spectively, that is, the Al content in the α phase should be higher than that of the matrix,
and the V content should be lower. EDS analysis was carried out on the facets in the
ultra-high-cycle-fatigue crack origin zones of the four kinds of microstructures, and the
results are shown in Figure 10. Among the 24 randomly selected facets, the V content is
lower than the matrix level, and the Al content in most facets is higher than the matrix
level, while the Al content in a few facets is lower, and the missing part is replaced by the
Ti element. Considering that the facets are about tens of microns, slightly smaller than the
primary α grain size, it is concluded that the facets in the crack origin zone are formed
by the cleavage of primary α grains. The formation of these primary α grains is mainly
attributed to the Al element-promoting effect, while a few of them hardly need the Al
element assistance. Combined with SEM and EDS analysis, it is concluded that the crack
initiation mechanism is the cleavage of primary α grains on the surface or subsurface,
corresponding to high-cycle-fatigue and ultra-high-cycle-fatigue, respectively, and the
fatigue failure mechanism is mainly transgranular quasi-cleavage brittle fracture.
The facets in the crack origin zone appear in the form of clusters, rather than the single
facet. Chandran [29] believed that the spatial α grain agglomeration leads to the formation
of fatigue cracks in Ti-10V-2Fe-3Al titanium alloy. In order to investigate whether the TC4
titanium alloy fatigue cracks originate from a single facet or facet clusters, four kinds of
microstructure ultra-high-cycle-fatigue origin zone facet clusters were analyzed by EDS.
If the fatigue cracks originate in facet clusters, the Al content should be significantly higher
than the matrix level, and the V content should be lower. As shown in Figure 11, the three
element contents are very close to the matrix level, so it is concluded that facet clusters
are not the cause of fatigue crack formation. Kun et al. [30] obtained a similar conclusion
by analyzing the Ti-8Al-1Mo-1V titanium alloy inverse pole diagram along the loading
direction. It was found that the grain orientation is random on the whole, but the adjacent
grain orientation is basically the same in the local microzone, indicating that the fatigue
cracks originate from the preferred texture of primary α grains rather than clusters. To sum
up, the fatigue crack initiation mechanism is the cleavage of primary α grains with specific
spatial orientation and crystal orientation for the four kinds of microstructures.

30
Materials 2021, 14, 5329

Figure 10. EDS analysis of facets in ultra-high-cycle-fatigue crack origin zone. (a,e) Equiaxed I,
σa = 218 MPa, Nf = 6.06 × 108 ; (b,d) equiaxed II, σa = 233 MPa, Nf = 3.16 × 108 ; (c,f) bimodal I,
σa = 225 MPa, Nf = 4.15 × 108 ; (g,h) bimodal II, σa = 233 MPa, Nf = 3.25 × 108 .

31
Materials 2021, 14, 5329

Figure 11. EDS analysis of facet clusters in ultra-high-cycle-fatigue crack origin zone. (a) Equiaxed
I, σa = 218 MPa, Nf = 6.06 × 108 ; (b) equiaxed II, σa = 233 MPa, Nf = 3.16 × 108 ; (c) bimodal I,
σa = 225 MPa, Nf = 4.15 × 108 ; (d) bimodal II, σa = 233 MPa, Nf = 3.25 × 108 .

4.2. Analysis of Microstructure Influence


The titanium alloy fatigue properties are mainly affected by the microstructure such
as primary α grains and are closely related to the forging process. On the one hand,
the increased forging temperature will strengthen the primary α phase diffusion behavior,
swallow the surrounding fine α phases, and cause the primary α grains to grow. On the
other hand, the increased forging temperature will promote the α→β phase transformation
behavior, resulting in the decrease in primary α grain size and content. The two mech-
anisms compete with each other under the deformation degree influence, resulting in
changes in the primary α content and grain size, as shown in Table 2.
As the α phase is more brittle than the β phase is, it is easier to form dislocation
accumulation at the α-β phase boundaries or primary α grain boundaries under fatigue
loading. The larger primary α grain boundaries increase the dislocation slip length and
aggravate the stress concentration at the tip of dislocation accumulation, thus promoting
the primary α grain cleavage failure [31]. Stanzl et al. [32] believed that the accumulated
irreversible slip in the process of low stress fatigue leads to the α grain cleavage fractures
in the slip surface with a high Schmidt factor, and the subsequent process of faceted crack
nucleation and propagation form a rough origin zone. Chai et al. [33] studied the subsurface
defect-free crack initiation mechanism and obtained a similar conclusion: the primary α
phase is the weak crack initiation part, and the longer primary α grain boundaries play the
role of internal notches.
The analysis shows that the fatigue life is related to the microstructure influence on
the high-cycle-fatigue and ultra-high-cycle-fatigue mechanism. The traditional high-cycle-
fatigue belongs to macro-scale fatigue, the material composition can be approximately
regarded as a uniform distribution, the fatigue failure is mainly controlled by surface stress,
and the microstructure influence is small. At the same stress level, the primary α phase
content in the equiaxed microstructure is higher than that in the bimodal microstructure,
which can tolerate more slip deformation and enhance the resistance to crack initiation,
so the high-cycle-fatigue performance is better. In contrast, the ultra-high-cycle-fatigue
belongs to micro-scale fatigue, material composition is no longer uniformly distributed,
fatigue failure is controlled by both surface stress and internal defects, and the microstruc-
ture influence is greater. The lower primary α phase content in the bimodal microstructure

32
Materials 2021, 14, 5329

means fewer defects, which reduces the number of stress concentration zones caused by
dislocation accumulation at the second phase interface, makes the slip deformation more
uniform, and increases the crack initiation life. In addition, the β transition tissue content
in the bimodal microstructure is higher and the creep resistance is greater. As shown in
Figure 1, the strip-like and fine needle-like secondary α phases are arranged longitudinally
and horizontally in the β matrix, isolating the primary α grains from each other, making the
crack propagation path more tortuous, blocking the facet crack propagation, converging
and combining into the main crack, and increasing the crack propagation life, so that the
bimodal ultra-high-cycle-fatigue properties are better.

4.3. Fatigue Life Prediction Based on Energy


In the linear elastic fracture mechanics, the stress intensity factor amplitude ΔK at the
crack tip is the main crack propagation control parameter, which can be calculated by the
Murakami model [34], as shown in Formula (1):

ΔK f ,orFC = n · σa (π area f ,orFC )1/2 (1)

For surface cracks, n = 0.65; for subsurface cracks, n = 0.5; σa is the stress amplitude;
and areaf,orFC represents the projected area of facet or facet clusters in the principal stress
direction. Based on the measurement of the facet characteristics in the bimodal I crack
origin zone, it is found that there is no obvious difference in most facet sizes, with an aver-
age of about 17.5 μm, while the facet cluster size increases with the fatigue life. The stress
intensity factors ΔKf and ΔKFC of facets and faceted clusters are calculated according to
Formula (1). As shown in Figure 12, ΔKf decreases with the fatigue life, which is much
smaller than the stress intensity factor threshold value for macroscopic crack propagation
(ΔKth = 5~6 MPa·m1/2 ). The ΔKFC remains basically unchanged with the increased fatigue
life, and the ΔKFC is about 4.53 MPa·m1/2 for surface failure and 5.49 MPa·m1/2 for subsur-
face failure, which is mainly attributed to the adverse environmental effects. Surface cracks
germinate and propagate in air and water fog, while subsurface cracks do so in vacuum,
so the ΔKFC is lower when surface failure occurs. To sum up, the amplitude of the stress
intensity factor ΔKFC of faceted clusters is analogized to the threshold ΔKth , which controls
the macroscopic stable crack propagation.

Figure 12. Calculation of stress intensity factor amplitude. (a) ΔKf and Nf ; (b) ΔKFC and Nf .

33
Materials 2021, 14, 5329

For high-cycle-fatigue and ultra-high-cycle-fatigue regimes, the crack initiation and


initial propagation stages occupy the vast majority, so the fatigue life can be estimated by
the energy-based crack nucleation life model, and the calculation formula is [35]:

4πμ3 h2 M3 c
Ni =   (2)
0.005d3 (Δσ − 2Mk) π 2 Δσ2 (1 − v)2 + ξ M2 μ2

where μ is the shear homogeneous matrix modulus, h is the slip band width, d is the
primary α grain size, Δσ is the stress amplitude, 2 Mk is defined as the fatigue strength
at 109 cycles, M is usually taken as 2, v is Poisson’s ratio, and ξ is a numerical constant.
It is known from Figure 12 that the stress intensity factor amplitude of the faceted clusters
ΔKFC is approximately constant, so the crack size c is represented by the facet cluster
size areaFC 1/2 . According to the S-N curve, the unknown parameters ξ and h are fitted
nonlinearly, and the results are 1.02 and 8.25 × 10−4 , respectively. Finally, the bimodal I
fatigue life prediction curve is obtained, as shown in Figure 13.

Figure 13. Results of the fatigue crack life prediction.

The life prediction curve related to the facet cluster size is basically consistent with
the fatigue data, which verifies that the fatigue failure process is mainly consumed in
the facet cluster formation stage, and only a small part of the propagation zone and final
rupture zone. It is also shown that the energy-based life prediction model is not only
suitable for axial tension–compression loading, but also for three-point bending loading.
The analysis shows that the crack origin zone shows similar facet morphology under the
two loading modes, and there is no essential difference in fatigue failure mechanism, so the
relevant characteristic parameters in the life prediction model are universal. As far as
the current research is concerned, it is reasonable to apply the axial tension–compression
fatigue analysis model to three-point bending fatigue, and the author will further analyze
and demonstrate this statement in the follow-up research.

5. Conclusions
In this paper, forged TC4 titanium alloy ultrasonic fatigue properties under three-point
bending were studied, and the main conclusions are as follows:
1. As the number of loading cycles increases, the crack initiation changes from surface
to subsurface initiation. The S-N curves have different shapes, and there is no tra-
ditional fatigue limit. The equiaxed high-cycle-fatigue performance is better, while
the bimodal ultra-high-cycle-fatigue performance is better. The best forging process
combination is 985 ◦ C with a 39.3% deformation degree.
2. The crack origin zones in the four kinds of microstructures show the mixed mor-
phology of facets, cleavage steps, and MPT characteristics. The crack originates
from the surface or subsurface facet, showing a transgranular quasi-cleavage fracture

34
Materials 2021, 14, 5329

mechanism. The ultra-high-cycle-fatigue striation spacing is smaller than that of


high-cycle-fatigue for the same microstructure, which reflects a higher crack propaga-
tion life.
3. The facet in the crack origin zone is formed by the cleavage of primary α grains.
Most of these primary α grains are attributed to the Al element-promoting effect,
while a few of them hardly need the Al element assistance. The fatigue cracks
originate from the preferred texture of primary α grains, rather than the primary α
grain clusters.
4. High-cycle-fatigue belongs to macro-scale fatigue, the fatigue failure is mainly con-
trolled by surface stress, and the influence of the microstructure is small. While
ultra-high-cycle-fatigue belongs to micro-scale fatigue, the fatigue failure is controlled
by both surface stress and internal defects, and the influence of the microstructure is
great. This makes the equiaxed and bimodal microstructure have better high-cycle-
fatigue and ultra-high-cycle-fatigue performance, respectively.
5. The stress intensity factor amplitude ΔKFC of faceted clusters can be analogized to
the threshold ΔKth to control the macroscopic stable crack propagation, and the life
prediction curve related to the faceted cluster size is basically consistent with the
fatigue data, which shows that it is reasonable to apply the axial tension–compression
fatigue analysis model to three-point bending fatigue, which is attributed to the
similar fatigue failure mechanism.

Author Contributions: B.W. and L.C. designed the experiments; B.W. and D.L. performed the
experiments and funding acquisition; L.C., B.W. and D.L. analyzed the experimental results; B.W.
wrote the paper. All authors have read and agreed to the published version of the manuscript.
Funding: This research was funded by the National Key Basic Research and Development Program
of China (973), grant number 2015CB057400.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Zimmermann, M. Diversity of damage evolution during cyclic loading at very high numbers of cycles. Int. Mater. Rev. 2012, 57,
73–91. [CrossRef]
2. Mayer, H. Recent developments in ultrasonic fatigue. Fatigue Fract. Eng. Mater. Struct. 2016, 39, 3–29. [CrossRef]
3. Oguma, H.; Nakamura, T. Formation mechanism of specific fracture surface region in the sub-surface fracture of titanium alloy.
Adv. Mater. Res. 2014, 891–892, 1436–1441. [CrossRef]
4. Günther, J.; Krewerth, D.; Lippmann, T.; Leuders, S.; Tröster, T.; Weidner, A.; Biermann, H.; Niendorf, T. Fatigue life of additively
manufactured Ti-6Al-4V in the very high cycle fatigue regime. Int. J. Fatigue 2017, 94, 236–245. [CrossRef]
5. Uematsu, Y.; Kakiuchi, T.; Hattori, K. EBSD assisted fractography of subsurface fatigue crack initiation mechanism in the
ultrasonic shot peened βeta-type titanium alloy. Fatigue Fract. Eng. Mater. Struct. 2018, 41, 2239–2248. [CrossRef]
6. Furuya, Y.; Nishikawa, H.; Hirukawa, H.; Nagashima, N.; Takeuchi, E. Catalogue of NIMS fatigue data sheets. Sci. Technol. Adv.
Mater. 2019, 20, 1055–1072. [CrossRef] [PubMed]
7. Hong, Y.; Sun, C.; Liu, X. A review on mechanisms and models for very-high-cycle fatigue of metallic materials. Adv. Mech. 2018,
48, 1–65.
8. Sakai, T.; Nakagawa, A.; Oguma, N.; Nakamura, Y.; Ueno, A.; Kikuchi, S.; Sakaida, A. A review on fatigue fracture modes of
structural metallic materials in very high cycle regime. Int. J. Fatigue 2016, 93, 339–351. [CrossRef]
9. Bao, X.; Cheng, L.; Ding, J.; Chen, X.; Lu, K.; Cui, W. The effect of microstructure and axial tension on three-point bending fatigue
behavior of TC4 in high cycle and very high cycle regimes. Materials 2020, 13, 68. [CrossRef]
10. Yang, K.; Zhong, B.; Huang, Q.; He, C.; Huang, Z.Y.; Wang, Q.; Liu, Y.J. Stress Ratio and Notch Effects on the Very High Cycle
Fatigue Properties of a Near-Alpha Titanium Alloy. Materials 2018, 11, 1778. [CrossRef]
11. Hong, Y.; Lei, Z.; Sun, C. Propensities of crack interior initiation and early growth for very-high-cycle fatigue of high strength
steels. Int. J. Fatigue 2014, 58, 144–151. [CrossRef]
12. Pan, X.; Hong, Y. High-cycle and very-high-cycle fatigue behaviour of a titanium alloy with equiaxed microstructure under
different mean stresses. Fatigue Fract. Eng. Mater. Struct. 2019, 42, 1950–1964. [CrossRef]
13. Oguma, H.; Nakamura, T. The effect of microstructure on very high cycle fatigue properties in Ti-6Al-4V. Scr. Mater. 2010, 63,
32–34. [CrossRef]
14. Jha, S.K.; Szczepanski, C.J.; Golden, P.J.; Porter, I.I.I.W.J.; John, R. Characterization of fatigue crack-initiation facets in relation to
lifetime variability in Ti-6Al-4V. Int. J. Fatigue 2012, 42, 248–257. [CrossRef]

35
Materials 2021, 14, 5329

15. Liu, X.; Chen, E.; Zeng, F. Mechanisms of interior crack initiation in very-high-cycle fatigue of high-strength alloys. Eng. Fract.
Mech. 2019, 212, 153–163. [CrossRef]
16. Liu, X.; Sun, C.; Hong, Y. Faceted crack initiation characteristics for high-cycle and very-high-cycle fatigue of a titanium alloy
under different stress ratios. Int. J. Fatigue 2016, 92, 434–441. [CrossRef]
17. Wu, Y.; Liu, J.; Wang, H.; Guan, S.; Yang, R.; Xiang, H. Effect of stress ratio on very high cycle fatigue properties of Ti-10V-2Fe-3Al
alloy with duplex microstructure. J. Mater. Sci. Technol. 2018, 34, 1189–1195. [CrossRef]
18. Huang, Z.Y.; Liu, H.Q.; Wang, H.M.; Wagner, D.; Khan, M.K.; Wang, Q.Y. Effect of stress ratio on VHCF behavior for a compressor
blade titanium alloy. Int. J. Fatigue 2016, 93, 232–237. [CrossRef]
19. Yang, K.; He, C.; Huang, Q.; Huang, Z.Y.; Wang, C.; Wang, Q.; Liu, Y.J.; Zhong, B. Very high cycle fatigue behaviors of a turbine
engine blade alloy at various stress ratios. Int. J. Fatigue 2017, 99, 35–43. [CrossRef]
20. Furuya, Y.; Takeuchi, E. Gigacycle fatigue properties of Ti-6Al-4V alloy under tensile mean stress. Mater. Sci. Eng. A 2014, 598,
135–140. [CrossRef]
21. Nie, B.; Zhao, Z.; Chen, D. Effect of basketweave microstructure on very high cycle fatigue behavior of TC21 titanium alloy.
Metals 2018, 8, 401. [CrossRef]
22. Li, W.; Li, M.; Sun, R.; Xing, X.; Wang, P.; Sakai, T. Faceted crack induced failure behavior and micro-crack growth based strength
evaluation of titanium alloys under very high cycle fatigue. Int. J. Fatigue 2020, 131, 105369. [CrossRef]
23. Deng, R.; Yang, G.; Mao, X.; Zhang, P.; Han, D. Effects of Forging Process and Following Heat Treatment on Microstructure and
Mechanical Properties of TC11 Titanium Alloy. Mater. Mech. Eng. 2011, 35, 58–62.
24. Pan, Q.; Yu, X.; Qi, Y.; Huang, Q.; Pan, G. Effect of Thermal Tensile Deformation and Solution Aging Treatment on Microstructure
of TC4-DT Titanium Alloy. Mater. Mech. Eng. 2019, 43, 21–26.
25. Tanaka, S.; Akahori, T.; Niinomi, M.; Nakai, M. Relationship between Microstructure and Fatigue Properties of Forged Ti-5Al-
2Sn-2Zr-4Mo-4Cr for Aircraft Applications: Engineering Materials and Their Applications. Mater. Trans. 2020, 61, 2017–2024.
[CrossRef]
26. Nikitin, A.; Bathias, C.; Palin-Luc, T. A new piezoelectric fatigue testing machine in pure torsion for ultrasonic gigacycle fatigue
tests: Application to forged and extruded titanium alloys. Fatigue Fract. Eng. Mater. Struct. 2015, 38, 1294–1304. [CrossRef]
27. Nikitin, A.; Palin-Luc, T.; Shanyavskiy, A. Crack initiation in VHCF regime on forged titanium alloy under tensile and torsion
loading modes. Int. J. Fatigue 2016, 93, 318–325. [CrossRef]
28. Sinha, V.; Pilchak, A.L.; Jha, S.K.; Porter, W.J.; John, R.; Larsen, J.M. Correlating Scatter in Fatigue Life with Fracture Mechanisms
in Forged Ti-6242Si Alloy. Metall. Mater. Trans. A 2018, 49, 1061–1078. [CrossRef]
29. Chandran, K.S.R. Duality of fatigue failures of materials caused by Poisson defect statistics of competing failure modes. Nat.
Mater. 2005, 4, 303–308. [CrossRef]
30. Yang, K.; Zhong, B.; Huang, Q.; He, C.; Huang, Z.Y.; Wang, Q.; Liu, Y.J. Stress ratio effect on notched fatigue behavior of a
Ti-8Al-1Mo-1V alloy in the very high cycle fatigue regime. Int. J. Fatigue 2018, 116, 80–89. [CrossRef]
31. JZuo, H.; Wang, Z.G.; Han, E.H. Effect of microstructure on ultra-high cycle fatigue behavior of Ti-6Al-4V. Mater. Sci. Eng. A 2007,
473, 147–152.
32. Stanzl-Tschegg, S.; Mughrabi, H.; Schoenbauer, B. Life time and cyclic slip of copper in the VHCF regime. Int. J. Fatigue 2007, 29,
2050–2059. [CrossRef]
33. Chai, G.; Zhou, N.; Ciurea, S.; Andersson, M.; Peng, R.L. Local plasticity exhaustion in a very high cycle fatigue regime. Scr.
Mater. 2012, 66, 769–772. [CrossRef]
34. Murakami, Y.; Kodama, S.; Konuma, S. Quantitative evaluation of effects of non-metallic inclusions on fatigue strength of high
strength steels. I: Basic fatigue mechanism and evaluation of correlation between the fatigue fracture stress and the size and
location of non-metallic inclusions. Int. J. Fatigue 1989, 11, 291–298. [CrossRef]
35. Li, W.; Xing, X.; Gao, N.; Wang, P. Subsurface crack nucleation and growth behavior and energy-based life prediction of a titanium
alloy in high-cycle and very-high-cycle regime. Eng. Fract. Mech. 2019, 221, 106705. [CrossRef]

36
materials
Article
Comparison of Hydrogen Embrittlement Susceptibility of
Different Types of Advanced High-Strength Steels
Sangwon Cho 1,† , Geon-Il Kim 1,† , Sang-Jin Ko 1 , Jin-Seok Yoo 1 , Yeon-Seung Jung 2 , Yun-Ha Yoo 2
and Jung-Gu Kim 1, *

1 School of Advanced Materials Science and Engineering, Sungkyunkwan University (SKKU),


Suwon 16419, Korea; [email protected] (S.C.); [email protected] (G.-I.K.);
[email protected] (S.-J.K.); [email protected] (J.-S.Y.)
2 Steel Solution Research Laboratory, POSCO Global R&D Center, Incheon 21985, Korea;
[email protected] (Y.-S.J.); [email protected] (Y.-H.Y.)
* Correspondence: [email protected]; Tel.: +82-31-290-7360
† These authors contributed equally to this study.

Abstract: This study investigated the hydrogen embrittlement (HE) characteristics of advanced
high-strength steels (AHSSs). Two different types of AHSSs with a tensile strength of 1.2 GPa were
investigated. Slow strain rate tests (SSRTs) were performed under various applied potentials (Eapp )
to identify the mechanism with the greatest effect on the embrittlement of the specimens. The SSRT
results revealed that, as the Eapp increased, the elongation tended to increase, even when a potential
exceeding the corrosion potential was applied. Both types of AHSSs exhibited embrittled fracture
behavior that was dominated by HE. The fractured SSRT specimens were subjected to a thermal
desorption spectroscopy analysis, revealing that diffusible hydrogen was trapped mainly at the
grain boundaries and dislocations (i.e., reversible hydrogen-trapping sites). The micro-analysis
results revealed that the poor HE resistance of the specimens was attributed to the more reversible
Citation: Cho, S.; Kim, G.-I.; Ko, S.-J.;
Yoo, J.-S.; Jung, Y.-S.; Yoo, Y.-H.; Kim,
hydrogen-trapping sites.
J.-G. Comparison of Hydrogen
Embrittlement Susceptibility of Keywords: advanced high-strength steel; hydrogen embrittlement; hydrogen trapping; thermal
Different Types of Advanced desorption spectroscopy
High-Strength Steels. Materials 2022,
15, 3406. https://doi.org/10.3390/
ma15093406
1. Introduction
Academic Editors: Lucjan Śnieżek
and Jaroslaw Galkiewicz The need to reduce environmental harm is a growing global concern. Accordingly,
the automotive industry is striving to improve fuel efficiency and reduce carbon dioxide
Received: 15 April 2022
emissions to protect the environment. Globally, the industry is pushing for fuel efficiency
Accepted: 7 May 2022
improvements via two routes: high-efficiency engines and lightweight body designs [1–3].
Published: 9 May 2022
A vehicle’s body weight accounts for 40% of its fuel efficiency factors; therefore, reduc-
Publisher’s Note: MDPI stays neutral ing this weight has the greatest impact on improving fuel efficiency. Generally, a 100-kg
with regard to jurisdictional claims in reduction in body weight lowers carbon dioxide emissions by 7.5 to 12.5 g/km, signifi-
published maps and institutional affil- cantly enhancing the fuel efficiency. To reduce the weight of cars, manufacturers may use
iations. nonferrous materials (e.g., resin, aluminum alloy, and magnesium alloy) [4–8]; specific
methods (e.g., the miniaturization of parts); or different types of high-strength steels [9–11].
Although nonferrous materials used in automotive structures such as aluminum alloy and
magnesium alloy are lighter than steel, they are also weaker, and their thicknesses must
Copyright: © 2022 by the authors.
be increased to maintain body stiffness. Additionally, lightweight materials must be used
Licensee MDPI, Basel, Switzerland.
in combination with other materials, such as carbon fiber-reinforced plastic, to maintain
This article is an open access article
distributed under the terms and
the required body stiffness. Therefore, research and development into various types of
conditions of the Creative Commons
advanced high-strength steels (AHSSs) are currently underway.
Attribution (CC BY) license (https://
Generally, the mechanical properties of steel are enhanced using methods such as
creativecommons.org/licenses/by/ solid solution, grain refinement, or precipitation; however, in the case of AHSS, phase
4.0/). transformation-based methods are also used. Enhancing the mechanical properties of an

Materials 2022, 15, 3406. https://doi.org/10.3390/ma15093406 37 https://www.mdpi.com/journal/materials


Materials 2022, 15, 3406

AHSS increases its corrosivity and sensitivity to a delayed fracture, i.e., stress–corrosion
cracking (SCC) and hydrogen embrittlement (HE), which are the main problems associated
with AHSS [12].
Over the past decades, a lot of research on HE in AHSSs has been conducted: effects
of the strength, microstructure, and types of defects of AHSSs. V. G. Khanzhin et al. [13,14]
studied the influence of precipitate and mechanical properties on HE. According to the
studies, the higher density of precipitates in a structure, the lower HE resistance, since the
secondary phase particles influence both the stage of initiating hydrogen cracks and the
crack growth kinetics to a critical value. Additionally, mechanical properties, their strength
and toughness, affects the nucleation of hydrogen cracks, possibility of their propagation,
and the kinetics of growth to a critical size.
In AHSS, the delayed fracture phenomenon is caused mainly by HE. Hydrogen inside
the steel is preferentially trapped in lattice defects, such as voids, dislocations, and grain
boundaries, as well as in various carbides and precipitates [15,16]. Additionally, after
entrapment, hydrogen is concentrated in certain areas by stress, leading to the propagation
of internal cracks and, eventually, to delayed fractures. However, the exact cause and
mechanism of the delayed fracture phenomenon have not been identified to date. This is
because, in addition to HE, a delayed fracture can result from the combined effects of other
variables, including the stresses acting on the steel, the microstructures, mechanical proper-
ties, surface conditions, and internal cracks. Further research is required to determine the
exact cause of HE. Therefore, this study uses slow strain rate tests (SSRTs), a microstruc-
tural analysis, and a thermal desorption spectroscopy (TDS) analysis to investigate the HE
mechanisms of two different types of AHSSs with the tensile strength of 1.2 GPa.

2. Materials and Methods


This study used two different types of AHSSs with a tensile strength of 1.2 GPa. Table 1
provides their chemical compositions. Figure 1 presents the specimens’ microstructural
images, and Figure 2 shows the mechanical properties obtained by the tensile test. Steel A
was comprised of fine grains with complex phases of ferrite, bainite, martensite, and a small
fraction of retained austenite, with Ti and/or Nb precipitates for enhancing the tensile
strength and ductility. Since this steel was cooled slowly after soaking in the austenite
region, its main phases were bainite and martensite. Steel B was also a multiphase AHSS
comprising ferrite, bainite, a relatively higher fraction of retained austenite, and a small
portion of martensite. Under an applied stress, the phase transformation of the retained
austenite increased the ductility of Steel B. As shown in Figure 2c, Steel B showed a uniform
strain-hardening rate range, which is evidence of a transformation-induced plasticity effect.

Table 1. Chemical compositions of the advanced high-strength steels for use in automobiles (wt.%).

Component C Si Mn Cr Ti Nb Fe
Steel A 0.11~0.18 0.4~0.7 2.2~2.7 0.0~0.1 0.01~0.02 0.01~0.02 Bal.
Steel B 0.11~0.18 1.2~1.9 2.4~2.7 0.20~0.45 0.01~0.02 - Bal.

Figure 1. Scanning electron microscopy images of the microstructure of (a) Steel A and (b) Steel B
etched with a 2% nital solution.

38
Materials 2022, 15, 3406

Figure 2. (a) Engineering stress–strain curves, (b) true stress–strain curves, and (c) strain-hardening
rate vs. true strain curves for Steel A and Steel B obtained by tensile tests.

2.1. Electrochemical Tests


Potentiodynamic polarization test was performed to analyze the corrosion behavior of
AHSSs and determine the applied potentials in the SSRTs. For the electrochemical tests,
the specimens were cut into dimensions of 1.5 × 1.5 cm2 , abraded up to #600 with silicon–
carbide paper, degreased with ethanol, and dried with N2 gas. The electrochemical test
environment used a modified Society of Automotive Engineers’ (M-SAE) solution at 25 ◦ C
(room temperature) (see Table 2 for the chemical compositions). All the electrochemical
experiments were performed with a triple-electrode electrochemical cell, as shown in
Figure 3. The counter electrode was a graphite rod, and the reference electrode was
a saturated calomel electrode (SCE). The open-circuit potential (OCP) was established
over 3 h. Potentiodynamic polarization tests were conducted with a potential sweep of
0.166 mV/s in accordance with ASTM G5. After the samples were stabilized in an M-SAE
solution at room temperature for 1 h, SSRTs were conducted under applied potentials (Eapp )
of −600, −750, and −1500 mVSCE based on the potentiodynamic polarization test.

Table 2. Chemical composition (wt.%) of the modified Society of Automotive Engineers’ solution.

NaCl CaCl2 NaHCO3 (NH4 )2 SO4 pH


0.5 0.1 0.075 0.35 7.3

39
Materials 2022, 15, 3406

Figure 3. Schematic diagram of the three-electrode cell configuration used in (a) potentiodynamic
polarization test and (b) SSRTs. CE, RE, and WE refer to counter electrode, reference electrode, and
working electrode, respectively.

2.2. Slow Strain Rate Tests


A schematic image of the specimen for the SSRTs is presented in Figure 4. First, the
critical strain rate was determined by various SSRTs in the OCP state; then, under the
Eapp values listed in Section 2.1, SSRTs were conducted at a strain rate of 10−5 . After
the tests, each fractured specimen was cleaned with ethanol and transferred into liquid
nitrogen as soon as possible. Then, the fracture surface was observed via scanning electron
microscopy (SEM), and the hydrogen desorption rate was determined by a TDS analysis
of the hydrogen content charged into the specimen. To enable the TDS analysis of the
hydrogen content, the specimen was cut up to 10 mm from the fracture surface. To calculate
the activation energy for hydrogen de-trapping, heating rates of 2 ◦ C/min and 4 ◦ C/min
were used.

Figure 4. Dimensions of the specimen used in slow strain rate tests.

2.3. Analyses for Hydrogen-Trapping Sites


The grain boundary areas and austenite phase fractions of the samples were measured
using electron backscattered diffraction (EBSD). X-ray diffraction (XRD) was performed at
a scan rate of 1◦ /min, and the dislocation density was calculated using the full width at
half-maximum (FWHM). Electron probe microanalysis (EPMA) and transmission electron
microscopy (TEM) were used to analyze the type and characteristics of the precipitates.
Before EPMA, each specimen was etched slightly with a 2% nital solution for 5 s.

40
Materials 2022, 15, 3406

3. Results and Discussion


3.1. Potentiodynamic Polarization Test
This study conducted potentiodynamic polarization tests to analyze the elec-tro-
chemical properties of the specimens; the results are presented in Figure 5 and Table 3.
Mild steel with a tensile strength of 270 MPa was used as a comparison material for the
AHSS. The AHSSs exhibited higher corrosion rates than the mild steel. The anodic polar-
ization curves of the AHSSs and the mild steel material were similar in shape; however,
the ca-thodic polarization curve of the AHSSs shifted more to the right compared with the
mild steel. This was because hydrogen evolution reactions are more likely in AHSS than in
mild steel due to the higher levels of precipitates, carbides, and grain boundary densi-ties.
Additionally, there are more phase types in AHSS compared with mild steel, result-ing in
higher corrosion rates due to the large micro-galvanic effect between phases [17].

Figure 5. Potentiodynamic polarization curves in a modified Society of Automotive Engineers’


solution.

Table 3. Potentiodynamic polarization test results.

Corrosion
Ecorr Icorr βa βc
Specimen Rate
(VSCE ) (A/cm2 ) (mV/Decade) (mV/Decade)
(mm/y)
Steel A −0.748 3.54 × 10−5 92 205 0.41
Steel B −0.755 3.38 × 10−5 61 241 0.39
Mild steel −0.765 1.70 × 10−5 89 419 0.20

The electrochemical properties of both AHSSs, e.g., corrosion potential and corrosion
current density, were almost identical. From the polarization curve of Steel A, the redox
reaction of hydrogen (Equation (1)) produced a higher hydrogen equilibrium potential
(E0 H2 O/H2 , −0.672 VSCE ) than corrosion potential (Ecorr , −0.728 VSCE ). Therefore, hydrogen
was also generated at the corrosion potential.

2H2 O + 2e− → H2 + 2OH− , E0H2 O/H2 = −0.672VSCE . (1)

3.2. Slow Strain Rate Tests


During an SSRT, the applied strain rate will cause differences in the occurrence of SCC
and HE behaviors [18]. When the strain rate is too high, there is insufficient time for SCC to
occur, resulting in only a tensile rupture. Conversely, at a low strain rate, the re-passivation

41
Materials 2022, 15, 3406

of the film before the propagation of a crack by anodic dissolution means that SCC does
not occur. However, HE does not require the breakdown of the passive film; instead, it
is caused by hydrogen trapping inside the steel. Therefore, the lower strain rate requires
more time for the hydrogen to be entrapped in the steel, making it more susceptible to HE.
Accordingly, when conducting SSRTs to determine the HE characteristics, an optimal strain
rate should be applied in consideration of the hydrogen trapping time. To determine the
critical strain rate, the SSRTs were performed at strain rates of 2 × 10−4 /s, 10−4 /s, 10−5 /s,
10−6 /s, and 5 × 10−7 /s in M-SAE solution under an OCP state (see Figure 6 for the results).
When the strain rate was 10−5 /s, the SSRT results revealed a relatively low elongation for
both AHSSs. Therefore, the final SSRTs were performed under a strain rate of 10−5 /s.

Figure 6. Elongation vs. strain rate in a modified Society of Automotive Engineers’ solution under an
open-circuit potential state.

Both specimens showed a decreasing elongation with a decreasing strain rate, which is
representative of a typical HE elongation–strain rate curve, in which a ductility minimum
is not expected [18]. A low strain rate provides sufficient time for lattice diffusion, which
allows hydrogen to easily enter the trapping sites.
To minimize the hydrogen generation reactions, the applied anodic potential should
be higher than the hydrogen reduction potential. Accordingly, as is shown in Figure 7 and
Table 4, the SSRTs in this study were conducted using various Eapp values. The amounts of
hydrogen evolution for the Eapp values of −1500 mVSCE were calculated by integrating the
base area of the current–time curves (Figure 7c,d). Since the current obtained with an Eapp
above −750 mVSCE was caused by corrosion, the hydrogen evolution amounts for Eapp
values of −750 mVSCE were calculated using Faraday’s law as follows [18]:

Ired,H2 O/H2 × t × a
m= (2)
n×F
where m is the reaction mass (hydrogen evolution amount, in grams), Ired is the current of
the reduction reaction at each Eapp (A), t is the time to fracture (s), a is the atomic weight
(g/mol), n is the number of electrons (n = 2 for Equation (1)), and F is the Faraday constant
(96,500 C/mol).

42
Materials 2022, 15, 3406

Figure 7. Stress–strain curves for (a) Steel A and (b) Steel B, and the current–time curves of (c) Steel
A and (d) Steel B obtained during slow strain rate tests.

Table 4. Slow strain rate test results.

Applied Yield Tensile HE Susceptibility Hydrogen


Elongation
Steel Potential Strength Strength Index, IHE Evolution Rate
(%)
(mVSCE ) (MPa) (MPa) (%) (g)
Air 781 1186 16.0 - -
Steel −600 703 1124 14.1 11.9 -
A −750 (Ecorr ) 699 1146 11.2 30.1 5.76 × 10−6
−1500 690 1041 4.6 71.0 1.23 × 10−3
Air 800 1202 21.6 - -
Steel −600 906 1150 16.4 24.0 -
B −750 (Ecorr ) 852 1119 13.3 38.3 2.62 × 10−6
−1500 851 1034 3.7 82.9 9.76 × 10−4

For both AHSSs, the lower Eapp was found to be correlated with a reduced elongation.
There was an increase in the amount of corrosion with a higher anodic overvoltage, while
the hydrogen evolution amount increased with the increasing cathodic overvoltage. A
slightly higher amount of hydrogen was generated on Steel A compared to on Steel B.
The HE sensitivity index (IHE ) indicates the ductility loss of the AHSSs according
to the Eapp . Since the ductility loss of AHSSs with cathodic applied potential is related
to HE, the IHE was used to compare the HE resistance. The IHE can be calculated using
Equation (3), in which a higher IHE is associated with increased HE sensitivities. The IHE of

43
Materials 2022, 15, 3406

Steel B was approximately 10% higher than that of Steel A. Therefore, compared with Steel
B, Steel A had a superior HE resistance.

εair − εsoln.
IHE = × 100 (3)
εair

where IHE is the HE sensitivity index (%), εair is the elongation tested in air, and εsoln. is the
elongation tested under an Eapp .

3.3. Fractography
To determine the fracture properties of the AHSS samples, after the SSRTs were
conducted, the fracture surfaces and sides of the specimens were observed by SEM. The
results are presented in Figures 8 and 9. Cracks were initiated and propagated from the
sides in all the specimens. In Steel A, uniform pitting corrosion was observed on the sides at
−600 mVSCE , while there was no changes at −750 and −1500 mVSCE (Figure 8b,f,j). Dimples
were observed at the crack initiation site at −600 mVSCE , cleavage occurred at the crack
initiation site at −750 mVSCE , and transgranular fracturing was noted at −1500 mVSCE
(Figure 8c,g,k). All the specimens exhibited dimpling at the center of their fracture surfaces
(Figure 8d,h,l). Steel A only exhibited ductile fracturing at −600 mVSCE , and the lower Eapp
values resulted in more brittle fracture behavior. Even under a potential of −1500 mVSCE ,
the center of the specimen exhibited ductile fracture behavior. Therefore, hydrogen did not
diffuse into the center of the specimen.

Figure 8. Fractography of Steel A at (a–d) −600 mVSCE, (e–h) −750 mVSCE , and (i–l) −1500 mVSCE .
(a,e,i) Entire sample, (b,f,j) side view, (c,g,k) crack initiation site, and (d,h,l) center. Red arrows
indicate the initiation of cracks and direction of propagation.

44
Materials 2022, 15, 3406

Figure 9. Fractography of Steel B at (a–d) −600 mVSCE , (e–h) −750 mVSCE , and (i–l) −1500 mVSCE .
(a,e,i) Entire sample, (b,f,j) side view, (c,g,k) crack initiation site, and (d,h,l) center. Red arrows
indicate the initiation of cracks and direction of propagation.

In Steel B, uniform corrosion and cracks occurred on the side of the specimen at
−600 mVSCE . Cracks without any corrosion were observed at −750 and −1500 mVSCE
(Figure 9b,f,j), and the lower Eapp values were correlated with a higher density of cracks.
Cleavage was observed at the crack initiation site at −600 mVSCE , while mixed intergranular
and transgranular fractures were seen at the crack initiation sites of −750 and −1500 mVSCE
(Figure 9c,g,k). The intergranular fracture was more obvious at −1500 mVSCE , and in all
the specimens, dimples occurred at the center of the fracture surfaces (Figure 9d,h,l). Steel
B exhibited brittle fractures at −600 mVSCE , and the lower Eapp values were associated
with more obvious brittle fracture behaviors. At −1500 mVSCE , the center of Steel B
demonstrated ductile fracture behavior. Thus, like Steel A, hydrogen did not diffuse into
the center of the specimen. Under the same Eapp , Steel B exhibited more brittle fracture
behavior than Steel A. The fractography results confirmed that, compared with Steel A,
Steel B was more susceptible to delayed fractures.

3.4. Hydrogen Trapping and Desorption Behaviors


To investigate the desorption behavior of diffusible hydrogen, the SSRT specimens
were analyzed by TDS at the Eapp values of −600, −750, and −1500 mVSCE . The results
are presented in Figure 10 and Table 5. To quantitatively analyze the desorbed hydrogen,
the area below the desorption rate vs. the temperature curve was integrated [19] (see
Table 5 for the results). Just 0.05 ppm of diffusible hydrogen was released in the as-received
specimens. Most of the diffusible hydrogen that accumulated during the steel manufac-
turing process (e.g., during acid cleaning) appeared to be released during machining and
storage. However, when the potential was applied, the lower Eapp was associated with the
higher hydrogen desorption rate. Under the same Eapp values of both AHSSs, the desorbed
diffusible hydrogen content of Steel B was higher than in Steel A, except for −1500 mVSCE .
In that case, Peak 3 of Steel A and Peak 2 of Steel B (located at approximately 220 ◦ C)
originated from the deformation field around the dislocation. In this study, as tensile defor-
mation was considered an error, the hydrogen de-trapping from these peaks was negligible.

45
Materials 2022, 15, 3406

Theoretically, the production of hydrogen did not occur at −600 mVSCE , although diffusible
hydrogen was detected. It is assumed that the hydrogen was accumulated from the 1-h
stabilizing process before the SSRTs were conducted.

Figure 10. Cont.

46
Materials 2022, 15, 3406

Figure 10. Hydrogen desorption rates obtained by thermal desorption spectroscopy at a heating
rate of 4 ◦ C/min in a fractured specimens of (a–c) Steel A and (d–f) Steel B at (a,d) −600, (b,e) −750,
(c,f) −1500 mVSCE , and (g) the as-received condition.

Table 5. Desorbed hydrogen contents for each peak.

Applied Potential Peak 1 Peak 2 Peak 3 Sum of Peaks


Steel
(mVSCE ) (ppm) (ppm) (ppm) (ppm)
As-received 0.0569 - - 0.0569 ± 0.0323
−600 0.0622 0.1511 0.0311 0.2485 ± 0.1262
Steel A
−750 0.0579 0.3693 0.0514 0.4787 ± 0.0145
−1500 0.2264 0.7735 0.0978 1.0977 ± 0.0968
As-received 0.0506 - - 0.0506 ± 0.0268
−600 0.1388 0.3122 0.0057 0.4568 ± 0.2070
Steel B
−750 0.1748 0.4445 - 0.6193 ± 0.1280
−1500 0.9622 0.0246 - 0.9868 ± 0.0052

To analyze the hydrogen-trapping sites in the steel specimens, the activation energy for
hydrogen de-trapping was calculated using Equation (4), as proposed by Kissinger [20–22]:
 
∂ ln ϕ/Tc2 E
= − aT (4)
∂(1/Tc ) R

where Tc is the temperature (K) at which the hydrogen desorption rate is maximal, ϕ is the
heating rate (K/min), EaT is the activation energy for hydrogen de-trapping (kJ/mol), and
R is the ideal gas constant (8.314 J/K·mol).
As is shown in Figure 10, the desorption curves were deconvoluted into two or three
peaks of Gaussian curves, indicating that diffusible hydrogen accumulated at more than
two or three trapping sites. According to the Kissinger equation, the slope of ln(ϕ/Tc 2 ) vs.
1/Tc curve for each peak represents the activation energies (see Figure 11 for the results).
The activation energies for Steels A and B corresponding to each peak are illustrated in
Table 6.

47
Materials 2022, 15, 3406

Figure 11. ln(ϕ/Tc 2 ) vs. 1/Tc curve for (a–c) Steel A and (d–f) Steel B at (a,d) −600, (b,e) −750, and
(c,f) −1500 mVSCE .

48
Materials 2022, 15, 3406

Table 6. Calculated activation energies for hydrogen de-trapping.

Applied Potential Peak 1 Peak 2 Peak 3


Steel
(mVSCE ) (kJ/mol) (kJ/mol) (kJ/mol)
−600 21.5 20.5 25.6
Steel A −750 24.8 20.7 21.7
−1500 21.1 27.9 22.3
−600 27.6 26.1 28.1
Steel B −750 17.8 23.4 -
−1500 28.6 32.4 -

Table 7 summarizes the activation energies for hydrogen de-trapping reported in pre-
vious related studies. Based on the published literature, the electrochemically accumulated
hydrogen corresponding to Peaks 1 and 2 in Steel A was associated with the grain bound-
ary and dislocation. Peak 3 was associated with the mechanical deformation by tensile
deformation that occurred during the SSRTs [23]. For Steel B, the hydrogen corresponding
to Peaks 1 and 2 at −600 and −750 mVSCE , respectively, was desorbed from the grain
boundary, dislocation, and ferrite–Fe3 C interface. In that specimen, the contributions from
the grain boundary and dislocation were indistinguishable in Peak 1 at −1500 mV, which
means that Peak 1 was the sum of the hydrogen desorbed from the grain boundary and
dislocation. Peak 3 (−1500 mVSCE ) was associated with mechanical deformation by tensile
deformation, which occurred during the SSRTs.

Table 7. Types of reversible and irreversible hydrogen-trapping sites reported in the literature.

Type of Trap Activation Energy (kJ/mol) References


Reversible hydrogen-trapping sites
Ferrite/Fe3 C 10.9 [16]
Grain boundary 17.2 [16]
Ferrite/Fe3 C interface 18.4 [16,24]
Grain boundary, Dislocation 21–29 [25–28]
Deformation field around
29 ± 5 [23]
dislocation
Irreversible hydrogen-trapping sites
Semi-coherent TiC 49.9 [28]
High-angle grain boundary 53–59 [29]
NbC interface 63–68 [30]
Incoherent TiC 85.7, 86.9 [28]

3.5. Analysis of Defects Acting as Hydrogen Trapping Sites


3.5.1. Electron Backscattered Diffraction Analysis
EBSD analysis was conducted to measure the grain boundary density and fraction of
retained austenite; the results are shown in Figure 12 and Table 8. Each value was measured
three times to derive the mean value. The average grain sizes measured by EBSD for Steel
A and Steel B were 2.79 and 4.03 μm, respectively. Mild steel has an average approximate
grain size of 22 μm [31]; therefore, these values indicate that AHSSs have a smaller grain
size than mild steel.

Table 8. EBSD analysis results (relative value).

High-Angle Low-Angle Retained


Average Grain
Specimen Grain Boundary Grain Boundary Austenite
Size
Length Length Fraction
Steel A 2.79 μm 15.73 mm 1.91 mm 9.5%
Steel B 4.03 μm 13.93 mm 1.99 mm 10.9%

49
Materials 2022, 15, 3406

Figure 12. Electron backscattered diffraction results of (a–c) Steel A and (d–f) Steel B. (a,d) Grain
mapping. (b,e) Low-angle grain boundary (red) and high-angle grain boundary (green) mapping.
(c,f) Face-centered cubic (red) and body-centered cubic (green) mapping.
When the misorientation of a grain boundary exceeds 15◦ , it is termed a high-angle
grain boundary; otherwise, it is a low-angle grain boundary. In this study, Steel A had
the longer high-angle grain length than Steel B, and the low-angle grain boundary lengths
in both specimens were similar. The low-angle grain boundary is a reversible hydrogen-
trapping site, suggesting that a longer low-angle grain boundary is more likely to induce
HE [32]. Since the high-angle grain boundary is an irreversible hydrogen-trapping site, the
longer high-angle grain boundary enhances the HE resistance. Therefore, in both AHSSs,
the diffusible hydrogen content charged in the grain boundary is almost identical, and the
non-diffusible hydrogen content charged in the grain boundary of Steel A is expected to
be high.
The conducted EBSD analysis reveals that the face-centered cubic (FCC) structure
reflected the retained austenite content. In Steel B, the retained austenite fraction was 10.9%,
which was 1.4% higher than in Steel A. Retained austenite is an irreversible hydrogen-
trapping site that enhances the HE resistance. However, in Steel B, the retained austenite
fraction is not proportional to the HE resistance; this is because retained austenite with
an FCC structure is transformed by tensile stress into martensite with a body-centered
tetragonal (BCT) structure. Since BCT structures have a lower hydrogen solubility and
faster diffusion rate than FCC structures, hydrogen accumulation via diffusion is easy in
the BCT structure [33]. Thus, the hydrogen charged on the retained austenite in Steel B
segregates during the tensile process and becomes susceptible to HE. Furthermore, the
austenite–matrix interface is an effective diffusible hydrogen-trapping site [34]. The higher
fraction of retained austenite increases the susceptibility to HE, i.e., it is expected that Steel
B will be more susceptible to HE than Steel A.

3.5.2. X-ray Diffraction


The dislocation density of the samples was determined using XRD (see Figure 13 for
the results). Both AHSSs were mainly comprised of α-Fe, although γ-Fe peaks were also
observed. Specifically, the γ-Fe peaks were higher in intensity in Steel B compared with
Steel A, which is consistent with the results of the EBSD analysis. The dislocation density is

50
Materials 2022, 15, 3406

defined as the length of dislocation lines per unit volume of crystal and can be calculated
using the Williamson–Smallman relationship [35], as in Equation (5) below:

1
δ= (5)
D2

Figure 13. X-ray diffraction results for (a) Steel A and (b) Steel B.

Here, δ is the dislocation density, and D is the size of crystalline domain, which is
similar to the grain size. Therefore, D can be calculated using Scherrer’s equation [36],
as follows:

D= (6)
β cos θ
in which k is the shape factor (=approx. 0.9), λ is the wavelength (Cu-Kα = 1.5406 Å), β is
the full width at half-maximum (FWHM) value, and θ is the position of the peaks. Using
the above expression, the dislocation density was calculated to be 3.488 × 1014 /m2 and
6.263 × 1014 /m2 for Steel A and Steel B, respectively, i.e., the dislocation density of Steel
B was twice that of Steel A. Since the low-angle grain boundary areas of the two AHSSs
were similar, the difference in the hydrogen content of the two types of AHSSs discharged
during TDS was attributed to the difference in the dislocation density.

3.5.3. Characterization of Precipitates


To characterize the type and size of the precipitates, EPMA and TEM analyses were
conducted. The results are presented in Figure 14. According to Figure 14a, the precipitate
of Steel A was rich in Ti and Nb and a (Nb, Ti) precipitate surrounded the Ti-rich precipitate.
However, in the precipitates of Steel B, only Ti was detected, while Nb was undetected.
The precipitates of both AHSSs were approximately 1 μm in size. The TEM images and
diffraction patterns for the two types of AHSSs are presented in Figure 14c,d. In Steel
A, extremely fine precipitates were distributed along the grain boundary. The electron
diffraction pattern and energy-dispersive X-ray spectroscopy analysis confirmed that the
precipitates were amorphous Ti and Fe carbides smaller than 10 nm in size. Only the
small fraction of Fe carbides was distributed randomly in the grain, and in Steel B, no
TiC precipitate was observed (Figure 14d). The EPMA and TEM results revealed the
presence of sub-micrometer (Nb, Ti)C and fine TiC precipitates in Steel A, although Steel B
contained only a sub-micrometer TiC precipitate. The small size of the carbide produced
a large effective area for hydrogen trapping [37,38]. Therefore, Steel A was able to trap
considerably more hydrogen in the TiC precipitate interface compared with Steel B. Since
Nb and Ti precipitates are powerful and irreversible hydrogen-trapping sites, they can
positively influence HE resistance, i.e., Steel A is expected to be more resistant to HE than
Steel B.

51
Materials 2022, 15, 3406

Figure 14. Identification of the types and sizes of precipitates on each AHSSs. Electron probe
microanalysis results for (a) Steel A and (b) Steel B. Transmission electron microscopy results for
(c) Steel A and (d) Steel B.

52
Materials 2022, 15, 3406

4. Conclusions
This study investigated the SCC and HE mechanisms of two AHSSs using SSRTs and
characterized their hydrogen-trapping behaviors using TDS, EBSD, and XRD. According to
the results of these investigations, the SCC and HE characteristics of the studied AHSSs
can be summarized as follows:
1. For both AHSSs, elongation decreased as the cathodic overvoltage increased, i.e.,
both types of AHSSs were fractured by the mechanism of HE. Even when the anodic
potential was applied, HE was more dominant than SCC. Although the HE sensitivity
of Steel B was higher than that of Steel A, both AHSSs were more sensitive to HE than
SCC.
2. In both AHSSs, the lower Eapp was associated with a strong brittle fracture behavior.
However, the center of each specimen exhibited ductile fracture behavior, because the
hydrogen did not diffuse into that region. It was clear that the fracture surface of Steel
B was more brittle than that of Steel A.
3. The lower Eapp was associated with the higher rate of hydrogen desorption. In both
AHSSs, diffusible hydrogen was trapped mainly at the grain boundary and dislocation.
4. The density of the irreversible hydrogen-trapping sites (high-angle grain boundaries
and TiC precipitates) was higher in Steel A than in Steel B. However, the density of
the reversible hydrogen-trapping sites (low-angle grain boundaries and dislocations)
was lower in Steel A than in Steel B. Therefore, compared to Steel A, Steel B was more
susceptible to HE.

Author Contributions: Conceptualization, S.C. and G.-I.K.; methodology, S.C. and G.-I.K.; validation,
S.C., G.-I.K., S.-J.K., J.-S.Y., Y.-S.J., Y.-H.Y. and J.-G.K.; formal analysis, S.C. and G.-I.K.; investigation,
S.C. and G.-I.K.; resources, S.C. and G.-I.K.; data curation, S.C., G.-I.K., S.-J.K. and J.-S.Y.; writing—
original draft preparation, S.C. and G.-I.K.; writing—review and editing, S.C., G.-I.K., S.-J.K., J.-S.Y.,
Y.-S.J., Y.-H.Y. and J.-G.K.; visualization, S.C. and G.-I.K.; supervision, J.-G.K.; project administration,
J.-G.K.; and funding acquisition, J.-G.K. All authors have read and agreed to the published version of
the manuscript.
Funding: This research was funded by the POSCO (Grant Number 2019Z083).
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Not applicable.
Acknowledgments: This research was supported by SungKyunKwan University and BK21 FOUR
(Graduate School Innovation) funded by the Ministry of Education (MOE, Korea) and the National
Research Foundation of Korea (NRF).
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Miller, W.; Zhuang, L.; Bottema, J.; Wittebrood, A.J.; De Smet, P.; Haszler, A.; Vieregge, A. Recent development in aluminium
alloys for the automotive industry. Mater. Sci. Eng. A 2000, 280, 37–49. [CrossRef]
2. Kulekci, M.K. Magnesium and its alloys applications in automotive industry. Int. J. Adv. Manuf. Technol. 2008, 39, 851–865.
[CrossRef]
3. Cole, G.; Sherman, A. Light weight materials for automotive applications. Mater. Charact. 1995, 35, 3–9. [CrossRef]
4. Friedrich, H.; Schumann, S. Research for a “new age of magnesium” in the automotive industry. J. Mater. Process. Technol. 2001,
117, 276–281. [CrossRef]
5. Musfirah, A.; Jaharah, A. Magnesium and aluminum alloys in automotive industry. J. Appl. Sci. Res 2012, 8, 4865–4875.
6. Gándara, M.J.F. Recent growing demand for magnesium in the automotive industry. Mater. Technol. 2011, 45, 633–637.
7. Tisza, M.; Czinege, I. Comparative study of the application of steels and aluminium in lightweight production of automotive
parts. Int. J. Lightweight Mater. Manuf. 2018, 1, 229–238. [CrossRef]
8. Hirsch, J. Recent development in aluminium for automotive applications. Trans. Nonferrous Met. Soc. China 2014, 24, 1995–2002.
[CrossRef]

53
Materials 2022, 15, 3406

9. Baluch, N.; Udin, Z.M.; Abdullah, C.S. Advanced high strength steel in auto industry: An overview. Eng. Technol. Appl. Sci. Res.
2014, 4, 686–689. [CrossRef]
10. Galán, J.; Samek, L.; Verleysen, P.; Verbeken, K.; Houbaert, Y. Advanced high strength steels for automotive industry. Rev. Metal.
2012, 48, 118. [CrossRef]
11. Funakawa, Y.; Nagataki, Y. High Strength Steel Sheets for Weight Reduction of Automotives. JFE Tech. Rep. 2019, 24, 1–5.
12. Lovicu, G.; Bottazzi, M.; D’aiuto, F.; De Sanctis, M.; Dimatteo, A.; Santus, C.; Valentini, R. Hydrogen embrittlement of automotive
advanced high-strength steels. Metall. Mater. Trans. A 2012, 43, 4075–4087. [CrossRef]
13. Khanzhin, V.; Nikulin, S.; Belov, V.; Rogachev, S.; Turilina, V.Y. Hydrogen embrittlement of steels: III. Influence of secondary-phase
particles. Russ. Metall. (Met.) 2013, 2013, 790–796. [CrossRef]
14. Khanzhin, V.; Turilina, V.Y.; Rogachev, S.; Nikitin, A.; Belov, V. Effect of various factors on hydrogen embrittlement of structural
steels. Met. Sci. Heat Treat. 2015, 57, 197–204. [CrossRef]
15. Hagi, H.; Hayashi, Y.; Ohtani, N. Diffusion coefficient of hydrogen in pure iron between 230 and 300 K. Trans. Jpn. Inst. Met. 1979,
20, 349–357. [CrossRef]
16. Choo, W.; Lee, J.Y. Thermal analysis of trapped hydrogen in pure iron. Metall. Mater. Trans. A 1982, 13, 135–140. [CrossRef]
17. Katiyar, P.K.; Misra, S.; Mondal, K. Comparative corrosion behavior of five microstructures (pearlite, bainite, spheroidized,
martensite, and tempered martensite) made from a high carbon steel. Metall. Mater. Trans. A 2019, 50, 1489–1501. [CrossRef]
18. Jones, D.A. Principles and Prevention of Corrosion, 2nd ed.; Prentice Hall: Upper Saddle River, NJ, USA, 1996.
19. Yoo, J.; Xian, G.; Lee, M.; Kim, Y.; Kang, N. Hydrogen embrittlement resistance and diffusible hydrogen desorption behavior of
multipass FCA weld metals. J. Weld. Join. 2014, 31, 112–118. [CrossRef]
20. Hurley, C.; Martin, F.; Marchetti, L.; Chêne, J.; Blanc, C.; Andrieu, E. Numerical modeling of thermal desorption mass spectroscopy
(TDS) for the study of hydrogen diffusion and trapping interactions in metals. Int. J. Hydrog. Energy 2015, 40, 3402–3414. [CrossRef]
21. Kirchheim, R. Bulk diffusion-controlled thermal desorption spectroscopy with examples for hydrogen in iron. Metall. Mater.
Trans. A 2016, 47, 672–696. [CrossRef]
22. Ryu, J.H.; Chun, Y.S.; Lee, C.S.; Bhadeshia, H.; Suh, D.W. Effect of deformation on hydrogen trapping and effusion in TRIP-assisted
steel. Acta Mater. 2012, 60, 4085–4092. [CrossRef]
23. Krieger, W.; Merzlikin, S.V.; Bashir, A.; Szczepaniak, A.; Springer, H.; Rohwerder, M. Spatially resolved localization and
characterization of trapped hydrogen in zero to three dimensional defects inside ferritic steel. Acta Mater. 2018, 144, 235–244.
[CrossRef]
24. Hong, G.-W.; Lee, J.-Y. The interaction of hydrogen and the cementite-ferrite interface in carbon steel. J. Mater. Sci. 1983, 18,
271–277. [CrossRef]
25. Wang, M.; Akiyama, E.; Tsuzaki, K. Effect of hydrogen and stress concentration on the notch tensile strength of AISI 4135 steel.
Mater. Sci. Eng. A 2005, 398, 37–46. [CrossRef]
26. Hong, G.-W.; Lee, J.-Y. The interaction of hydrogen with dislocation in iron. In Perspectives in Hydrogen in Metals; Elsevier:
Amsterdam, The Netherlands, 1986; pp. 427–435.
27. Yamasaki, S.; Takahashi, T. Evaluation method of delayed fracture property of high strength steels. J. Iron Steel Inst. Jpn. 1997, 83,
454–459. [CrossRef]
28. Wei, F.; Hara, T.; Tsuzaki, K. Precise determination of the activation energy for desorption of hydrogen in two Ti-added steels by a
single thermal-desorption spectrum. Metall. Mater. Trans. B 2004, 35, 587–597. [CrossRef]
29. Pressouyre, G. A classification of hydrogen traps in steel. Metall. Mater. Trans. A 1979, 10, 1571–1573. [CrossRef]
30. Wallaert, E.; Depover, T.; Arafin, M.; Verbeken, K. Thermal desorption spectroscopy evaluation of the hydrogen-trapping capacity
of NbC and NbN precipitates. Metall. Mater. Trans. A 2014, 45, 2412–2420. [CrossRef]
31. Jangir, D.K.; Verma, A.; Sankar, K.M.; Khanna, A.; Singla, A. Influence of grain size on corrosion resistance and electrochemical
behaviour of mild steel. Int. J. Res. Appl. Sci. Eng. Technol. 2018, 6, 2875–2881. [CrossRef]
32. Oger, L.; Malard, B.; Odemer, G.; Peguet, L.; Blanc, C. Influence of dislocations on hydrogen diffusion and trapping in an
Al-Zn-Mg aluminium alloy. Mater. Des. 2019, 180, 107901. [CrossRef]
33. Kang, H.-J.; Kang, N.-H.; Park, S.-J.; Chang, W.-S. Evaluation of hydrogen delayed fracture for high strength steels and weldments.
J. Weld. Join. 2011, 29, 33–39. [CrossRef]
34. Szost, B.A.; Vegter, R.H.; Rivera-Díaz-del-Castillo, P.E. Hydrogen-trapping mechanisms in nanostructured steels. Metall. Mater.
Trans. A 2013, 44, 4542–4550. [CrossRef]
35. Shahmoradi, Y.; Souri, D.; Khorshidi, M. Glass-ceramic nanoparticles in the Ag2O–TeO2–V2O5 system: Antibacterial and
bactericidal potential, their structural and extended XRD analysis by using Williamson–Smallman approach. Ceram. Int. 2019, 45,
6459–6466. [CrossRef]
36. Bykkam, S.; Ahmadipour, M.; Narisngam, S.; Kalagadda, V.R.; Chidurala, S.C. Extensive studies on X-ray diffraction of green
synthesized silver nanoparticles. Adv. Nanopart 2015, 4, 1–10. [CrossRef]
37. Yamasaki, S.; Manabe, T.; Hirakami, D. Analysis of Hydrogen State in Steel and Trapping Using Thermal Desorption Method. Nippon
Steel & Sumitomo Metal Technical Report. 2017, pp. 38–43. Available online: https://www.nipponsteel.com/en/tech/report/
nssmc/pdf/116-08.pdf (accessed on 18 January 2022).
38. Turk, A.; San Martín, D.; Rivera-Díaz-del-Castillo, P.E.; Galindo-Nava, E.I. Correlation between vanadium carbide size and
hydrogen trapping in ferritic steel. Scr. Mater. 2018, 152, 112–116. [CrossRef]

54
materials
Article
Effect of Strain Rate and Temperature on Tensile and Fracture
Performance of AA2050-T84 Alloy
Nagaraj Ekabote 1 , Krishnaraja G. Kodancha 1, *, T. M. Yunus Khan 2,3 and Irfan Anjum Badruddin 3

1 School of Mechanical Engineering, KLE Technological University, Hubballi 580031, India;


[email protected]
2 Research Center for Advanced Materials Science (RCAMS), King Khalid University, 9004,
Abha 61413, Saudi Arabia; [email protected]
3 Department of Mechanical Engineering, College of Engineering, King Khalid University, 394,
Abha 61421, Saudi Arabia; [email protected]
* Correspondence: [email protected]; Tel.: +91-98-8659-6953

Abstract: AA2050-T84 alloy is widely used in primary structures of modern transport aircraft.
AA2050-T84 is established as a low-density aluminum alloy with improved Young’s modulus, less
anisotropy, and temperature-dependent mechanical properties. During flights, loading rate and
temperature variation in aircraft engine subsequent parts are commonly observed. The present work
focuses on the effect of loading rate and temperature on tensile and fracture properties of the 50 mm
thick (2-inch) AA2050-T84 alloy plate. Quasi-static strain rates of 0.01, 0.1, and 1 s−1 at −20 ◦ C,
24 ◦ C and 200 ◦ C are considered. Tensile test results revealed the sensitivity of mechanical properties
towards strain rate variations for considered temperatures. The key tensile properties, yield, and
ultimate tensile stresses were positive strain rate dependent. However, Young’s modulus and
elongation showed negative strain rate dependency. Experimental fracture toughness tests exhibited
the lower Plane Strain Fracture Toughness (KIC ) at −20 ◦ C compared to 24 ◦ C. Elastic numerical
Citation: Ekabote, N.; Kodancha, fracture analysis revealed that the crack driving and constraint parameters are positive strain rate
K.G.; Khan, T.M.Y.; Badruddin, I.A. dependent and maximum at −20 ◦ C, if plotted and analyzed over the stress ratio. The current results
Effect of Strain Rate and Temperature concerning strain rates and temperatures will help in understanding the performance-related issues
on Tensile and Fracture Performance of AA2050-T84 alloy reported in aircraft applications.
of AA2050-T84 Alloy. Materials 2022,
15, 1590. https://doi.org/10.3390/ Keywords: AA2050-T84 alloy; strain rate effect; plane strain fracture toughness; temperature effect;
ma15041590 strain rate effect; constraint effect
Academic Editors: Jaroslaw
Galkiewicz and Lucjan Śnieżek

Received: 26 January 2022 1. Introduction


Accepted: 18 February 2022
Modern aircraft predominantly use lightweight structures to improve the performance-
Published: 20 February 2022
to-weight ratio. The low-density Aluminum alloy is popular among aircraft structures
Publisher’s Note: MDPI stays neutral owing to its durable mechanical properties and ease of manufacturability [1,2]. Composites
with regard to jurisdictional claims in pose tough competition to Aluminum alloys due to their tailor-made properties suited for
published maps and institutional affil- specific applications. The unpredictable behavior of composites for change in temperature
iations. and time, restricted its usage to secondary and tertiary aircraft structures [3]. Currently,
the modern transport aircraft primary structures are built by Al-Li alloys. The Lithium
addition to aluminum with improved manufacturing methods resulted in the enhancement
of specific strength and stiffness of the alloy [4]. However, the higher cost of Al-Li alloy
Copyright: © 2022 by the authors.
restricted its usage to only aerospace industries.
Licensee MDPI, Basel, Switzerland.
This article is an open access article
The apprehensions related to diversifying mechanical properties of 1st and 2nd gen-
distributed under the terms and
eration Al-Li alloys directed the complete withdrawal from aerospace applications [2].
conditions of the Creative Commons
Some noteworthy limitations were anisotropic behavior, cracking during manufacturing,
Attribution (CC BY) license (https:// and thermal instability-driven lower fracture toughness [2]. The skillful and sophisticated
creativecommons.org/licenses/by/ fabrication methods steered the evolution of 3rd generation Al-Li alloys. The spars and ribs
4.0/). of modern transport aircraft are fabricated by a 3rd generation Al-Li alloy, AA2050-T84 [1].

Materials 2022, 15, 1590. https://doi.org/10.3390/ma15041590 55 https://www.mdpi.com/journal/materials


Materials 2022, 15, 1590

AA2050-T84 alloy exhibits exceptional tensile, fatigue, and fracture toughness behavior
suited for damage tolerance property requirements of wing structures [1]. Notably, the
anisotropic behavior and temperature-dependent property variations of AA2050-T84 alloy
were also reported [5,6]. During flights, the fuel pressure at various altitudes of an aircraft,
wing lift, and drag loads at different operating conditions may result in load rate and
temperature variations on spars and ribs. The sensitivity of AA2050-T84 alloy to these load
and temperature variations are essential to claim its suitability to modern transport aircraft
wing parts. The following paragraphs discuss the strain rate and temperature effect on
various ductile and brittle material properties reported in the literature.
Mirza et al. [7] have conducted tensile tests on mild steel and aluminum at various
quasi-static (lower strain rates up to 1 s−1 ) strain rates. The results have shown negligible
dependency of ductility on strain rate variations. Clausen et al. [8] have reported the
negative strain rate sensitivity of AA5083–H116 through tensile tests for quasi-static strain
rate variations. The strain rate dependency was related to dynamic strain aging at lower
strain rates and temperatures, resulting in serrated stress-strain curves. Singh et al. [9] have
reported increased flow stress with the rise in test temperature and strain rates on titanium
alloys. The observed trend was attributed to the dynamic strain aging of the alloy. Through
experiments and numerical analysis, Khan et al. [10] have investigated the influence of
strain rate and temperature on Al2024-T351. The results inferred a strong temperature
dependency and negligible strain rate effect on fracture strength of the Al2024-T351 alloy.
Anderson et al. [11] have witnessed the sensitivity of DP 780 steel towards the strain rate
variations. The tensile stress-strain response was steady and almost negligible concerning
quasi-static strain rates. The alterations in failure surface morphology were noticed with
changes in strain rates.
The experimental tensile results on DP590 and TRIP 780 steel by Roth et al. [12] have
shown that ductility increases with loading speed. Rincon et al. [13] have studied the
influence of temperature (between −90 ◦ C to 270 ◦ C) on tensile behavior of an as-cast A319
alloy and noticed the silicon dominant brittle fracturing regardless of temperature variation.
Natesan et al. [14] have reported the variation in strain rate effect at different temperatures
on the deformation behavior of A356-T7 cast aluminum alloys. The yield stress and strain
hardening of Aluminum alloy 7075-W exhibited the positive load rate effect and negative
temperature effect through plasticity experiments [15]. Hafley et al. [5] and Chemin et al. [6]
have reported AA2050-T84 alloy tensile and fracture properties sensitivity to temperature
variations. In summary, the material properties of various alloys of steel and aluminum
generally exhibit reliance on strain rate and temperature and are noteworthy.
The dependency of fracture behavior on strain rate and temperature mainly alters the
state of stress near the crack front. The state of stress variation at the crack front due to
specimen type, geometry, and load type was defined by a term constraint. The constraint
level at the crack component/structure drives the selection of standard test specimens for
fracture toughness tests [16,17]. In Linear Elastic Fracture Mechanics (LEFM), the variation
of the state of stress near the crack was measured by popular constraint parameters [18–26].
T11 and T33 are used to measure the in-plane and out-of-plane constraints in LEFM. The
variations of T11 and T33 concerning specimen type, geometry, and load were well docu-
mented [16,18–26]. However, the strain rate effect at different temperatures on tensile and
fracture behavior of the AA2050-T84 alloy is essential to claim its suitability to primary
structures of the aircraft wing. Fracture toughness standard methods recommend the
single value of fracture toughness for quasi-static load variations [27]. Notably, the above
literature study shows the strong dependency of material properties on strain rate and
temperature variations. Moreover, critical cracks were observed in AA2050-T84 made spars
and ribs of Airbus-380 aircraft after a few flights [28]. The present literature findings de-
mand a fracture study based on constraints near the crack of AA2050-T84 alloy at different
strain rates and temperatures.
In the present work, AA2050-T84 alloy tensile behavior at different strain rates and
temperatures are experimentally studied. Experimental fracture toughness tests are con-

56
Materials 2022, 15, 1590

ducted using Compact Tension (C(T)) specimens at various temperatures. Furthermore,


the effect of strain rate on fracture characterizing parameters in LEFM such as Stress
Intensity Factor (KI ), T11, and T33 are analyzed numerically for different temperatures.
Finally, the AA2050-T84 alloy tensile and fracture behavior dependency on strain rates and
temperatures are compared and evaluated for compatibility for aircraft wing structures.

2. Material and Test Details


2.1. AA2050-T84 Alloy
This study uses a 50 mm thick (2-inch) AA2050-T84 alloy plate to extract the test
specimens. The chemical composition of AA2050-T84 alloy in wt% as obtained from the
supplier is shown in Table 1. Copper is used in AA2050-T84 to provide high strength,
suited for aircraft applications [29]. The Lithium addition is (<1%) restricted to balance
between density reduction and increase in Young’s modulus of the alloy [3,29].

Table 1. Chemical composition of 50 mm thick (2-inch) AA2050-T84 alloy plate (wt%).

Cu Mg Mn Zn Fe Ti Si Li Zr Ag Al
3.743 0.369 0.372 0.025 0.045 0.040 0.039 0.798 0.087 0.398 Base

2.2. Tensile and Fracture Toughness Test


ASTM E8/E8M-21, the standard test method for tension testing of metallic materi-
als [30], was used for the tensile specimen preparation and testing of AA2050-T84 alloy.
Round specimens were extracted in the rolling (along the length of the plate) direction of
the AA2050-T84 plate. Figure 1 shows the tensile test specimen dimensions (in mm) used in
this study. The main dimensions of the specimen are, gauge diameter (D0 = 6 mm), gauge
length (L0 = 30 mm), and overall specimen length (L = 65 mm). The tensile specimens were
designed, keeping the L0 /D0 ratio to 5.

Figure 1. Tensile test specimen. All dimensions are in mm.

ASTM E399-20a, the standard test method for linear-elastic plane strain fracture
toughness of metallic materials [27], was used. The commonly used fracture test specimen
for primary aircraft structures is the compact tension (C(T)) specimen shown in Figure 2.
The standard dimensions (in mm) are specimen width (W = 25.4 mm), specimen height
(H = 2W), and specimen thickness (B) = crack length (a) = 0.5W. The C(T) specimen is
extracted, ensuring the crack length in the rolling direction and load application in the
transverse directions of the plate.
Flight durations and operations cause the temperature variations of the wing parts.
Furthermore, the effect of these variations depends on alloy type and its ductile to brit-
tle transition temperature [31]. However, temperature variations will be high near the
aircraft engine (wing components), and the experienced load rates are dynamic. In the
present study, the quasi-static strain rates considered in the tensile tests were 0.01, 0.1, and
1 s−1 . The temperatures considered were −20 ◦ C (Sub-zero temperature), 24 ◦ C (Room
temperature), and 200 ◦ C (High temperature) [32]. In the tensile and fracture toughness
tests, the low-temperature chamber with liquid nitrogen and a high-temperature furnace
with forced convection heating was used to maintain the sub-zero and high temperatures.

57
Materials 2022, 15, 1590

The Servo Electric Universal Testing Machine (UTM) (BISS, Bangalore, India) with 50 kN
capacity was used for tensile and fracture toughness tests. In tensile testing, the applied
load and deformations were recorded continuously through the load cell and extensometer,
respectively. However, along with these, Crack Opening Displacement (COD) gauge (BISS,
Bangalore, India) was used to record the relative displacement of two knife edges of the C(T)
specimen in the fracture toughness test. These data were further processed to extract the
tensile properties and fracture toughness of the AA2050-T84 alloy as per standards [27,30].

Figure 2. Compact Tension (C(T)) Specimen. All dimensions are in mm.

2.3. Finite Element Analysis


The crack driving forces and constraints of the C(T) specimen were investigated at
different load rates and temperatures using 3D linear elastic finite element analysis (FEA).
Half-symmetry is modeled and analyzed using Abaqus (6.14, 2014, Dassault Systemes Simu-
lia Corp., Providence, RI, USA). The Poisson’s ratio (v) and Young’s modulus (E) obtained
from experimentally conducted tensile results at different strain rates and temperatures
are used for linear elastic fracture analysis. The output parameters viz. stress intensity
factor (KI ) and constraint parameter (T11 ) were extracted using the counter-integral method
mentioned in Abaqus post-processor [33]. T33 is calculated by using Equation (1). In
Equation (1), ε33 is a strain in the z-direction (thickness direction) extracted along the crack
front. The material property input and the KI extraction details were adopted as similar to
the work of [22,33].
T33 = Eε33 + vT11 (1)
Half symmetry C(T) meshed model with supports and loading is shown in Figure 3.
20-noded hexahedral elements with reduced integration were used for the meshing. A
fine mesh near the crack front was used to encapsulate the crack characteristics effectively.
Singularity at the crack front was emulated by shifting the mid-side nodes of crack sur-
rounding elements towards the crack front. The crack edge (crack front) surrounded by
these nodes is defined as contour integral. The output parameters are calculated along
the user-defined contour integrals (in the present analysis, it is 10 contours). The detailed
procedure to define the crack front and contour integrals to obtain crack driving parameters
is available in the Abaqus manual [33]. Y-symmetry was imposed along the ligament (the
uncracked portion in the crack plane), and the tensile load was applied through the hole to
simulate Mode-I.

58
Materials 2022, 15, 1590

Figure 3. C(T) Meshed Model with Boundary conditions.

3. Results and Discussions


The following sections discuss the experimental and numerical analysis of the AA2050-
T84 alloy.

3.1. Experimental Analysis


3.1.1. Tensile Test Analysis
The experimental tensile tests were conducted at varying temperatures and strain rates.
A total of 27 tensile tests were conducted, comprising 3 experiments for each strain rate per
temperature. The stress-strain curves for different strain rates and temperatures are shown
in Figure 4a. The corresponding variation of average tensile properties viz. yield stress
(σys ), ultimate tensile stress (σut ), Young’s modulus (E), and % elongation (% et) extracted
from the stress-strain curves along with error bars are presented in Figure 4b–e. The
positive strain rate dependency was observed for tensile yield stress for the temperatures
considered in the study, as shown in Figure 4b. The highest yield stress variation of 2%
between 0.01 s−1 to 0.1 s−1 and 0.1 s−1 to 1 s−1 was observed at −20 ◦ C. However, the
lowest yield stress variation between successive strain rates, around 0.5%, was noticed at
24 ◦ C. It was observed that the yield stress decreased from −20 ◦ C to room temperature
and further increased slightly at 200 ◦ C, indicating the V-shaped behavior for temperature
variation.
Similarly, ultimate tensile stress exhibited the positive strain rate dependency at
various temperatures, as shown in Figure 4c. However, strain rate has minimal effect on
ultimate tensile stress as the difference observed between successive strain rates for all
temperatures is less than 1%. The ultimate tensile stress is inversely proportional to the
temperature for all the strain rates and is in line with the observations of Hafley et al. [5]
and Chemin et al. [6]
Figure 4d, shows the variation of Young’s modulus at various strain rates and temper-
atures. Young’s modulus showed negative strain rate sensitivity at room and higher tem-
peratures. However, the strain rate effect on Young’s modulus was negligible (around 1%)
at −20 ◦ C. The maximum Young’s modulus difference of about 10% was observed at 200 ◦ C
between strain rates 0.01 and 1 s−1 . Furthermore, Young’s modulus difference was around
5% at room temperature for successive strain rate variations. This reveals that the strain
rate sensitivity towards Young’s modulus was in the decreasing order of temperatures
200 ◦ C:24 ◦ C:−20 ◦ C. At 200 ◦ C. Young’s modulus values were minimal and almost similar
in values at −20 ◦ C.

59
Materials 2022, 15, 1590

700
AA 2050-T84 alloy
(a)
600

500
−20 °C @ SR 0.01 s-1

True stress (MPa)


−20 °C @ SR 0.1 s-1
400
−20 °C @ SR 1 s-1

300 24 °C @ SR 0.01 s-1


24 °C @ SR 0.1 s-1
200 24 °C @ SR 1 s-1

100
200 °C @ SR 0.01 s-1
200 °C @ SR 0.1 s-1
200 °C @ SR 1 s-1
0
0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14 0.16
True strain
560

550 @ −20 °C 640 @ −20 °C


@ 24 °C (b) @ 24 °C
(c)
540 @ 200 °C 620 @ 200 °C
Ultimate stress (MPa)
Yield stress (MPa)

530
600
520

510 580

500
560
490

480 540

470
520
0.01 0.1 1 0.01 0.1 1
Strain rate (s-1) Strain rate (s-1)

80 16
@ −20 °C @ −20 °C
@ 24 °C (d) @ 24 °C (e)
75
@ 200 °C @ 200 °C
Young's modulus (GPa)

14
Elongation (%)

70

12
65

10
60

55 8
0.01 0.1 1 0.01 0.1 1
-1
Strain rate (s ) Strain rate (s-1)

Figure 4. Strain rate and Temperature effect on: (a) True stress-strain curves (b) Yield stress; (c) Ulti-
mate stress; (d) Young’s modulus; (e) elongation.

Overall, a reduction between 6% and 10% is noticed in Young’s modulus for the
temperatures studied. The exact thickness of the plate with room temperature and −54 ◦ C
has been studied by Chemin et al. [6], revealing the same trend with a 2.5% reduction in
Young’s modulus. With a 100 mm (4-inch) plate, Hafley et al. [5] noticed a 9–11% reduction
in Young’s modulus when studied at different locations for room and −196 ◦ C temperature.
Strong interatomic bonding between the atoms at room temperature may be the
probable reason for the highest value of Young’s modulus. Farraro and McLellan [34]

60
Materials 2022, 15, 1590

have reported that the lower values of Young’s modulus at elevated temperature indicate
weakened interatomic bonding between the atoms.
Figure 4e shows the % elongation of AA2050-T84 alloy at different strain rates and
temperatures. Negative strain rate dependency on % elongation as similar to Young’s
modulus was observed for various temperatures. The maximum % elongation was noticed
at room temperature, indicating higher ductility than other temperatures.

3.1.2. Fracture Toughness Test Analysis


The fracture toughness tests were conducted as per ASTM E399-20a at different
temperatures. ASTM E399-20a essentially elucidates the procedure of obtaining the single
value, Plane Strain Fracture Toughness (KIC ), for metallic materials under quasi-static
strain rates. The C(T) specimen was fatigue pre-cracked to emulate the natural crack
characteristics. The effect of loading and material properties (mainly yield stress) strongly
influences fatigue crack growth [35]. The pre-cracking load details are shown in Table 2.
The load ratio (σmin /σmax ) was maintained at 0.1 to attain the a/W range in between
0.45 and 0.55.

Table 2. Pre-cracking details of fracture toughness test.

Crack Crack Maximum Minimum Alternating


Mean Stress
Length Length/Width Stress Stress Stress
σ mean (MPa)
a (mm) a/W σ max (MPa) σ min (MPa) σ average (MPa)
12.94 0.51 250 25 137.5 112.5

The pre-cracked C(T) specimen was tested under Mode-I (opening mode) loading
through tensile load application at the holes. A minimum of 3 successful fracture tough-
ness tests was conducted at each temperature at the strain rate 0.01 s−1 . The KIC (aver-
age of 3 test samples) obtained from experiments for −20 ◦ C and 24 ◦ C are 904.28 and
1059.36 MPa mm1/2 , respectively. The error bar for the KIC is depicted in Figure 5.

1100
50 mm thick AA 2050-T84 alloy
@ Strain rate 0.01 s-1
1050
KIC (MPa mm1/2)

1000

950

900

í20 24
Temperature (°C)

Figure 5. Plane Strain Fracture Toughness (KIC ) at various temperatures.

At 200 ◦ C, all the 3 test results were invalid as the ample crack deviation was observed
from the crack plane. However, the KIC has reduced by about 15% from room to sub-
zero temperature. Similarly, in the work of Chemin et al. [6], a reduction of the order
of magnitude 16% in KIC was noticed in the rolling direction of the AA2050-T84 alloy
plate from room to cryogenic (−56 ◦ C) temperature. The decrease of KIC at sub-zero
temperatures of AA2050-T84 alloy can be attributed to surface hardening inside the grain
and validated by the grain microstructure of the alloy [6,36]. In summary, the positive

61
Materials 2022, 15, 1590

temperature dependency was witnessed for KIC of AA2050-T84 alloy at strain rate 0.01 s−1 .
The economic limitations on conducting further fracture toughness tests at other strain
rates impelled us to adopt the numerical analysis.

3.2. Linear Elastic Fracture Analysis


ASTM E399-20a, to predict the KIC of metallic materials, recommended the single
toughness value for lower strain rate variations. However, the variation of tensile properties
of AA2050-T84 alloy at different lower strain rates and temperatures was substantial. The
elastic fracture analysis was carried out at varied strain rates and temperatures using the
Abaqus software. The current numerical procedure was adopted from Kudari et al. [22]
and Kavale et al. [37] The KI values extracted through-thickness direction of the crack at
24 ◦ C conditions, and experimental KIC are shown in Figure 6. The experimental KIC value
was emulated through numerical fracture analysis with less than 1% error, as observed
in Figure 6.
1200
Experimental KIC Applied load = 7039 N
1150

1100 a/W = 0.51, B/W = 0.5


KI (MPa mm1/2)

@ 24 °C, SR 0.01 s-1


1050

1000 Numerically obtained KI

950

Crack front center


900

-8 -6 -4 -2 0 2 4 6 8
Crack Front, z (mm)

Figure 6. Stress intensity factor (KI ) along the crack front at 24 ◦ C.

Similarly, at −20 ◦ C, the validation of the numerical procedure was executed with less
than 1% error. For all the numerical analysis, the center of the crack front was associated
with the largest value of the crack characterizing parameter. Thus, the KI values at the
center of the specimen are used for analysis in further discussions.

3.2.1. Effect of Strain Rate


In linear elastic fracture analysis, the experimental load associated with KIC of the alloy is
the applied load at respective temperatures. However, the KIC at 200 ◦ C was unavailable, and
hence for the numerical analysis, the assumed load applied up to KI = 1200 MPa mm1/2 . The
applied stress (σapplied ) was determined using the relationship mentioned in Equation (2) [27].
The extracted values of KI at the crack front center for various strain rates and temperatures
are plotted against the stress ratio (σapplied /σys ) as shown in Figure 7. Positive strain rate
dependency of the KI was observed for all temperatures considered in this study. At room
temperature, a steady increase of 0.6% in KI was witnessed with the rise in strain rate.
However, for strain rate 1 s−1 , the maximum of 1.88% increase in KI was noticed at −20 ◦ C.
The strain rate sensitivity on KI was found to be maximum at −20 ◦ C and minimum at 24 ◦ C,
as the same trend was noticed for tensile yield stress values. The results of the KI are in line
with the yield stress variations of the alloy for all strain rates at different temperatures. The
nominal variations of KI (within the stress ratio of KIC ) indicate the negligible dependency
of fracture characterizing parameters on strain rates for AA2050-T84 alloy. Moreover, ASTM

62
Materials 2022, 15, 1590

E399-20a recommendation to use the single value of KIC for quasi-static strain rates seems to
be justifying as the difference in numerical KI values was minimal.

Papplied  a 
KI = √ f (2)
B W W

1200
@ −200 C, 3D @ 24 °C, 3D
1000
a/W = 0.5, B/W = 0.5 a/W = 0.5, B/W = 0.5
1000

800
KI, (MPa mm1/2)

KI, (MPa mm1/2)


800

600
600

400
400
SR 0.01 s-1 SR 0.01 s-1
SR 0.1 s-1 SR 0.1 s-1
200 200
SR 1 s-1 SR 1 s-1

0.04 0.08 0.12 0.16 0.20 0.04 0.08 0.12 0.16 0.20
Stress Ratio Stress Ratio
(a) (b)
1200
@ 200 °C, 3D
a/W = 0.5, B/W = 0.5
1000
KI, (MPa mm1/2)

800

600

400
SR 0.01 s-1
SR 0.1 s-1
200
SR 1 s-1

0.04 0.08 0.12 0.16 0.20


Stress Ratio
(c)

Figure 7. KI at specimen thickness center obtained using FEA for different strain rates vs. Stress ratio
(a) @ −20 ◦ C; (b) 24 ◦ C; (c) 200 ◦ C.

Further, the effect of strain rate on crack tip/front constraints has been evaluated
through T11 and T33 . The values of T11 and T33 are found to vary along the thickness
similar to KI variation and maximum being at the center of the specimen. One can infer that
the crack-front constraint is high at the center than at the surface; the material may fail at the
center than on surface or shows instability at the center of the specimen. As the constraint
parameters do not have a unique value for the specimen thickness, maximum values at
the center of the specimen are considered for further analysis. T11 and T33 variations for
different strain rates within the purview of KIC (or stress ratio) of the alloy at respective
temperatures are plotted in Figure 8. The nature of variation was identical at all strain rates
for both constraint parameters. However, the increase in stress ratio resulted in positive
T11 and negative T33 values at all strain rates and temperatures. It is clear from Figure 8,
that the applied stress was directly proportional to T11 and inversely proportional to T33 .
The negative strain along the thickness resulted in the negative T33 [20]. This is in close
agreement with the findings of Kudari et al. [22] for IF steel C(T) specimen.

63
Materials 2022, 15, 1590

150
T11 Stress T33 Stress T11 Stress T33 Stress
100
SR 0.01 s-1 SR 0.01 s-1 SR 0.01 s-1 SR 0.01 s-1
100 SR 0.1 s-1 SR 0.1 s-1 SR 0.1 s-1 SR 0.1 s-1
SR 1 s-1 SR 1 s-1 SR 1 s-1 SR 1 s-1
50
T11 & T33 (MPa)

T11 & T33 (MPa)


50

0
0

-50 -50
a/W = 0.5, B/W = 0.5 a/W = 0.5, B/W = 0.5
@ 20 °C, 3D @ 24 °C, 3D
-100 -100
0.00 0.04 0.08 0.12 0.16 0.20 0.00 0.04 0.08 0.12 0.16 0.20
Stress ratio Stress ratio
(a) (b)
200
T11 Stress T33 Stress
SR 0.01 s-1 SR 0.01 s-1
SR 0.1 s-1 SR 0.1 s-1
100 SR 1 s-1 SR 1 s-1
T11 & T33 (MPa)

-100 a/W = 0.5, B/W = 0.5


@ 200 °C, 3D

0.00 0.04 0.08 0.12 0.16 0.20 0.24 0.28


Stress ratio
(c)

Figure 8. T11 and T33 at specimen thickness center obtained using FEA for different strain rates vs.
Stress ratio (a) @ −20 ◦ C; (b) 24 ◦ C; (c) 200 ◦ C.

At 24 ◦ C and 200 ◦ C, both constraint parameters were unaffected (very marginal


difference) by the strain rate variation, as observed in Figure 8b,c. However, T33 variations
between the strain rates at −20 ◦ C were relatively substantial. At −20 ◦ C, between the strain
rates, a difference of 2.4% was found for T33 . T33 variation depended on Poisson’s ratio
and Young’s modulus (material property) of the alloy at different strain rates. Eventually,
the crack driving and constraint parameters were less sensitive to strain rate variations.
Strain rate effect on T11 variation is negligible as in-plane constraint depends on
specimen type, geometry, and loading type only. Furthermore, the variation of hydrostatic
stress along the uncracked ligament is studied at different strain rates. It is observed that
no variations are found at different strain rates. The negligible variation of T11 can also be
attributed to the uniform state of stress at the crack front and minimal variation of yield
stress (or stress ratio as depicted in the graph) between the strain rates. However, the T33
variation is quite measurable at different strain rates for −20 ◦ C, as shown in Figure 8a,
owing to the variations in material property (both Young’s modulus and yield stress).
Positive T11 results in a lower plastic zone at the crack tip and influences the specimen’s
unstable crack growth. Similarly, negative T33 results in loss of constraint at the crack tip.
Since the crack front plasticity is restricted in LEFM regime, the state of stress may
be unaltered due to strain rate variations at identical temperatures. The current serrated
stress-strain curves may affect the plasticity ahead of the crack front and can be accounted
in Elastic-Plastic Fracture Mechanics (EPFM) regime.

64
Materials 2022, 15, 1590

3.2.2. Effect of Temperature


The variation of KI at different temperatures for quasi-static strain rates is plotted in
Figure 9. The KI variation is linearly increased with an increase in stress ratio as expected,
and the nature of variation was the same for all temperatures considered. At the peak stress
ratio, the difference between KI at 200 ◦ C and −20 ◦ C is 5.1%, 5.16%, and 7.31% at strain rates
0.01, 0.1, and 1 s−1 , respectively. However, the variation of KI was minimum (around 4%)
for the temperatures 24 ◦ C and 200 ◦ C at all strain rates. Notably, the temperature effect was
highest at strain rate 1 s−1 and in the decreasing order of 1: 0.1: 0.01 s−1 . The substantial
variation of KI at sub-zero temperatures indicates that for identical load conditions, the
AA2050-T84 alloy is more prone to fracture failure than the other two temperatures as it
possesses lower KIC.
1400 1400
@ SR 0.01 s-1, 3D @ SR 0.1 s-1, 3D
1200 a/W = B/W = 0.5 1200 a/W = B/W = 0.5

1000 1000
KI (MPa mm1/2)

KI (MPa mm1/2)
800 800

600 600

400 400
−20 °C −20 °C
24 °C 24 °C
200 200
200 °C 200 °C
0 0
0.00 0.05 0.10 0.15 0.20 0.25 0.00 0.05 0.10 0.15 0.20 0.25
Stress Ratio Stress Ratio
(a) (b)
1400

@ SR 1 s-1, 3D
1200 a/W = B/W = 0.5

1000
KI (MPa mm1/2)

800

600

400
−20 °C
24 °C
200
200 °C

0
0.00 0.05 0.10 0.15 0.20 0.25
Stress Ratio
(c)

Figure 9. KI at specimen thickness center obtained using FEA for different temperature vs. Stress
ratio (a) SR 0.01 s−1 ; (b) SR 0.1 s−1 ; (c) SR 1 s−1 .

Similarly, Figure 10 shows the variation of T11 and T33 at different temperatures
for quasi-static strain rates. The variation of T11 and T33 were almost identical at 24 ◦ C
and 200 ◦ C at quasi-static strain rates. At 200 ◦ C, the T11 was less subtle and owed
lower in-plane constraints than the −20 ◦ C, and 24 ◦ C. Chemin et al. [6] have related
dislocations gathered along the grain boundaries, led to stress concentrations under loading
and promoted the lower fracture toughness of the AA2050-T84 alloy at sub-zero (−56 ◦ C)
temperature. Similarly, in the current analysis, the sensitivity of fracture toughness and
in-plane constraint against the stress ratio is highest at −20 ◦ C. The sensitivity may be

65
Materials 2022, 15, 1590

associated to crack front stress concentrations at grain boundaries at −20 ◦ C and can
be accounted through constraint parameters as shown in the current numerical analysis.
However, the T11 and T33 variations seem identical for strain rates 0.01 and 1 s−1 through
Figure 10a,c at −20 ◦ C. Eventually, T11 and T33 in combination with KI is maximum at 1 s−1
compared to 0.01 s−1 . Thus, the highest constraint associated with numerically obtained KI
of the AA2050-T84 alloy at −20 ◦ C is 1 s−1 . This behavior also resulted in the lower KIC at
−20 ◦ C than 24 ◦ C.

80 T11 80 T11
−20 °C −20 °C
60 60
24 °C 24 °C
40 200 °C 40 200 °C
T11 & T33 (MPa)

T11 & T33 (MPa)


20 20

0 @ SR 0.01 s-1, 3D 0 @ SR 0.1 s-1, 3D


a/W = B/W = 0.5 a/W = B/W = 0.5
-20 -20

-40 -40
T33 T33
-60 -60
−20 °C −20 °C
-80 24 °C -80 24 °C
200 °C 200 °C
-100 -100

0.00 0.05 0.10 0.15 0.20 0.25 0.00 0.05 0.10 0.15 0.20 0.25
Stress ratio Stress ratio
(a) (b)
80 T11
−20 °C
60
24 °C
40 200 °C
T11 & T33 (MPa)

20

0 @ SR 1 s-1, 3D
a/W = B/W = 0.5
-20

-40
T33
-60
−20 °C
-80 24 °C
200 °C
-100

0.00 0.05 0.10 0.15 0.20 0.25


Stress ratio
(c)

Figure 10. T11 and T33 at specimen thickness center obtained using FEA for different temperature vs.
Stress ratio (a) SR 0.01 s−1 ; (b) SR 0.1 s−1 ; (c) SR 1 s−1 .

In summary, the major constraint loss was observed for temperature variation com-
pared to quasi-static strain rate variations. Moreover, at −20 ◦ C, AA2050-T84 alloy pos-
sesses lower KIC with the highest in-plane crack tip/front constraint compared to the other
two temperatures. This behavior of the alloy makes it vulnerable to fracture failure in
cryogenic (sub-zero) applications at a strain rate 1 s−1 .

4. Conclusions
In the present study, the tensile and fracture behavior of the 50 mm thick (2-inch)
AA2050-T84 plate was considered at various temperatures for quasi-static strain rates.
Tensile tests revealed the sensitivity of mechanical properties towards the strain rates and

66
Materials 2022, 15, 1590

temperatures. Positive strain rate dependency was observed for temperatures considered
on yield stress and ultimate tensile stress of the alloy. A maximum of 2% increase in yield
stress was noticed between strain rates at −20 ◦ C. The lowest strain rate sensitivity of
around 0.5% was witnessed at room temperature. Notably, the ultimate stress variation
between the strain rates for temperatures was less than 1%. However, Young’s modulus
and % elongation were negative strain rate dependent. The maximum decrease of Young’s
modulus up to 10% was noticed at 200 ◦ C. The minimum Young’s modulus variation was
witnessed at −20 ◦ C between the strain rates.
Temperature sensitivity towards tensile behavior of the AA2050-T84 alloy was noticed.
The maximum yield stress variation of 10–11% was witnessed between room temperature
and −20 ◦ C. Notably, the yield stress increase was only up to 3% between room temperature
and 200 ◦ C. Similarly, for ultimate stress, the variation was up to 7.5% between room
temperature and −20 ◦ C. However, the reduction of Young’s modulus up to 18% was
noticed between room temperature and 200 ◦ C. This implies that yield and ultimate stress
are quite substantial at −20 ◦ C compared to other temperatures, making the AA2050-T84
alloy vulnerable at sub-zero temperatures. Moreover, an increase in strain rate prompts the
decrease in % elongation, implying the brittle behavior of the alloy at higher strain rates.
The crack driving and constraint parameters are less sensitive to strain rate variations.
However, at −20 ◦ C, crack characterizing and constraint parameters to strain rate variations
were moderately considerable. The temperature effect is highest at strain rate 1 s−1 and in
the decreasing order of 1:0.1:0.01 s−1 .
Overall, the AA2050-T84 alloy tensile and fracture performance obtained through
experimental and numerical analyses exhibited the dependency on strain rates and tem-
peratures. Hence, these mechanical properties of the alloy strongly influence the dam-
age tolerance design of spars and wings of the aircraft. The authors believe that the
current results concerning strain rates and temperatures will help in understanding the
performance-related issues of AA2050-T84 alloy reported in aircraft applications.

Author Contributions: Conceptualization, N.E. and K.G.K.; methodology, N.E. and K.G.K.; software,
N.E.; validation, N.E. and K.G.K.; formal analysis, N.E.; investigation, N.E.; resources, N.E. and
K.G.K.; data curation, K.G.K.; writing—original draft preparation, N.E.; writing—review and editing,
N.E. and K.G.K.; visualization, K.G.K., T.M.Y.K. and I.A.B.; supervision, K.G.K., T.M.Y.K. and I.A.B.;
project administration, K.G.K., T.M.Y.K. and I.A.B.; funding acquisition, T.M.Y.K. and I.A.B. All
authors have read and agreed to the published version of the manuscript.
Funding: This research was funded by King Khalid University through grant number (R.G.P.
2/107/41).
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: This study didn’t report any data.
Acknowledgments: The authors extend their appreciation to the Deanship of Scientific Research
at King Khalid University for funding this work through research groups program under grant
number (R.G.P. 2/107/41). The authors also thank the financial help provided by KLE Technological
university through the capacity buildings grants.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Wanhill, R.J.; Bray, G.H. Aerostructural design and its application to aluminum–lithium alloys. In Aluminum-Lithium Alloys,
1st ed.; Prasad, N.E., Gokhale, A., Wanhill, R.J.H., Eds.; Butterworth-Heinemann: Oxford, UK, 2014; pp. 27–58.
2. Lynch, S.P.; Wanhill, R.J.; Byrnes, R.T.; Bray, G.H. Fracture toughness and fracture modes of aerospace aluminum–lithium alloys.
In Aluminum-Lithium Alloys, 1st ed.; Prasad, N.E., Gokhale, A., Wanhill, R.J.H., Eds.; Butterworth-Heinemann: Oxford, UK, 2014;
pp. 415–455.
3. Edgar, A.; Starke, J.R. Historical Development and Present Status of Aluminum Lithium Alloys. In Aluminum-Lithium Alloys,
1st ed.; Prasad, N.E., Gokhale, A., Wanhill, R.J.H., Eds.; Butterworth-Heinemann: Oxford, UK, 2014; pp. 3–26.

67
Materials 2022, 15, 1590

4. Dorin, T.; Vahid, A.; Lamb, J. Aluminium lithium alloys. In Fundamentals of Aluminium Metallurgy, 1st ed.; Lumley, R.M., Ed.;
Woodhead Publishing: Duxford, UK, 2018; pp. 387–438.
5. Hafley, R.A.; Domack, M.S.; Hales, S.J.; Shenoy, R.N. Evaluation of Aluminum Alloy 2050-T84 Microstructure and Mechanical Properties
at Ambient and Cryogenic Temperatures; Technical Report for NASA; Langley Research Center: Hampton, VA, USA, August 2011.
6. Chemin, A.E.; Afonso, C.M.; Pascoal, F.A.; Maciel, C.D.; Ruchert, C.O.; Bose Filho, W.W. Characterization of phases, tensile
properties, and fracture toughness in aircraft-grade aluminum alloys. Mater. Des. Process. Commun. 2019, 1, 1–3. [CrossRef]
7. Mirza, M.S.; Barton, D.C.; Church, P. The effect of stress triaxiality and strain-rate on the fracture characteristics of ductile metals.
J. Mater. Sci. 1996, 31, 453–461. [CrossRef]
8. Clausen, A.H.; Børvik, T.; Hopperstad, O.S.; Benallal, A. Flow and fracture characteristics of aluminium alloy AA5083–H116 as
function of strain rate, temperature and triaxiality. Mater. Sci. Eng. A 2004, 364, 260–272. [CrossRef]
9. Singh, N.; Singh, V. Effect of temperature on tensile properties of near-α alloy Timetal 834. Mater. Sci. Eng. A 2008, 485, 130–139.
[CrossRef]
10. Khan, A.S.; Liu, H. Strain rate and temperature dependent fracture criteria for isotropic and anisotropic metals. Int. J. Plast. 2012,
37, 1–5. [CrossRef]
11. Anderson, D.; Winkler, S.; Bardelcik, A.; Worswick, M.J. Influence of stress triaxiality and strain rate on the failure behavior of a
dual-phase DP780 steel. Mater. Des. 2014, 60, 198–207. [CrossRef]
12. Roth, C.C.; Mohr, D. Effect of strain rate on ductile fracture initiation in advanced high strength steel sheets: Experiments and
modeling. Int. J. Plast. 2014, 56, 19–44. [CrossRef]
13. Rincon, E.; Lopez, H.F.; Cisneros, M.M.; Mancha, H.; Cisneros, M.A. Effect of temperature on the tensile properties of an as-cast
aluminum alloy A319. Mater. Sci. Eng. A 2007, 452, 682–687. [CrossRef]
14. Natesan, E.; Ahlström, J.; Manchili, S.K.; Eriksson, S.; Persson, C. Effect of Strain Rate on the Deformation Behaviour of A356-T7
Cast Aluminium Alloys at Elevated Temperatures. Metals 2020, 10, 1239. [CrossRef]
15. Pandya, K.S.; Roth, C.C.; Mohr, D. Strain rate and temperature dependent fracture of aluminum alloy 7075: Experiments and
neural network modeling. Int. J. Plast. 2020, 135, 102788. [CrossRef]
16. Ekabote, N.; Kodancha, K.G.; Kudari, S.K. Suitability of Standard Fracture Test Specimens for Low Constraint Conditions. Conf.
Ser. Mater. Sci. Eng. 2021, 1123, 012033. [CrossRef]
17. Ekabote, N.; Kodancha, K.G.; Revankar, P.P. Elastic-plastic fracture analysis of anisotropy effect on AA2050-T84 alloy at different
temperatures: A numerical study. Frat. Integrita Strutt. 2022, 16, 78–88.
18. Gupta, M.; Alderliesten, R.C.; Benedictus, R. A review of T-stress and its effects in fracture mechanics. Eng. Fract. Mech. 2015,
134, 218–241. [CrossRef]
19. O’dowd, N.P.; Shih, C.F. Family of crack-tip fields characterized by a triaxiality parameter—I. Structure of fields. J. Mech. Phys.
Solids. 1991, 39, 989–1015. [CrossRef]
20. Hutař, P.; Seitl, S.; García, T.E.; Fernández-Canteli, A. Experimental and numerical analysis of in-and out-of plane constraint
effects on fracture parameters: Aluminium alloy 2024. Appl. Comput. Mech. 2013, 7, 53–64.
21. Fernández-Canteli, A.; Giner, E.; Fernández-Zúniga, D.; Fernández-Sáez, J. A unified analysis of the in-plane and out-of-plane
constraints in 3-D linear elastic fracture mechanics. In Proceedings of the 19th European Conference on Fracture, Kazan, Russia,
26–31 August 2012.
22. Kudari, S.K.; Kodancha, K.G. 3D Stress intensity factor and T-stresses (T11 and T33) formulations for a Compact Tension specimen.
Frat. Integrita Strutt. 2017, 11, 216–225. [CrossRef]
23. Wang, X. Elastic T-stress for cracks in test specimens subjected to non-uniform stress distributions. Eng. Fract. Mech. 2002,
69, 1339–1352. [CrossRef]
24. Neimitz, A.; Galkiewicz, J. Fracture toughness of structural components: Influence of constraint. Int. J. Press. Vessel. Pip. 2006,
83, 42–54. [CrossRef]
25. Galkiewicz, J.; Janus-Galkiewicz, U. The Numerical Analysis of the In-Plane Constraint Influence on the Behavior of the Crack
Subjected to Cyclic Loading. Materials 2021, 14, 1764. [CrossRef] [PubMed]
26. Kim, Y.K.; Oh, B.T.; Kim, J.H. Effects of Crack Tip Constraint on the Fracture Toughness Assessment of 9% Ni Steel for Cryogenic
Application in Liquefied Natural Gas Storage Tanks. Materials 2020, 13, 5250. [CrossRef]
27. ASTM E399-20; Standard Test Method for Linear-Elastic Plane-Strain Fracture Toughness of Metallic Materials. ASTM Interna-
tional: West Conshohocken, PA, USA, 2020.
28. Brian Falzon. The Airbus A380 Wing Cracks: An Engineer’s Perspective. Available online: https://theconversation.com/the-
airbus-a380-wing-cracks-an-engineers-perspective-5318 (accessed on 22 January 2022).
29. Rioja, R.J.; Liu, J. The evolution of Al-Li base products for aerospace and space applications. Metall. Mater. Trans. A 2012,
43, 3325–3337. [CrossRef]
30. ASTM E8/E8M-21; Standard Test Methods for Tension Testing of Metallic Materials. ASTM International: West Conshohocken,
PA, USA, 2021.
31. Mouritz, A.P. Introduction to Aerospace Materials; Woodhead Publishing Limited: Cambridge, UK, 2012; pp. 454–468.
32. Ekabote, N.; Kodancha, K.G. Temperature and test specimen thickness (TST) effect on tensile and fracture behavior of AA2050-T84
alloy. Mater. Today Proc. 2021, in press. [CrossRef]

68
Materials 2022, 15, 1590

33. Dassault Systemes. Abaqus Analysis User Guide. Available online: http://wufengyun.com/v6.14/books/usb/default.htm?
startat=pt05ch22s02abm03.html (accessed on 22 January 2022).
34. Farraro, R.; McLellan, R.B. Temperature dependence of the Young’s modulus and shear modulus of pure nickel, platinum, and
molybdenum. Metall. Trans. A 1977, 8, 1563–1565. [CrossRef]
35. Borges, M.F.; Antunes, F.V.; Prates, P.A.; Branco, R.; Vojtek, T. Effect of Young’s modulus on fatigue crack growth. Int. J. Fatigue
2020, 132, 105375. [CrossRef]
36. Abd El-Aty, A.; Xu, Y.; Guo, X.; Zhang, S.H.; Ma, Y.; Chen, D. Strengthening mechanisms, deformation behavior, and anisotropic
mechanical properties of Al-Li alloys: A review. J. Adv. Res. 2018, 10, 49–67. [CrossRef] [PubMed]
37. Kavale, S.M.; Kodancha, K.G.; Ekabote, N. Effect of Poisson’s ratio on KI, T11 and T33 for SENB and CT specimen–A FE study.
Procedia Struct. Integr. 2019, 14, 584–596. [CrossRef]

69
materials
Review
Void-Induced Ductile Fracture of Metals:
Experimental Observations
Wiktor Wciślik 1, * and Sebastian Lipiec 2

1 Faculty of Civil Engineering and Architecture, Kielce University of Technology, 25-314 Kielce, Poland
2 Faculty of Mechatronics and Mechanical Engineering, Kielce University of Technology, 25-314 Kielce, Poland
* Correspondence: [email protected]

Abstract: The paper presents a literature review on the development of microvoids in metals, leading
to ductile fracture associated with plastic deformation, without taking into account the cleavage
mechanism. Particular emphasis was placed on the results of observations and experimental studies
of the characteristics of the phenomenon itself, without in-depth analysis in the field of widely
used FEM modelling. The mechanism of void development as a fracture mechanism is presented.
Observations of the nucleation of voids in metals from the turn of the 1950s and 1960s to the present
day were described. The nucleation mechanisms related to the defects of the crystal lattice as well as
those resulting from the presence of second-phase particles were characterised. Observations of the
growth and coalescence of voids were presented, along with the basic models of both phenomena. The
modern research methods used to analyse changes in the microstructure of the material during plastic
deformation are discussed. In summary, it was indicated that understanding the microstructural
phenomena occurring in deformed material enables the engineering of the modelling of plastic
fracture in metals.

Keywords: ductile fracture; material testing; void nucleation; growth; coalescence; microstructure;
material characterisation

Citation: Wciślik, W.; Lipiec, S.


Void-Induced Ductile Fracture of
1. Introduction
Metals: Experimental Observations.
Materials 2022, 15, 6473. https://
Ductile fracture criteria can be classified based on the physical sense of the quantities
doi.org/10.3390/ma15186473 defining them. If the criterion includes values derived by macroscopic analysis based on
mechanics of deformable bodies, without precisely defined areas of fracture initiation, a
Academic Editor: Javad Mola
global criterion is formulated. The most known and commonly used global ductile fracture
Received: 30 July 2022 criteria (also known as phenomenological) were determined on the basis of such quantities
Accepted: 14 September 2022 as J integral and crack tip opening displacement δT.
Published: 18 September 2022 Another approach involves analysis of the stress and strain (or other parameter) fields
in the most stressed areas of the material (process zone). The stress and strain state in
Publisher’s Note: MDPI stays neutral
the process zone determines the strength of the whole element. Such an approach and
with regard to jurisdictional claims in
arising criteria are local. The criterion values of local models depend only on the material
published maps and institutional affil-
iations.
properties.
An undoubted advantage of using a local approach to the analysis of the fracture is
the independence of the obtained results on the geometry of the samples used. The local
approach is based on actual physical phenomena (degradation of the microstructure) that
Copyright: © 2022 by the authors. occur in the material subjected to load. In general, understanding and describing these
Licensee MDPI, Basel, Switzerland. phenomena allow for their inclusion in the cracking description of any element type, and
This article is an open access article thus the development of a universal, comprehensive method of predicting the durability of
distributed under the terms and structural elements and assessing the safety of their work.
conditions of the Creative Commons Conducting an analysis according to a local approach requires the use of hybrid meth-
Attribution (CC BY) license (https:// ods: obtaining data from experimental research (with accompanying in situ analyses, e.g.,
creativecommons.org/licenses/by/ video extensometer, recording of acoustic emission signals, using computed tomography
4.0/).

Materials 2022, 15, 6473. https://doi.org/10.3390/ma15186473 71 https://www.mdpi.com/journal/materials


Materials 2022, 15, 6473

techniques) and supplementing them with numerical, metallographic and fractographic


analyses.
Among the physical phenomena taken into account in the local approach to fracture,
the nucleation and growth of microvoids in the structure of the material subjected to
plastic deformation is of particular practical importance. Observation of this phenomenon
is the basis for formulating theoretical models of nucleation, growth and coalescence
of voids, which in turn enables the establishment of models of plastic materials with
microdamages. These issues are therefore crucial for assessing the safety of structural
elements. This article is a literature review on the microstructure observations and void
growth in deformed metals. Particular emphasis was placed on ductile fracture associated
with substantial plastic deformation, without taking into account cleavage cracking. The
results of pioneering works from the second half of the twentieth century, as well as
the latest results obtained with advanced research tools (such as microtomography), are
discussed. The structure of the article is divided into sections according to the different
mechanisms in ductile fracture: Section 2 provides general information on void-induced
failure; further, the nucleation of voids (Section 3) and subsequent stages of their growth
and coalescence (Section 4) are described.

2. Cracking of Metals by the Development of Voids


Void initiators are primarily defects in the crystalline structure (point, linear and
planar defects) and second-phase particles, which can be introduced into the material
microstructure intentionally (e.g., metal matrix composites) or constitute an undesirable
contamination of the material. All these discontinuities, being local stress concentrators,
under plastic deformation create voids with sizes in the order of tenths of a micrometre.
As the plastic strain increases, the voids created in this way increase their size many times
and then coalesce, forming a larger defect and leading to failure. An example of a sample
structure in which a fracture developed as a result of the growth and coalescence of voids
is presented in Figure 1a.
Figure 1b presents a microscopic photograph of the fracture surface of the S355 steel
sample damaged due to the development of voids. The characteristic dimple microstructure
is observed, in which the voids of a few μm in size are surrounded by plastically deformed
ligaments (light bands in Figure 1b) [2].

ȱ
(a)ȱ

Figure 1. Cont.

72
Materials 2022, 15, 6473

ȱ
(b)ȱ

Figure 1. Void-induced failure of metals: (a) Formation of a macroscopic crack due to void growth
and coalescence in X52 steel [1]; (b) Fracture surface of a tensile S355 steel specimen and dimple
microstructure resulting from void development (authors’ own study).

A separate issue is the macroscopic observation of fracture surfaces at void-induced


failure. Ductile fracture of a round tensile bar has been relatively well-documented in the
literature [3]. Initially, voids develop mainly in the centre of the sample, whereby the crack
propagates approximately perpendicular to the sample axis, in the plane of the minimum
cross section (in the neck area). As the crack approaches the sample surface, the influence
of constraints decreases, so the crack deviates from its initial plane. As a result, the fracture
of the sample takes the form of a cup–cone fracture (Figure 2), in which the central part is
flat (cup) and an inclined area (cone) is formed around it.

ȱ
Figure 2. An example of a S355 steel specimen failure due to void development. Characteristic
cup–cone shape is clearly visible [authors’ own study].

An in-depth analysis of microstructures with the use of different methods allowed


for the development of the systematics of the mechanisms of the formation of voids. This
applies to both the nucleation of the voids and their subsequent development.
The initial experimental observations of the development of voids in plastically de-
formed metals took place at the turn of the 1950s and 1960s [4]. Since that time, a number
of research methods have been developed to observe the expansion of voids in metals.
The most common studies include the use of optical microscopy and scanning electron
microscopy (SEM). The first attempts of 3D characterisation of voids involved the use of
indirect methods, such as change in elastic stiffness [5,6] and change in resistivity and
density of the sample [7,8]. A common disadvantage of indirect methods is the ability to
measure only the volume fraction of voids, without taking into account their distribution,

73
Materials 2022, 15, 6473

shape, size and other geometric features [9]. The use of X-ray tomography offers much
more possibilities in this respect.
Section 3 describes the results obtained with the above-mentioned methods, taking into
account different nucleation mechanisms and different types of void initiators. Section 4
presents the observation of the growth and coalescence of voids, as well as characterises
the basic analytical models of these phenomena.

3. Void Nucleation in Metals


3.1. Overview
In general, voids are nucleated in the vicinity of stress concentrators on a microscopic
scale [10]. In the literature, the most frequently mentioned mechanisms include the for-
mation of voids at the intersection of the slip bands [11], at the grain boundaries [12], at
twin boundaries and at vacancy clusters [13], but most importantly around inclusions and
precipitates [14–16].

3.2. Nucleation with No Particles Involved


In the absence of material discontinuities, the formation of voids is observed as planar
slip band decohesion or grain boundary decohesion (homogeneous nucleation). One of the
first widely known studies in which the participation of dislocations in the formation of
voids was experimentally confirmed was developed by Gardner et al. [12], who noticed
that the dislocation structures in iron and beryllium crystals evolved into cells at high
strains. The boundaries between individual cells were characterised by sufficient surface
energy to create a void, even in the absence of other internal stress concentrators.
The observations of the development of voids in the Nb-Cr-Ti alloy [11] showed that
the voids formed in this way are characterised by a flattened shape.
The concept of void nucleation as a result of vacancy condensation and the presence of
dislocation boundaries has been known for a long time [17,18], but only modern research
techniques allowed for a better understanding and documentation of this phenomenon [19].
For example, in [20], by investigating the fracture mechanism of copper containing copper
oxide particles, it was found that the voids are first nucleated at the nanoscale, most
often without any relation to the presence of the second-phase particles. An example of a
photograph of nanovoids, made with the use of the high-angle annular dark-field scanning
transmission electron microscopy (HAADF-STEM), is shown in Figure 3.

ȱ
Figure 3. (a) HAADF-STEM image presenting nanoscale voids (black areas) in copper subjected
to tension (plastic strain); (b) Enlargement of the boxed area in (a), from [20]. White rectangle in
(b) indicates the twin boundary.

In the next stage, only a small quantity of the nanovoids grew to the microscale,
contributing to the initiation of cracking. It was noted that all microscale voids were
associated with a dislocation boundary. By carefully analysing the microstructure of the

74
Materials 2022, 15, 6473

material in the neck area, the authors distinguished three basic groups of microvoids,
depending on their location and the mechanism of formation, namely: (i) voids related to
intragranular, inclusion-free dislocation boundaries; (ii) voids associated with the inclusion-
free intersection between one or more dislocation boundaries and one or more grain
boundaries; (iii) voids associated with an inclusion intersected by a dislocation boundary.
As mentioned before, voids can be initiated by grain separation. This mechanism is
favoured by the following conditions: high share of the hydrostatic stress; small value of
the void spacing to void diameter ratio; and high value of the precipitate-free zone (PFZ)
thickness to void spacing ratio [21].

3.3. The Role of Second-Phase Particles


In technical alloys, voids are often initiated by the failure of second-phase particles,
randomly distributed over the material matrix (heterogeneous nucleation). The presence
of a hard particle locally limits elastic and plastic deformation of the matrix, which in
turn causes local stress concentration. With the increase in plastic strain, the stress value
also increases, which ultimately leads to the particle cracking or its separation from the
matrix [22]. As the result, a void is formed (Figure 4).

ȱ
Figure 4. Two mechanisms of void nucleation in AZ31 magnesium alloy: (a) Particle fragmenta-
tion/fracture; (b) Matrix–particle separation [23].

The occurrence of one of the two mentioned mechanisms depends largely on the
mechanical properties of the matrix and the particle (strength, ductility) and the strength of
the matrix–particle interface. In general, the phenomenon of particle separation is primarily
observed in relatively soft, ductile matrices. However, high yield stress and hardening
exponent of the matrix as well as the high particle stiffness promote particle cracking.
Moreover, the nucleation mechanism depends on the particle geometry (size, shape)
and its orientation with respect to the direction of the principle tensile stresses. Larger
particles usually break, as do elongated particles parallel to the principal stress direction.
In practice, engineering materials have different void populations; therefore, both mech-
anisms occur simultaneously. The local stress state is also of great importance, namely
the predominance of normal stress over shear stress, which most often forces the particle
fracture [24,25].
Due to the heterogeneous stress state changing with time, the nucleation of voids is
continuous across the entire range of plastic strain. In other words, the development of
microdamage includes the simultaneous growth and coalescence of existing voids as well
as the nucleation of new ones [26].

75
Materials 2022, 15, 6473

Among the experimental studies, the process of nucleation of voids and microcrack
formation by fracture of silicon particles in Al-Si alloys is relatively well-documented.
The most important results described in the literature are summarised in [27]. As shown
in [28,29], cracking of silicon particles and their separation from the matrix were observed
already at strain values of 1–2%, while the quantity of damaged particles increased linearly
with the increasing strain [30]. According to the results of observations described in [31],
a maximum of 10% of the particles were damaged in the samples subjected to tension or
compression. In general, larger particles crack first [32], which is often explained by a
greater probability of internal defects in this type of particles. Smaller-sized particles tend
to detach from the matrix, initiating voids.
As noted in [30], the phenomenon of fracture of silicon particles in the Al-7Si-0.4Mg
alloy subjected to tension and bending occurs mainly in the case of elongated particles.
The authors also analysed the structure of the alloy deformed to failure. In the areas with
a homogeneous deformation (a few mm from the fracture surface) between 3 and 10% of
the Si particles cracked. However, in the immediate vicinity of the fracture surface, the
proportion of fractured particles locally increased to about 15–20%. In the coarser structures,
the fracture of the particles was sudden and occurred at low strains. The development of
microdamage in the finer structures was gradual.

3.4. Estimating Particle Strength


To understand the phenomenon of void nucleation by fracture of the second-phase
particles, it is important to estimate their mechanical properties, especially strength. The
first and best-known attempts were described in [33,34]. In [34], the microscopic maps of
the locations of cracked and separated MnS inclusions were compared with the numerically
determined stress and strain distributions, which gave rise to assessment of the critical
fracture stress of MnS particles of about 1120 MPa. Additionally, in a similar manner, the
critical stress of separation of the MnS particle and the matrix was determined, where the
stress value was about 810 MPa.
A similar approach was used in [35], but this time the X-ray technique was used
to assess the particle deformation. The tests were carried out on a sample made of an
aluminium alloy, subjected to pure bending. Using the X-ray method, the strain values of
silicon particles were determined, and then the stresses in the particles were determined as
a function of the measured strains.
The typical research methods used so far did not take into account the internal defects
of the particles; therefore, the obtained results are not precise.
The significant progress made in recent years in the field of methods of microstructural
materials testing allowed for a more in-depth analysis of this issue. An interesting attempt
to determine the strength of silicon particles in the A356 aluminium alloy is described
in [36]. In order to expose the silicon particles, the ground surface of the sample was
subjected to deep etching with a mixture of phosphoric, acetic and nitric acids. As a result,
the particles to be tested were exposed to a height of several dozen micrometres. In the next
step, using the focused ion beam (FIB) method, the geometric notch of the selected particles
was cut. The prepared particles were subjected to an eccentric compression test in which
the load was transferred to the particle by a tungsten needle. The test scheme is shown in
Figure 5. The entire course of the process, up to the particle fracture, was recorded using
a scanning electron microscope (SEM) with the video recording. In the next stage, based
on the values of the measured force, the evaluation of the stress values in the particle was
performed, using the simplified analytical method and with the use of the finite element
method (FEM). For the defect-free particles, the particle strength was determined to be
around 16 GPa. Detailed visual inspection of particles using SEM, combined with FEM
modelling, made it possible to evaluate the effect of different types of particle defects on
their strength. It was found that the presence of defects can reduce the strength of the
silicon particles to 2–3 GPa.

76
Materials 2022, 15, 6473

ȱ
Figure 5. Strength test of silicon microparticles in A356 aluminium alloy: (a) Scheme of the microme-
chanical test; (b) C-shaped particle and the tungsten tip before test; (c) Last frame before particle
fracture, from [36].

The article [37] also provides an example of particle strength assessment by means
of a microscopic three-point bending test. First, the silicon particles were extracted from
the Al-Si alloy by dissolving the aluminium matrix. Then, after cleaning and selecting the
particles, microscopic beams were cut from them using the above-mentioned FIB technique.
The beam prepared in this way was placed on a steel base with a cut-out hole and subjected
to load. Figure 6 illustrates the tested sample. The results obtained during the experiment
were compared with the results of analytical and numerical calculations, which resulted in
determining the strength of silicon at the level of about 9 GPa, assuming no particle defects.

ȱ
Figure 6. (a) Eutectic silicon particle extracted from the Al-Si alloy; (b) Microscopic three-point
bending specimen prepared from the particle in (a), from [37].

A significant technical problem during the microscale strength tests is the measurement
of deformations and stresses. Recently, the Raman spectroscopy technique offers a wide
range of analysis of stress values on a microscopic scale, and thus also the evaluation of
particle strength. The Raman effect is related to inelastic light scattering. After filtering,
the light scattered on the sample goes to the spectrometer, which records its spectrum. The
spectrum is presented as a function of the Raman shift, defined as the difference between
the frequency of the scattered light and the input light [38]. In materials science, spectrum
analysis enables the recording of the stress values, especially in the case of uniaxial stress
states. However, the authors of [39] present the methodology of plane stress analysis in Si
wafers on a microscopic scale.
A more complex analysis is discussed in [40], where the stress state in eutectic silicon
particles in the Al-Si alloy was analysed. The alloy was tested in as-received condition
under uniaxial loading. Cracking of silicon particles was observed already at the stress
value of 600 MPa. Importantly, the use of Raman spectroscopy enables the assessment of
the effect of particle size and its neighbourhood on the strength.

77
Materials 2022, 15, 6473

As already mentioned above, the second mechanism of void formation involves the
decohesion of the matrix and the second-phase particles. Due to technical difficulties in a
detailed, experimental study of this phenomenon, there are relatively few works of this
type. Thus, numerical analyses are of particular importance. The solutions described in
the literature are most often theoretical in nature. For example, in [41–43] a hypothetical
aluminium matrix–silicon particle interface strength was determined in the range of about
4–7 GPa in tension, while the shear strength was only about 300 MPa. As this paper deals
primarily with experimental research, these issues will not be described in detail here. More
information on the simulation of this phenomenon, and the use of cohesive models, can be
found in [44–47].

3.5. Effect of Martensite Cracking in DP Steels


As it has been shown in many studies, for example [48], the mechanism of void
formation in dual-phase (DP) steels is only slightly based on the fracture and separation of
the second-phase particles, because the fracture of martensite plays the most important role.
The development of microdamage of DP1000 steel subjected to tension was analysed in
detail in [48]. Strength tests were carried out inside the scanning electron microscope (SEM)
chamber and stopped at regular intervals, each time photographing the microstructure of
the material in the selected area. Cracking of the particles was observed at overall small
strain, of the order of 2%. With the increase in plastic deformation, the voids created
in this way grew, but no crack development in its vicinity was observed. As the strain
value was increased further, the martensite phase cracked. Due to increasing deformation,
the crack turned into a void, which, being a stress concentrator, became the cause of
crack propagation in the ferrite. The high intensity of this phenomenon accounted for its
dominant role in the failure of DP1000 steel.
Using advanced research methods, the authors of [48] attempted to estimate the
local values of strains and stresses accompanying martensite cracking and void initiation.
The obtained photographs of the microstructure were processed using the digital image
correlation (DIC) technique. The photograph of the undeformed structure was divided into
subset windows, then an algorithm was applied to track these areas in the photographs of
the deformed structure. The results obtained in this way made it possible to determine the
vectors of displacements and local deformations. Further, the experimentally determined
values of displacements were used as boundary conditions in the finite element method
(FEM) model of the tested sample, which allowed for the estimation of martensite cracking
stress at the level of about 1700 MPa.
The dominant role of martensite cracking in the formation of voids in DP steels, mainly
of the coarse structure, was also emphasised in [49]. This is the leading mechanism at
low strains. At a later stage, voids were formed mainly by the decohesion of ferrite and
martensite. The latter mechanism dominated in steels with the finer structure in the entire
range of deformation. The occurrence of the decohesion is mainly attributed to the lower
deformability of martensite.
Observations of martensite cracking in DP600 steel under uniaxial tension at low strain
values were also described in [50], although as the authors point out, the importance of this
mechanism in the entire process of void nucleation is not great. The process of fracture and
separation of second-phase particles led to the formation of a few voids, which, however,
were characterised by large dimensions, and therefore their area fraction was significant.
The dominant mechanism for the formation of voids was, as in [51], decohesion at the
interface between ferrite and martensite, observed in the entire range of deformation of the
tensile sample. The voids were initiated mainly at the interfaces perpendicular to tensile
stresses and enlarged along the ferrite grains. The authors of [50] also noticed that with the
increase in strain, the mean size of the voids decreased, which indicates a high intensity of
nucleation of new voids also immediately before failure.
The authors of [52] drew similar conclusions. While examining the development of
microdamage in DP600 and DP800 steels under uniaxial tension, it was noticed that at low

78
Materials 2022, 15, 6473

strain, the void initiators were the globular aluminium oxide inclusions, but with higher
deformations the voids were formed near to the ferrite–martensite interfaces as well as in
the ferrite matrix and close to martensite islands.

3.6. Quantitative Description of Void Nucleation


Regardless of the single void nucleation analysis, it is important to evaluate the void
nucleation globally, determining the number of nucleated voids as a function of remote
strain and the location of the analysed area. In this case, it is particularly important to con-
tinuously track changes in the microstructure of the material throughout the deformation
range up to the failure.
In recent years, the widespread use of the X-ray microtomography method gave great
opportunities in this regard, and has contributed to a much better understanding of the
phenomenon of failure and void development [9,53].
For example, in [54], the tomography method was used to assess changes in the
microstructure of JIS SUM24L free-cutting steel under uniaxial tension. The tests were
carried out on tensile specimens subjected to uniaxial stress state. The tests were interrupted
at various stages, each time taking tomographic photographs of the microstructure in one
selected area. In order to precisely determine the value of strain, especially after necking,
changes in the width of the specimen were recorded using tomographic images. As
part of the microtomographic research, in the first step, the region of interest (ROI) was
distinguished, along with the voids present in the unstrained material. The photographs
were then binarised, indicating base material and voids. In the next stage, a 3D labelling
algorithm was applied to the binarised images, and then the volume and position of each
of the detected voids was determined. In this way, over four thousand voids and the
second-phase particles were marked in the ROI, which allowed for their tracking in the
entire range of given deformation.
A separate issue was the development of an algorithm that allows for tracking each of
the voids in subsequent stages, with increasing values of strain. The adopted procedure
included the determination of a transformation matrix that was calculated by minimising
the sum of distance difference between corresponding pairs of objects detected in subse-
quent stages of loading. Due to the heterogeneity of deformation and the different quality
of individual photos, the obtained results were not fully accurate, hence the matching
probability parameter Mp was introduced into the analysis. The algorithm developed in
this way took into account the translation and rotation of the voids related to the occurrence
of plastic deformation. A detailed description of the voids tracking algorithm is presented
in [55,56].
The results of the observation of microstructure changes indicate that the nucleation of
voids was continuous throughout the analysed range of deformations, up to failure. This
phenomenon is well-illustrated by the graph in Figure 7, where the number of voids in
the unit volume [mm3 ] of ROI was determined as a function of true strain. From the very
beginning, a constant increase in the number of voids is visible. Just before failure, one can
see a flattening of the curve, which seemingly means a reduction in the intensity of void
nucleation at this stage. In fact, as noted by the authors of [54], nucleation of voids is still
present; however, at this stage, the ductile fracture process is controlled primarily by the
void coalescence phenomenon, which explains the flattening of the curve in Figure 7.
Figure 8 presents tomographic cross sections of ROI, made at different strains. The
vertical axis in the individual figures is equated with the direction of loading. Dark areas
represent voids, while lighter areas represent the matrix. In Figure 8b (with the strain 0.23),
the neck is clearly visible.

79
Materials 2022, 15, 6473

Figure 7. Experimentally determined relation between density of voids and plastic strain in JIS
SUM24L free-cutting steel, from [54].

ȱ
Figure 8. Free-cutting steel (JIS SUM24L grade) microstructure at different strains: (a) 0; (b) 0.23;
(c) 0.50; (d) 0.64, from [54].

Detailed analysis of the individual pictures shows that the nucleation of voids occurs
primarily at the interface between the matrix and the particles. The voids are also initiated
at the places where the particles of the second phase break. As can be seen from the
comparison of the subsequent photographs, some voids seem to disappear as the strain
increases. This is due to their rotation and displacement, because the strains are not uniform
throughout the tested sample. The void-tracking algorithm described above allowed for
the inclusion of this phenomenon in the void development analysis.
Additionally, on the basis of the obtained results of tomographic examinations, changes
in the volume and diameter of voids as a function of plastic strain were determined. While,
as predicted, the total volume of the voids in the sample increased with increasing strain,
the mean diameter of the voids was almost constant, regardless of the strain level. The
authors of [54] indicate the nucleation of new small voids at higher strain levels as a possible
reason.
The authors of [57] also emphasised the key role of small voids in the initiation of
fracture. Using the 4D X-ray microtomography technique (3D + time), changes in the
microstructure of SA508 steel were analysed in the entire range of tensile strains up to
failure. It was observed that large Al2 O3 and MnS2 particles (with sizes ranging from
several to tens of micrometres) cracked or separated from the matrix at zero values of
plastic strain (elastic range). Considering the fact that such particles were scattered and
spaced far apart, the voids they initiated did not coalesce, and therefore their contribution
to the fracture initiation was insignificant. On the other hand, the elongation of voids along

80
Materials 2022, 15, 6473

the tensile axis was observed, but the their contraction in the perpendicular direction was
not significant. Moreover, the rotation of the elongated voids took place as a result of the
increase in the value of shear stresses after the formation of the neck.
On the contrary, small particles of cementite (with a size of the order of 100–500 nm)
detached from the matrix when the remote strain of the order of two was achieved. The
initiation of small voids was, however, sudden. A large accumulation of small voids
favoured their coalescence, which led to formation of a crack. The microtomographic image
of the microstructure of the sample before failure is shown in Figure 9. In the central part
of the region of interest, a cluster of small voids initiating a crack is clearly visible.

ȱ
Figure 9. Microtomographic 3D image of the distribution of voids in SA508 steel, at the onset of
failure, in the central part a cluster of small voids initiating a macroscopic defect is present, from [57].
The colour scale indicates the distance between void and sample centre.

Larger elongated voids predominate at a greater distance from the centre of the
sample. The two largest voids (marked in the figure as stringer 1 and 2) were formed from
an agglomerate of particles. Additionally, the authors of the paper carefully analysed the
void nucleation around the inclusion marked as “inclusion 1” in Figure 9, indicating the
mechanism of matrix–particle decohesion.

3.7. Effect of Stress State on Void Nucleation Intensity


In the previously mentioned work [10], the influence of the type of load on the intensity
of void nucleation was determined. Figure 10 illustrates an exemplary dependence of the
number of nucleated voids per unit volume of the cast Al-Si-Mg sample as a function of
strain for various loading conditions. The results were obtained using the proprietary
analytical void nucleation model. The lowest intensity of void nucleation was obtained
for compression and torsion. As expected, among the simple load cases, the voids in the
sample subjected to tension showed the highest nucleation intensity. The simultaneous
action of tension and torsion resulted in the most intense nucleation of the voids.

81
Materials 2022, 15, 6473

Figure 10. Effect of the loading type on the intensity of void nucleation in cast Al-Si-Mg alloy as a
function of strain, based on [10].

Regardless of the above, the authors of [10] defined the material constants of void
nucleation for simple loading conditions (tension/compression, torsion).
Recently, the authors of [58] conducted a thorough experimental analysis of the impact
of the stress state on void nucleation in DP780 and CP800 steels. In order to obtain different
components of the stress state, various strength tests were carried out: simple shear, hole
tension, v-bending and biaxial tension. In each case, the tests were stopped, recording
the material microstructure in the region of interest, using microtomography. Nucleation
intensity was measured as the average number of nucleated voids in 1 mm3 of material in
the process zone. At failure, the highest number of voids (about 30,000/mm3 ) was observed
in the biaxial tension specimens. In the case of hole tension, this value was much lower
and ranged from about 9500 to 18,000, depending on the material tested. The specimens
subjected to shearing were characterised by the lowest nucleation intensity, i.e., at failure,
values between 3000 and 4000 voids in 1 mm3 of material were recorded.
One of the most frequently analysed issues related to the dependence of void nucle-
ation on stress state is the influence of stress triaxiality T on the value of the strain needed
to initiate the void. Stress triaxiality T describes the effect of the spherical component of
the stress tensor (hydrostatic tension or compression) and is defined as the quotient of the
mean stress (arithmetic mean of the principal stresses) and the Huber von Mises stress.
The works published so far, for example, [59,60], unanimously indicate that the increase in
triaxiality (increase in the hydrostatic pressure share) is accompanied by an exponential
decrease in the value of nucleation strain.
The author of [61] drew similar conclusions, at the same time indicating the large
influence of the Lode parameter on nucleation and the growth of voids. The Lode parameter
takes into account the influence of the third stress tensor invariant. According to Yu [61], the
value of the Lode parameter does not significantly affect the value of the void nucleation
strain; however, it plays an important role, as interfacial cracks nucleate from different
positions for different Lode parameters and propagate in different patterns. This is due to
the fact that the Lode parameter changes the principal stress distribution, even at constant
triaxiality.
Han et al. [62], studying the development of voids in QP980 steel under shear load,
noticed that a large number of small voids (less than 5 μm in size) was formed at phase
interfaces. In turn, a few microvoids generated from inclusions had more than 5 μm.
The phenomenon of the development of voids under shear was also analysed in detail
by the authors of [63], also indicating the low intensity of nucleation in these conditions. The
combination of microtomographic tests with FEM simulation allowed for the determination
of the mechanism of ductile fracture of FB600 steel, initiated by separation of the matrix from
CaO particles. The voids created in this way grew towards the largest local deformations,
forming microcrack-like defects. As noted, a shear-band type of failure was formed on the

82
Materials 2022, 15, 6473

microscopic scale even with a small volume fraction of voids, of the order of 0.015%. The
void volume fraction measured before failure did not exceed 0.1%.
In recent decades, computer simulations have made a huge contribution to under-
standing the phenomena of void development [64]. As this article focuses primarily on
experimental observations, the review of FEM results will not be discussed in detail here.
However, it is worth paying attention to molecular dynamics simulation [65,66], which
offers new possibilities compared to traditional continuum solutions, as it enables material
modelling at the atomic level. For example, in [67], the mechanism of decohesion of the
AlCu2 particle and the aluminium matrix was analysed. In the first stage, the breaking of
the bonds between single-inclusion and matrix atoms was observed, which initiated the
particle separation. In the next stage, the crack grew steadily, with no dislocation involved.
The fracture development in this case was driven by the lattice trapping phenomenon. After
the fracture reached a critical size, nucleation of Shockley partial dislocations at the crack tip
was observed. Then, the dislocations moved from the particle towards the matrix, whereby
the rate of crack propagation increased suddenly, leading to the complete separation of
the particle and the matrix. The authors called this stage of separation dislocation-mediated
delamination.

4. Void Growth and Coalescence in Metals


4.1. Mechanisms of Growth and Coalescence of Voids
In the course of the realisation of ductile failure, after nucleation (characterised in the
previous section of the article) due to plastic strain and hydrostatic stresses, the voids in
the material increase [68,69]. With the action of strain, the voids grow, change their shape
and move, changing their position. With stable void growth, plastic strain forms relatively
uniformly in the material. However, from a certain point, the strain localises between
adjacent voids. Outside the plane of strain localisation, the material undergoes elastic
unloading. The occurrence of local strain localisation limits the ductility of materials. Two
mechanisms for the realisation of strain localisation have been identified. The first involves
the softening of strain through factors such as microstructural changes, thermal interactions
and damage evolution. There is a local degradation of the load-carrying capacity of the
material, resulting in strain localisation in a thin band [70–73]. The initiation of strain
localisation depends on a number of factors, including stress state, material properties and
material porosity [74–76].
The second mechanism of strain localisation is associated with the phenomenon of
void coalescence. There is a local instability, conditioned by the interaction between adjacent
voids in the material. The moment of localisation of plastic strain is identified with the
beginning of the process of void coalescence. Once the coalescence process has started, the
kinematics of void enlargement differ significantly from the kinematics of void growth prior
to this mode of instability. For void coalescence induced by macroscopic strain localisation,
the width of the localisation band is narrower. This is due to the restriction of deformation
to areas (ligaments) between adjacent voids [76]. The process of void coalescence becomes
the direct cause of the initiation and growth of a ductile crack in the material. There are
three ways to realise void coalescence [77]:
(1) Internal necking (Figure 11a,d);
(2) Shear coalescence (Figure 11b,e);
(3) Necklace coalescence (Figure 11c).

83
Materials 2022, 15, 6473

ȱ
Figure 11. Methods of realising void coalescence in materials (example in steel): (a,d) Internal
necking; (b,e) Shear coalescence; (c) Necklace coalescence [24,78,79].

The internal necking mechanism is the most commonly observed way of the material
void coalescence process, initially observed by Argon [33]. With the internal necking
process, the plane of localisation is almost perpendicular to the main direction of the
applied load. During the process of coalescence of voids, a reduction in the area of the
intervacancy ligament can be observed, which is similar to the necking phenomenon during
the tensile test of the specimen under uniaxial loading [77,80–83]. A large contribution
to understanding the process of coalescence of voids according to the internal necking
mechanism was made by micromechanical analyses involving numerical calculations using
the finite element method [71,84–87].
The second distinguished method of coalescence is the mechanism involving localised
shear occurring between initial voids (of large size), when the voids are distributed along a
line inclined at 45◦ to the main direction of loading (Figure 11b). The stress state accompa-
nying the development of the failure process according to the void mechanism is the subject
of many scientific studies. The stress state is most often defined by the stress triaxiality
factor. However, the value cannot unambiguously describe the effect of stress triaxiality on
void growth and coalescence. For a given value of the triaxiality factor, more than one stress
state exists. The necessary information is very often extended by the magnitude of the Lode
parameter. The development of the stress state in ductile fracture was the subject of the
work. These quantities were determined on the basis of experimental results and numerical
calculations. Selected results will be cited later in this work (Section 4.3). Coalescence in
this case can occur according to the so-called “void sheeting” mechanism [88]. Plasticity is
localised in shear bands containing small secondary voids; large primary voids are con-
nected through coalescence of these secondary voids. This coalescence mechanism is often
observed in high-strength materials with low-to-medium strain-hardening capacity [77,89].
A third possible mechanism for the coalescence of voids is called necklace coalescence
(Figure 11c). It is the least commonly occurring in materials. The mechanism involves
localisation in a direction parallel to the action of the applied load. This mechanism
was observed in areas of voids, which are distributed in elongated concentrations. The
mechanism of necklace coalescence is considered to be of major importance in the process
of occurrence of ductile delamination cracking [77,90,91].

84
Materials 2022, 15, 6473

Many elements influence the development of the growth process and the coalescence
of voids in ductile failure. These may include the state of stress in the material (which can
be defined by the stress triaxiality factor, the Lode parameter), the contribution of shear
stress and the level of plastic strain, taking into account the level of critical strain [92].
On the basis of the above quantities and a wide range of experimental and fractographic
studies, numerical calculations, various failure models have been proposed, a selection of
which will be discussed later in this paper.

4.2. Classical Models for the Growth and Coalescence of Voids


Of the classic local models describing the process of void growth, mention should be
made of the solution proposed by McClintock [93]. In his work, he analysed the growth
of a cylindrical void in a rigid plastic material under axisymmetric loading, assuming a
plane strain condition. The model assumes that the relative volume of voids reaching a
critical value will result in crack initiation. As a continuation of McClintock’s study, an
approximate solution for spherical void growth was proposed by Rice and Tracey [94].
The void growth model developed by Rice and Tracey allowed for the formulation of a
fracture criterion, specifying that a crack will be initiated if the normalised void radius
reaches a critical size. Similar results were obtained in [34,95–97]. In the cited papers, it
was shown that the level of plastic strain and stress triaxiality have a significant impact on
the realisation of the void growth process. The cited works formed the basis for further
research and development of further local ductile fracture models.
A well-known model describing the phenomenon of void growth in ductile fracture
is that by Gurson [98]. The author, with a similar methodology to the work of Rice and
Tracey [94], developed an approach to analyse plastic flow in porous material assuming
material continuity. Gurson included in the model the interaction between voids and
the effect of void growth on material softening. A modification of Gurson’s model was
proposed by Tvergaard and Needleman [3]. The authors made a change in the definition of
the relative volume of voids in the material and added an acceleration factor to account for
the phenomenon of void coalescence. This resulted in a revised description of the plastic
flow of the material in the initial stage of ductile fracture. Expanding on the ideas presented
in the Gurson and GNT models, there have been many studies on the analysis of the growth
and coalescence of voids during ductile fracture [71,99–106].
With analytical models for describing the process of void coalescence, it is worth
mentioning Brown–Embury [107] and Thomason [71,106,108] as base models. These are
models developed at the micromechanical scale. The Brown–Embury model refers to a
perfectly ductile material and assumes the presence of shear bands at 45◦ between voids.
The model relates the possibility of void coalescence depending on the diameter of the voids
R and the distances between their centres (X). The criterion assumes that for a given void
form factor, there is a minimum relative distance between voids, below which coalescence
cannot be initiated, regardless of the stress state. Thomason’s void coalescence model was
developed for elastic–perfectly plastic materials, using solutions for slip lines. For the
axisymmetric problem, Thomason’s model assumed that the average normal stresses affect
the void if the stresses reach a certain specified value.
The aforementioned classical models relating to the growth process and the coalescence
of voids assume a number of simplifications. They do not address all aspects of strain and
ductile failure of real materials. Thus, there is a major role for developing in situ studies of
the development of voids in materials, with a particular focus on metals.

4.3. Experimental Verification of the Growth and Coalescence Process of Voids


The much more complex actual mechanism of ductile failure was highlighted in [109],
taking into account the development of critical strain levels and stress triaxiality. In addition
to the processes of void initiation, growth and coalescence (for tension dominated loading)
and the occurrence of shear and void coalescence (for shear dominated loading), the authors
point to seven types of micromechanisms that occur sequentially or are complementary in

85
Materials 2022, 15, 6473

the failure process. An example of this complex type of ductile failure can be observed in
the occurrence of a cup-and-cone failure on a tensile specimen, particularly for structural
steels with medium-strength characteristics and high levels of ductility (Figure 12).

ȱ
Figure 12. A cup-and-cone failure on a cylindrical, tensile specimen: view of break surface with
scheme of the break plane profile (for S355 steel) [110].

Based on an analysis of the results of experimental uniaxial tensile tests on materials


such as aluminium, nickel and copper, seven different types of ductile failure mechanisms
were demonstrated (Figure 13) [109]:
(1) Intervoid necking (occurring with initiation, growth and coalescence of voids): for
triaxiality stress factor T ≥ 0.33.
(2) Intervoid shearing, for which the initiation and subsequent elongation of voids along
shear bands is characteristic. The consequence is the coalescence of voids and the
formation of macrocracks in the planes of the shear bands. The mechanism occurs for
a triaxiality stress factor T less than 0.33, located in a single plane.
(3) Void sheeting, when shear develops between existing voids in the material and the
simultaneous process of nucleation and coalescence of new voids. The final ductile
fracture that develops connects the existing voids in the material. The mechanism
occurs on multiple planes, for stress triaxiality less than 0.33.
(4) The Orowan alternating slip (OAS) mechanism, which assumes the occurrence of void
nucleation at the intersection of slip bands and the consequent growth of prismatic
voids in the results of alternating slip along shear bands. OAS is characterised by its
occurrence on multiple planes, for a triaxiality factor T less than 0.33.
(5) Destruction by specimen necking (T ≥ 0.33).
(6) Shear leading to the destruction of a specimen, occurring in a single plane. It is realised
by the sliding of the material along a single slip band. Consequently, it will lead to a
loss of cohesion and failure of the specimen (triaxiality factor T ≤ 0.33).
(7) Multiplanar shear occurring along multiple shear bands; also referred to as the
slipping-off mechanism (T ≤ 0.33).
The nature of the interaction in the failure process between these seven mechanisms
depends on, among other things: the local state of stress, the type of microstructure of the
material under consideration (microstructural changes that occur during deformation can
eliminate or create void nucleation sites in the material) and the strain-hardening capacity
of the material (a low strain-hardening capacity favours the interlaminar shear mechanism,
while a high one predisposes to the occurrence of void sheeting). For engineering materials
in the occurrence of ductile failure, the first three mechanisms identified by Noell et al. [109]
will be most relevant.

86
Materials 2022, 15, 6473

ȱ ȱ ȱ

ȱ ȱ

ȱ ȱ ȱ

Figure 13. Types of ductile failure mechanisms: (1) Intervoid necking, (2) Intervoid shearing, (3) Void
sheeting, (4) The Orowan alternating slip, (5) Necking to a point, (6) Single-plane catastrophic shear,
(7) Multiplane catastrophic shear [109].

There are many analytical models in the literature describing ductile failure according
to the void mechanism. An important aspect is the possibility of their experimental
validation. Particularly difficult to validate experimentally is the phenomenon of void
coalescence [111–114]. The use of X-ray computed tomography is helpful during in situ
testing. In order to experimentally verify the growth process and the coalescence of voids,
X-ray tomography was used for the uniaxial tensile testing of copper and Glidcop alloy
in [80]. For this purpose, specimens were specially prepared for testing by laser-drilling
holes in the material. Figure 14 shows the resulting images of growth and coalescence of
the modelled voids. In the case of copper, the process of void coalescence occurred for true
strain levels of more than 100%, while for the Glidcop alloy the strain was in the range
of 50%. No secondary nucleation of voids was observed for copper. The occurrence of
ripple marks on the surface of the voids was identified with the realisation of slip in the
specimen. With the Glidcop alloy, coalescence was realised by nucleation of secondary
voids on alumina particles between the modelled holes. Based on the results obtained,
an attempt was made to verify the selected models of growth and coalescence of voids.
Those models that provide quantitative verification of the results were considered. Good

87
Materials 2022, 15, 6473

agreement was shown for the void growth model proposed by Rice and Tracey when taking
into account the change in stress triaxiality in the specimen with progressive strain. When
verifying the void coalescence process, greater discrepancies from experimental results
were shown for the Brown and Embury model (differences of about 50%) than for the
Thomason model (differences of 2–40%). With discrepancies in the Brown and Embury
model, constraints that cause delayed void coalescence may have an impact. In the case
of Glidcop alloy specimens, Thomason’s model overestimates the level of critical strain
at failure due to the failure to account for the secondary void nucleation occurring in the
material.

ȱ
ȱ ȱ
(a)ȱ (b)ȱ (c)ȱ

ȱ
ȱ
ȱ
(d)ȱ (e)ȱ (f)ȱ

Figure 14. Growth and coalescence of modelled holes in copper specimens at true strain levels:
(a) 0.00; (b) 0.77; (c) 1.01; in Glidcop alloy specimens at true strain levels: (d) 0.00; (e) 0.45; (f) 0.50 [80].

The authors of [77] proposed a criterion for the occurrence of void coalescence in
material. It is based on a modification of the assumptions of the classical Thomason
criterion (and its extension by Benzerga [81]). The criteria mentioned concern the analysis
of the action of normal stresses, excluding the contribution of shear components. However,
shear stresses can play an important role in the process of void coalescence and subsequent
crack initiation in a material. An RVE model was used, which (unlike previous models)
is affected by shear components. A positive calibration of the proposed criterion for void
coalescence with the results of numerical calculations for the three-dimensional model was
carried out. The calibration considered the values of void form factor and void distribution.
The classical Thomason criterion (with a modification by Benzerga) was generalised to any
loading condition. Modifications of Thomason’s criterion for the void coalescence process,
based on—among other things—the results of numerical calculations, are included in [82].
When analysing the growth process and the coalescence of voids in the material, an
important aspect is to determine the volume contribution of the voids and their initial
shape [115–117]. This allowed, among other things, for the subsequent modelling of the cor-
rect shape in the numerical model. In a number of papers, authors based on the Rice–Tracey
model and the assumption of a spherically shaped void obtained oversimplified results.
With the development of computed tomography methods, especially high-resolution μXCT,

88
Materials 2022, 15, 6473

it has been possible to determine the true shapes of voids present in the material at the
microscale level and to include them in the model describing the growth and coalescence
of voids [118–123]. An example of an analysis image using the μXCT technique with
the determined shape of the voids and their representation in the numerical calculation
programme is shown in Figure 15 [124]. In this paper, on the basis of μXCT observations
and numerical calculations, an attempt was made to determine the influence of the initial
void shape (spherical, cylindrical, elliptical) on the macroscopic description of the void
growth process in the material as an important stage of material destruction according to
the ductile mechanism [125]. The effects of stress triaxiality, shape factor and the orientation
and initial volume fraction of voids were taken into account. The influence of the initial
void shape on the growth process is important at low levels of stress triaxiality. An increase
in stress triaxiality reduces the difference in volume growth of voids with different shapes.
For stress triaxiality above 2, almost the same increase in voids of different shapes was
recorded. At low triaxiality (0.33), the volume contribution of the spherical void increases
the most. The initial orientation of the voids has a strong influence on the growth process. If
there is an orientation of the voids with an area located normal to the direction of stretching,
greater void growth was observed. The initial volume proportion of voids showed less
influence on the nature of their growth compared to the other parameters analysed [124].

ȱ
Figure 15. Shapes of voids in the material determined using computed tomography techniques (for
structural steel) [124].

The influence of the shape of the void on the nature of its growth can be determined
using other advanced research methods. In [126], a research methodology was used
to complement the discrete dislocation plasticity (DDP) method with calculations using
XFEM [127,128]. Higher stress levels, strain hardening and void growth rates occurred
under biaxial loading (compared to uniaxial loading). With a constant initial proportion
of void volume, it was observed that elliptical-shaped voids showed larger surface areas
relative to cylindrical voids. The voids with a larger surface area in relation to volume
showed a tendency to grow faster, but with a lower proportion of strain hardening.
The Rice–Tracey void growth model with Huang [97] corrections was verified on the
basis of CT (compact tension) analysis. Specimens from three steels (single-phase ferritic
steel, two-phase steel and steel with martensitic microstructure) subjected to tension were
analysed. The analyses showed an increase in voids from a few μm to 30 μm. Attention
was drawn to the change in the initially spherical shapes of the voids and to the need to
account for the effect of the change in stress triaxiality in the Rice–Tracey model, depending
on the microstructure and characteristics of the material under study [129]. Relationships of
the initial shape of voids and the distance between voids to the level of fracture toughness
were developed [130], which are shown in Figure 16.

89
Materials 2022, 15, 6473

ȱ
ȱ
(a)ȱ (b)ȱ

Figure 16. Dependence of fracture toughness level on initial porosity: (a) Depending on the initial
shape of the void; (b) Depending on the anisotropic void spacing [130].

Researchers are also using advanced computed tomography techniques to verify


growth models and in situ void fusion: SRCT and SRCL [131]. The idea of measurements
according to these techniques is explained in [132,133]. The use of the SRCT technique
carries limitations, primarily with regard to the specimens used. Specimens must have
a cross-section size of approximately 1 mm for a resolution of the order of micrometres.
This has consequences in terms of the use of nonstandard specimens and the influence
of specimen geometry on the results obtained (influence of the presence of a plane stress
region and the development of a plastic zone). In such cases, it becomes helpful to use the
SRCL technique for in situ studies. This allows for a high-quality 3D image to be obtained
from specimens with larger dimensions than SRCT and the inclusion of areas on the sides
of the specimens. Using the SRCL technique combined with numerical calculations of
the CT (compact tension) specimen [131], parameters were determined in the GNT model
relating to the realisation of failure by void mechanism in the AA6061 aluminium alloy.
Accurate observation of the nucleation, growth and fusion of voids is possible through
the use of TEM (transmission electronic microscope) observations in experimental studies.
Compared to the SEM, the TEM technique allows for the dynamic growth of voids in the
material to be captured and for the void bonding process in particular to be observed at the
submicron scale level. In [134], the TEM technique was used to analyse the growth and
bonding of voids in different types of materials: copper and aluminium-copper alloy.

4.4. Consideration of the Stress State in the Realisation of the Ductile Failure Process
Analysis using numerical calculations (e.g., finite element method) helps understand
the growth process and the coalescence of voids in detail. Very often, an elementary cell
model (representative RVE volume model) is used, which in its structure contains a void
or particle with specific material characteristics. Classical studies consider the analysis of
cylindrical voids, spherical voids and spherical particles. Parameters that can be taken
into account in the RVE loading process are the stress triaxiality parameter (T), the Lode
parameter (L) and the shear coefficient (S). There has been much work involving studies of
the effect of stress triaxiality on the development on ductile failure, including a description
of the void growth and coalescence process. However, it has become an important task
to determine the simultaneous influence of the three aforementioned parameters of stress
triaxiality, Lode parameter and shear rate when analysed using RVE. A solution to the task
posed was proposed in [77]. The RVE model was used in [135] to extend Gurson’s proposal
to high porosity materials. The variables in the model were the volume proportion of voids
and the value of the stress triaxiality factor. It was shown that the distribution of plastic
deformation depends on the volume fraction of voids; a concentration of deformation
occurs for a fraction with a large volume of voids. A small distance between individual
voids leads to a faster coalescence process and fracture.

90
Materials 2022, 15, 6473

The contribution of the critical level of plastic strain, the stress triaxiality factor and
the Lode parameter to the realisation of the ductile fracture process has been highlighted
in a number of works by Wierzbicki and coworkers [136–138]. On the basis of numerous
experimental studies and numerical calculations, the dependence of the critical strain at
ductile fracture on the stress triaxiality was determined (Figure 17). Compression, tension
and shear specimens of aluminium alloy 2024-T351 characterised by various levels of
stress triaxiality factor from −0.33 to 1 were analysed. Based on the analyses, the authors
concluded that shear failure occurs for compression specimens with negative values of
stress triaxiality factor. For notched tensile specimens, the observed failure character
was dependent on the level of stress triaxiality. For high triaxiality there was a failure
mechanism involving initiation, growth and coalescence of voids. For low levels of stress
triaxiality, failure was a combination between two mechanisms: void and shear. Work
by Wierzbicki and colleagues and other researchers has highlighted the need to calibrate
the material relationship used in the numerical calculation programme, particularly when
analysing high-plasticity materials.

ȱ
Figure 17. Numerically and experimentally determined dependence between critical plastic strain
and the level of stress triaxiality factor (for 2024-T351 aluminium alloy) [139].

The influence of the Lode parameter value on the void growth and coalescence process
has formed the basis of a number of papers [92,137,140–144]. Barsoum and Faleskog pro-
posed a micromechanical model based on experimental studies and numerical calculations
of a three-dimensional elementary cell containing a single spherical void [145]. The macro-
scopic stress state was defined by two variable quantities: the stress triaxiality factor and
the Lode parameter. On the basis of the results obtained, it was determined that the effect
of the Lode parameter on the change in the shape of voids and the rate of their growth
increases as the level of the stress triaxiality factor decreases. For a stress triaxiality factor
of T = 1, there is an increase and a change in the shape of the void due to shear strain.
The void undergoes a change in shape from spherical to ellipsoidal, up to the onset of
plastic localisation, so as to further obtain a ‘penny’ shape. For triaxiality stress level 2,
there is an almost spherical increase in voids. The influence of the triaxiality parameter and
Lode on the formation of plastic localisation was also determined. High values of the T
parameter were accompanied by low levels of critical strain. The minimum value of critical
strain occurred at a triaxiality factor of 0, while the maximum value occurred at T = 1. This
marks the shear state as the most critical state from the point of view of destruction, where

91
Materials 2022, 15, 6473

material failure will occur at an angle close to 45◦ to the plane of occurrence of the lowest
principal stress [145].

5. Summary and Conclusions


Due to its great practical importance, the phenomenon of ductile failure associated
with the development of voids in metals has been the subject of interest of researchers
since the turn of the 1950s and 1960s. Over about 6 decades, many research techniques
have been developed that have contributed to a better understanding of the nature of this
phenomenon.
It was indicated that void formation in metals is induced by material discontinuities
on the nano- and microscopic scale. The void initiators mainly include (i) intersection
of the slip bands, (ii) grain boundaries, (iii) twin boundaries, (iv) vacancy clusters and
(v) second-phase particles (inclusions and precipitates).
The void formation mechanism depends primarily on the material purity. In the
absence of second-phase particles, voids are formed mainly in the vicinity of defects of
the crystal structure (point, linear and planar defects). Materials that are commonly used
in engineering (mainly steels and aluminium alloys) most often show the presence of
inclusions, which, as stress concentrators, are void initiators. Depending on the geometrical
and strength parameters of particles and matrix, voids are formed by particles’ separation
or fracture.
As has been demonstrated, the local values of the accompanying stresses, measured
on a microscopic scale, are much higher than those measured macroscopically, and range
from several hundred to several thousand MPa.
After the nucleation process, voids grow with increasing strain and change their
shape and position in the material. The moment of localisation of strain in the immediate
vicinity of voids is identified with the initiation of the coalescence process. Three ways of
realising the coalescence of voids in the material are characterised: internal necking, shear
coalescence and necklace coalescence. Selected analytical models relating to the growth
process and the coalescence of voids under ductile failure mechanism are cited in this
paper. An important element is the possibility of experimental verification of the classical
models. This verification is possible by using advanced measurement methods (computed
tomography, microstructure analysis, etc.) for in situ testing.
Despite enormous progress in this subject, there are still many unresolved issues.
The course and intensity of the void development are significantly influenced by the load
conditions, most importantly the stress state. However, the literature rarely mentions other
factors that undoubtedly determine the development of voids, e.g., dynamic phenomena,
temperature, etc.
It has not been unequivocally determined whether the stress or strain criteria must
be satisfied for the void nucleation. Moreover, the literature lacks unambiguous sets of
parameters describing the development of voids in materials with practical engineering
application. Little is known yet about the development of voids under low-triaxiality
conditions.
It should be strongly emphasised here that the observations of the development of
voids are not only of cognitive importance, but above all they become the basis for the
formulation of theoretical models, which further allows for the description of defective
materials. It is an issue of great practical importance, thanks to which it becomes possible
to perform an engineering safety assessment of structural elements containing defects,
operating in pre-failure conditions. In the future, an even more accurate understanding
of the ductile failure process according to the void formation mechanism will be possible
with the development of advanced testing and measurement methods. These can be used
for in situ research. An important aspect of supplementing the information obtained about
the void mechanism will be the inclusion of the results of advanced numerical calculations.

92
Materials 2022, 15, 6473

Author Contributions: Conceptualisation, W.W. and S.L.; investigation, W.W. and S.L.; resources,
W.W. and S.L.; data curation, W.W. and S.L.; writing—original draft preparation, W.W. and S.L.;
writing—review and editing, W.W. and S.L.; visualisation, W.W. and S.L.; supervision, W.W. and S.L.;
project administration, W.W. and S.L.; funding acquisition, W.W. and S.L. All authors have read and
agreed to the published version of the manuscript.
Funding: The APC was funded by Ministry of Science and Higher Education, grant number
01.0.08.00/1.02.001SUBB.MPKM.22.001.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: No new data were created or analysed in this study.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Benzerga, A.A.; Besson, J.; Pineau, A. Anisotropic Ductile Fracture: Part I: Experiments. Acta Mater. 2004, 52, 4623–4638.
[CrossRef]
2. Wcislik, W. Experimental Determination of Critical Void Volume Fraction FF for the Gurson Tvergaard Needleman (GTN) Model.
Proc. Struct. Integr. 2016, 2, 1676–1683. [CrossRef]
3. Tvergaard, V.; Needleman, A. Analysis of Cup-Cone Fracture in a Round Tensile Bar. Acta Metall. 1984, 32, 157–169. [CrossRef]
4. Plateau, J.; Henry, G.; Crussard, C. Quelques nouvelles applications de la microfractographie. Rev. Met. Paris 1957, 54, 200–216.
[CrossRef]
5. Schmitt, J.H.; Jalinier, J.M. Damage in Sheet Metal Forming—I: Physical Behavior. Acta Metall. 1982, 30, 1789–1798. [CrossRef]
6. Corbin, S.F.; Wilkinson, D.S. Influence of Matrix Strength and Damage Accumulation on the Mechanical Response of a Particulate
Metal Matrix Composite. Acta Metall. Mater. 1994, 42, 1329–1335. [CrossRef]
7. Pardoen, T.; Delannay, F. Assessment of Void Growth Models from Porosity Measurements in Cold-Drawn Copper Bars. Metall.
Mater. Trans. A 1998, 29, 1895–1909. [CrossRef]
8. El Guerjouma, R.; Baboux, J.-C.; Ducret, D.; Godin, N.; Guy, P.; Huguet, S.; Jayet, Y.; Monnier, T. Non-Destructive Evaluation of
Damage and Failure of Fibre Reinforced Polymer Composites Using Ultrasonic Waves and Acoustic Emission. Adv. Eng. Mater.
2001, 3, 601–608. [CrossRef]
9. Landron, C.; Maire, E.; Adrien, J.; Bouaziz, O. Damage characterization in dual-phase steels using x-ray tomography. In The
Optical Measurements, Modeling, and Metrology; Proulx, T., Ed.; Springer: New York, NY, USA, 2011; Volume 5, pp. 11–18.
10. Horstemeyer, M.F.; Gokhale, A.M. A Void–Crack Nucleation Model for Ductile Metals. Int. J. Solids Struct. 1999, 36, 5029–5055.
[CrossRef]
11. Chan, K.S.; Davidson, D.L. Evidence of Void Nucleation and Growth on Planar Slip Bands in a Nb-Cr-Ti Alloy. Metall. Mater.
Trans. A 1999, 30, 579–585. [CrossRef]
12. Gardner, R.N.; Pollock, T.C.; Wilsdorf, H.G.F. Crack Initiation at Dislocation Cell Boundaries in the Ductile Fracture of Metals.
Mater. Sci. Eng. 1977, 29, 169–174. [CrossRef]
13. Nowak, Z. The Method of Identification in Mechanics of Ductile Materials with Defects; IFTR Reports; Institute of Fundamental
Technological Research Polish Academy of Sciences: Warsaw, Poland, 2006.
14. Crussard, C.; Plateau, J.; Tamhankar, R.; Henry, G.; Lajeunesse, D. A comparison of ductile and fatigue fractures. In Proceedings
of the Fracture, Swampscott, MA, USA, 12–16 April 1959; pp. 524–561.
15. Puttick, K.E. Ductile Fracture in Metals. Philos. Mag. A J. Theor. Exp. Appl. Phys. 1959, 4, 964–969. [CrossRef]
16. Rogers, H.C. The Tensile Fracture of Ductile Metals. Trans. Metall. Soc. AIME 1960, 218, 498–506.
17. Orowan, E. Fracture and Strength of Solids. Rep. Prog. Phys. 1948, XII, 185–232. [CrossRef]
18. Koppenaal, T.J. Porosity in Plastically Deformed Cu-10 at.% Al Single Crystals. Acta Metall. 1961, 9, 1078–1079. [CrossRef]
19. Noell, P.; Carroll, J.; Hattar, K.; Clark, B.; Boyce, B. Do Voids Nucleate at Grain Boundaries during Ductile Rupture? Acta Mater.
2017, 137, 103–114. [CrossRef]
20. Noell, P.J.; Sabisch, J.E.C.; Medlin, D.L.; Boyce, B.L. Nanoscale Conditions for Ductile Void Nucleation in Copper: Vacancy
Condensation and the Growth-Limited Microstructural State. Acta Mater. 2020, 184, 211–224. [CrossRef]
21. Pardoen, T.; Dumont, D.; Deschamps, A.; Brechet, Y. Grain Boundary versus Transgranular Ductile Failure. J. Mech. Phys. Solids
2003, 51, 637–665. [CrossRef]
22. Ashby, M.F.; Gandhi, C.; Taplin, D.M.R. Overview No. 3 Fracture-Mechanism Maps and Their Construction for f.c.c. Metals and
Alloys. Acta Metall. 1979, 27, 699–729. [CrossRef]
23. Henseler, T.; Osovski, S.; Ullmann, M.; Kawalla, R.; Prahl, U. GTN Model-Based Material Parameters of AZ31 Magnesium Sheet
at Various Temperatures by Means of SEM In-Situ Testing. Crystals 2020, 10, 856. [CrossRef]
24. Pineau, A.; Benzerga, A.A.; Pardoen, T. Failure of Metals I: Brittle and Ductile Fracture. Acta Mater. 2016, 107, 424–483. [CrossRef]
25. Wciślik, W.; Pała, R. Some Microstructural Aspects of Ductile Fracture of Metals. Materials 2021, 14, 4321. [CrossRef]

93
Materials 2022, 15, 6473

26. Cox, T.B.; Low, J.R. An Investigation of the Plastic Fracture of AISI 4340 and 18 Nickel-200 Grade Maraging Steels. Metall. Mater.
Trans. B 1974, 5, 1457–1470. [CrossRef]
27. Tiryakioğlu, M. Intrinsic and Extrinsic Effects of Microstructure on Properties in Cast Al Alloys. Materials 2020, 13, 2019. [CrossRef]
[PubMed]
28. Gangulee, A.; Gurland, J. On the Fracture of Silicon Particles in Aluminum-Silicon Alloys. Trans. Metall. Soc. AIME 1967, 239,
269–272.
29. Finlayson, T.; Griffths, J.; Viano, D.; Fitzpatrick, M.; Oliver, E.; Wang, Q. In stresses in the eutectic silicon particles of strontium-
modified A356 castings loaded in tension. In Shape Casting: 2nd International Symposium; The Minerals, Metals & Materials Society:
Warrendale, PA, USA, 2007; pp. 127–134.
30. Caceres, C.H.; Griffiths, J.R. Damage by the Cracking of Silicon Particles in an Al-7Si-0.4 Mg Casting Alloy. Acta Mater. 1996, 44,
25–33. [CrossRef]
31. Yeh, J.-W.; Liu, W.-P. The Cracking Mechanism of Silicon Particles in an A357 Aluminum Alloy. Metall. Mater. Trans. A 1996, 27,
3558–3568. [CrossRef]
32. Poole, W.J.; Charras, N. An Experimental Study on the Effect of Damage on the Stress–Strain Behaviour for Al–Si Model
Composites. Mater. Sci. Eng. A 2005, 406, 300–308. [CrossRef]
33. Argon, A.S.; Im, J.; Safoglu, R. Cavity Formation from Inclusions in Ductile Fracture. Metall. Mater. Trans. A 1975, 6, 825.
[CrossRef]
34. Beremin, F.M. Cavity Formation from Inclusions in Ductile Fracture of A508 Steel. Metall. Mater. Trans. A 1981, 12a, 723–731.
[CrossRef]
35. Coade, R.W.; Griffiths, J.R.; Parker, B.A.; Stevens, P.J. Inclusion Stresses in a Two-Phase Alloy Deformed to a Plastic Strain of 1%.
Philos. Mag. A 1981, 44, 357–372. [CrossRef]
36. Mueller, M.G.; Žagar, G.; Mortensen, A. In-Situ Strength of Individual Silicon Particles within an Aluminium Casting Alloy. Acta
Mater. 2018, 143, 67–76. [CrossRef]
37. Mueller, M.; Fornabaio, M.; Žagar, G.; Mortensen, A. Microscopic Strength of Silicon Particles in an Aluminium–Silicon Alloy.
Acta Mater. 2016, 105, 165–175. [CrossRef]
38. Grodecki, K.; Bozek, R.; Strupinski, W.; Wysmolek, A.; Stepniewski, R. Micro-Raman Spectroscopy of Graphene Grown on
Stepped 4H-SiC (0001) Surface. Appl. Phys. Lett. 2009, 100, 261604. [CrossRef]
39. Harris, S.J.; O’Neill, A.E.; Yang, W.; Gustafson, P.; Boileau, J.; Weber, W.H.; Majumdar, B.; Ghosh, S. Measurement of the State of
Stress in Silicon with Micro-Raman Spectroscopy. J. Appl. Phys. 2004, 96, 7195–7201. [CrossRef]
40. Harris, S.; O’Neill, A.; Boileau, J.; Donlon, W.; Su, X.; Majumdar, B. Application of the Raman Technique to Measure Stress States
in Individual Si Particles in a Cast Al–Si Alloy. Acta Mater. 2007, 55, 1681–1693. [CrossRef]
41. Ward, D.K.; Curtin, W.A.; Qi, Y. Aluminum-Silicon Interfaces and Nanocomposites: A Molecular Dynamics Study. Compos. Sci.
Technol. 2006, 66, 1151–1161. [CrossRef]
42. Ward, D.K.; Curtin, W.A.; Qi, Y. Mechanical Behavior of Aluminum-Silicon Nanocomposites: A Molecular Dynamics Study. Acta
Mater. 2006, 54, 4441–4451. [CrossRef]
43. Noreyan, A.; Qi, Y.; Stoilov, V. Critical Shear Stresses at Aluminum–Silicon Interfaces. Acta Mater. 2008, 56, 3461–3469. [CrossRef]
44. Needleman, A. Some Issues in Cohesive Surface Modeling. Procedia IUTAM 2014, 10, 221–246. [CrossRef]
45. Galkiewcz, J. Cohesive Model Application to Micro-Crack Nucleation and Growth. Procedia Struct. Integr. 2016, 2, 1619–1626.
[CrossRef]
46. Gałkiewicz, J. Microscopically Based Calibration of the Cohesive Model. J. Theor. Appl. Mech. 2015, 53, 477–485. [CrossRef]
47. Kossakowski, P.; Wciślik, W. The effect of stress state triaxiality on the value of microvoid nucleation strain in S235JR steel.
˛ Mech. 2013, 3, 15–21.
Przeglad
48. Alharbi, K.; Ghadbeigi, H.; Efthymiadis, P.; Zanganeh, M.; Celotto, S.; Dashwood, R.; Pinna, C. Damage in Dual Phase Steel
DP1000 Investigated Using Digital Image Correlation and Microstructure Simulation. Model. Simul. Mater. Sci. Eng. 2015, 23,
085005. [CrossRef]
49. He, X.J.; Terao, N.; Berghezan, A. Influence of Martensite Morphology and Its Dispersion on Mechanical Properties and Fracture
Mechanisms of Fe-Mn-C Dual Phase Steels. Met. Sci. 1984, 18, 367–373. [CrossRef]
50. Avramovic-Cingara, G.; Saleh, C.A.R.; Jain, M.K.; Wilkinson, D. Void Nucleation and Growth in Dual-Phase Steel 600 during
Uniaxial Tensile Testing. Metall. Mater. Trans. A 2009, 40, 3117–3127. [CrossRef]
51. Szewczyk, A.F.; Gurland, J. A Study of the Deformation and Fracture of a Dual-Phase Steel. Metall. Mater. Trans. A 1982, 13,
1821–1826. [CrossRef]
52. Santos, R.O.; da Silveira, L.B.; Moreira, L.P.; Cardoso, M.C.; da Silva, F.R.F.; dos Santos Paula, A.; Albertacci, D.A. Damage
Identification Parameters of Dual-Phase 600–800 Steels Based on Experimental Void Analysis and Finite Element Simulations. J.
Mater. Res. Technol. 2019, 8, 644–659. [CrossRef]
53. Wu, S.C.; Xiao, T.Q.; Withers, P.J. The Imaging of Failure in Structural Materials by Synchrotron Radiation X-Ray Microtomography.
Eng. Fract. Mech. 2017, 182, 127–156. [CrossRef]
54. Seo, D.; Toda, H.; Kobayashi, M.; Uesugi, K.; Takeuchi, A.; Suzuki, Y. In Situ Observation of Void Nucleation and Growth in a
Steel Using X-Ray Tomography. ISIJ Int. 2015, 55, 1474–1482. [CrossRef]

94
Materials 2022, 15, 6473

55. Kobayashi, M.; Toda, H.; Kawai, Y.; Ohgaki, T.; Uesugi, K.; Wilkinson, D.; Kobayashi, T.; Aoki, Y.; Nakazawa, M. High-Density
Three-Dimensional Mapping of Internal Strain by Tracking Microstructural Features. Acta Mater. 2008, 56, 2167–2181. [CrossRef]
56. Toda, H.; Maire, E.; Aoki, Y.; Kobayashi, M. Three-Dimensional Strain Mapping Using in Situ X-Ray Synchrotron Microtomogra-
phy. J. Strain Anal. Eng. Des. 2011, 46, 549–561. [CrossRef]
57. Guo, Y.; Burnett, T.L.; McDonald, S.A.; Daly, M.; Sherry, A.H.; Withers, P.J. 4D Imaging of Void Nucleation, Growth, and
Coalescence from Large and Small Inclusions in Steel under Tensile Deformation. J. Mater. Sci. Technol. 2022, 123, 168–176.
[CrossRef]
58. Pathak, N.; Adrien, J.; Butcher, C.; Maire, E.; Worswick, M. Experimental Stress State-Dependent Void Nucleation Behavior for
Advanced High Strength Steels. Int. J. Mech. Sci. 2020, 179, 105661. [CrossRef]
59. Wcislik, W. Experimental and Numerical Determination and Analysis of Selected Parameters of the Gurson-Tvergaard-Needleman
Model for S355 Steel and Complex Stress States. Ph.D. Thesis, Kielce University of Technology, Kielce, Poland, 2014.
60. Testa, G.; Bonora, N.; Ruggiero, A.; Iannitti, G.; Gentile, D. Stress Triaxiality Effect on Void Nucleation in Ductile Metals. Fatigue
Fract. Eng. Mater.Struct. 2020, 43, 1473–1486. [CrossRef]
61. Yu, Q. Influence of the Stress State on Void Nucleation and Subsequent Growth around Inclusion in Ductile Material. Int. J. Fract.
2015, 193, 43–57. [CrossRef]
62. Han, S.; Chang, Y.; Wang, C.; Han, Y.; Dong, H. Experimental and Numerical Investigations on the Damage Induced in the
Shearing Process for QP980 Steel. Materials 2022, 15, 3254. [CrossRef]
63. Tancogne-Dejean, T.; Roth, C.; Morgeneyer, T.; Helfen, L.; Mohr, D. Ductile Damage of AA2024-T3 under Shear Loading:
Mechanism Analysis Through In-Situ Laminography. Acta Mater. 2020, 205, 116556. [CrossRef]
64. Shakoor, M.; Bernacki, M.; Bouchard, P.-O. Ductile Fracture of a Metal Matrix Composite Studied Using 3D Numerical Modeling
of Void Nucleation and Coalescence. Eng. Fract. Mech. 2017, 189, 110–132. [CrossRef]
65. Nguyen, T.D.; Plimpton, S.J. Aspherical Particle Models for Molecular Dynamics Simulation. Comput. Phys. Commun. 2019, 243,
12–24. [CrossRef]
66. Lucchetta, A.; Brach, S.; Kondo, D. Effects of Particles Size on the Overall Strength of Nanocomposites: Molecular Dynamics
Simulations and Theoretical Modeling. Mech. Res. Commun. 2021, 114, 103669. [CrossRef]
67. Zhao, Q.Q.; Boyce, B.L.; Sills, R.B. Micromechanics of Void Nucleation and Early Growth at Incoherent Precipitates: Lattice-
Trapped and Dislocation-Mediated Delamination Modes. Crystals 2021, 11, 45. [CrossRef]
68. Anderson, T.L. Fracture mechanics. In Fundamentals and Applications, 3rd ed.; Taylor and Francis Group: London, UK, 2005.
69. Dzioba, I. Modelowanie i Analiza Procesu P˛ekania w Stalach Ferrytycznych; Politechnika Świ˛etokrzyska: Kielce, Poland, 2012.
70. Fressengeas, C.; Molinari, A. Inertia and Thermal Effects on the Localization of Plastic Flow. Acta Metall. 1985, 33, 387–396.
[CrossRef]
71. Pardoen, T.; Hutchinson, J.W. An Extended Model for Void Growth and Coalescence. J. Mech. Phys. Solids 2000, 48, 2467–2512.
[CrossRef]
72. Mercier, S.; Molinari, A. Predictions of Bifurcation and Instabilities during Dynamic Extension. Int. J. Solids Struct. 2003, 40,
1995–2016. [CrossRef]
73. Aretz, H. Numerical Analysis of Diffuse and Localized Necking in Orthotropic Sheet Metals. Int. J. Plast. 2007, 23, 798–840.
[CrossRef]
74. Barsoum, I. The Effect of Stress State in Ductile Failure. Ph.D. Thesis, Royal Institute of Technology, Stockholm, Sweden, 2008.
75. Needleman, A.; Rice, J.R. Limits to ductility set by plastic flow localization. In Mechanics of Sheet Metal Forming: Material
Behavior and Deformation Analysis; Koistinen, D.P., Wang, N.-M., Eds.; Springer US: Boston, MA, USA, 1978; pp. 237–267, ISBN
978-1-4613-2880-3.
76. Reboul, J.; Srivastava, A.; Osovski, S.; Vadillo, G. Influence of Strain Rate Sensitivity on Localization and Void Coalescence. Int. J.
Plast. 2020, 125, 265–279. [CrossRef]
77. Tekoglu, C.; Leblond, J.-B.; Pardoen, T. A Criterion for the Onset of Void Coalescence under Combined Tension and Shear. J. Mech.
Phys. Solids 2012, 60, 1363–1381. [CrossRef]
78. Wang, J.; Liang, J.; Wen, Z.; Yue, Z. Void Configuration-Induced Change in Microstructure and Deformation Mechanisms of
Nano-Porous Materials. J. Appl. Phys. 2019, 126, 085106. [CrossRef]
79. Besson, J. Continuum Models of Ductile Fracture: A Review. Int. J. Damage Mech. 2010, 19, 3–52. [CrossRef]
80. Weck, A.; Wilkinson, D.S.; Maire, E.; Toda, H. Visualization by X-Ray Tomography of Void Growth and Coalescence Leading to
Fracture in Model Materials. Acta Mater. 2008, 56, 2919–2928. [CrossRef]
81. Benzerga, A.A. Micromechanics of Coalescence in Ductile Fracture. J. Mech. Phys. Solids 2002, 50, 1331–1362. [CrossRef]
82. Scheyvaerts, F.; Pardoen, T.; Onck, P.R. A New Model for Void Coalescence by Internal Necking. Int. J. Damage Mech. 2010, 19,
95–126. [CrossRef]
83. Gologanu, M.; Leblond, J.-B.; Perrin, G.; Devaux, J. Theoretical Models for Void Coalescence in Porous Ductile Solids: I—
Coalescence “in Layers”. Int. J. Solids Struct. 2001, 38, 5581–5594. [CrossRef]
84. Morin, L.; Leblond, J.-B.; Benzerga, A.A. Coalescence of Voids by Internal Necking: Theoretical Estimates and Numerical Results.
J. Mech. Phys. Solids 2015, 75, 140–158. [CrossRef]
85. Koplik, J.; Needleman, A. Void Growth and Coalescence in Porous Plastic Solids. Int. J. Solids Struct. 1988, 24, 835–853. [CrossRef]

95
Materials 2022, 15, 6473

86. Richelsen, A.B.; Tvergaard, V. Dilatant Plasticity or Upper Bound Estimates for Porous Ductile Solids. Acta Metall. Mater. 1994, 42,
2561–2577. [CrossRef]
87. Srivastava, A.; Needleman, A. Void Growth versus Void Collapse in a Creeping Single Crystal. J. Mech. Phys. Solids 2013, 61,
1169–1184. [CrossRef]
88. Bandstra, J.; Koss, D.; Geltmacher, A.; Matic, P.; Everett, R. Modeling Void Coalescence during Ductile Fracture of a Steel. Mater.
Sci. Eng. A 2004, 366, 269–281. [CrossRef]
89. Chang, Z.; Chen, J. A New Void Coalescence Mechanism during Incremental Sheet Forming: Ductile Fracture Modeling and
Experimental Validation. J. Mater. Process. Technol. 2021, 298, 117319. [CrossRef]
90. Gologanu, M.; Leblond, J.-B.; Devaux, J. Theoretical Models for Void Coalescence in Porous Ductile Solids: II—Coalescence “in
Columns”. Int. J. Solids Struct. 2001, 38, 5595–5604. [CrossRef]
91. Pala, R.; Dzioba, I. Influence of Delamination on the Parameters of Triaxial State of Stress before the Front of the Main Crack. AIP
Conf. Proc. 2018, 2029, 020052. [CrossRef]
92. Ganjiani, M.; Homayounfard, M. Development of a Ductile Failure Model Sensitive to Stress Triaxiality and Lode Angle. Int. J.
Solids Struct. 2021, 225, 111066. [CrossRef]
93. McClintock, F.A. A Criterion for Ductile Fracture by Growth of Hole. J. Appl. Mech. 1968, 35, 353–371. [CrossRef]
94. Rice, J.R.; Tracey, D.M. On the Ductile Enlargement of Voids in Triaxial Stress Fields. J. Mech. Phys. Solids 1968, 17, 201–217.
[CrossRef]
95. Budianski, B.; Hutchinson, J.W.; Slutski, S. Void Growth and Collapse in Viscous Solids; Mechanics of Solids; Pergamon Press: Oxford,
UK, 1982.
96. Becker, R.; Smelser, R.E.; Richmond, O. The Effect of Void Shape on the Development of Damage and Fracture in Plane Strain
Tension. J. Mech. Phys. Solids 1989, 37, 111–129. [CrossRef]
97. Huang, Y. Accurate Dilatation Rates for Spherical Voids in Triaxial Stress Fields. J. Appl. Mech. 1991, 58, 1084–1086. [CrossRef]
98. Gurson, A.L. Continuum Theory of Ductile Rupture by Void Nucleation and Growth: Part 1—Yield Criteria and Flow Rules for
Porous Ductile Media. J. Eng. Mater. Technol. 1977, 99, 2–15. [CrossRef]
99. Berdin, C.; Hausild, P. Damage mechanisms and local approach to fracture part I: Ductile fracture. In Transferability of Fracture
Mechanical Characteristics; NATO Science Series: Brno, Czech Republic, 2001.
100. Gologanu, M.; Leblond, J.B.; Devaux, J. Approximate Models for Ductile Metals Containing Nonspherical Voids—Case of
Axisymmetric Prolate Ellipsoidal Cavities. J. Mech. Phys. Solids 1993, 41, 1723–1754. [CrossRef]
101. Gologanu, M.; Leblond, J.B.; Devaux, J. Approximate Models for Ductile Metals Containing Nonspherical Voids—Case of
Axisymmetric Oblate Ellipsoidal Cavities. J. Eng. Mater. Technol. 1994, 116, 290–297. [CrossRef]
102. Gologanu, M.; Leblond, J.B.; Perrin, G.; Devaux, J. Recent Extensions of Gursons Model for Porous Ductile Metals; Continuum
Micromechanics; Springer: Berlin/Heidelberg, Germany, 1995.
103. Gao, X.; Wang, T.; Kim, J. On Ductile Fracture Initiation Toughness: Effects of Void Volume Fraction, Void Shape and Void
Distribution. Int. J. Solids Struct. 2005, 42, 5097–5117. [CrossRef]
104. Gao, X.; Kim, J. Modeling of Ductile Fracture: Significance of Void Coalescence. Int. J. Solids Struct. 2006, 43, 6277–6293. [CrossRef]
105. Zhang, Z.; Skallerud, B. Void Coalescence with and without Prestrain History. Int. J. Damage Mech. 2010, 19, 153–174. [CrossRef]
106. Besson, J. Local Approach to Fracture; Ecole des Mines de Paris: Paris, France, 2004.
107. Brown, L.M.; Embury, J.D. The initiation and growth of voids at second phase particles. In Proceedings of the Third International
Conference on Strength of Metals and Alloys, ICSMA 3, Cambridge, UK, 20–25 August 1973; Institute of Metals: Cambridge, UK,
1973; pp. 164–169.
108. Thomason, P.F. A Three-Dimensional Model for Ductile Fracture by the Growth and Coalescence of Microvoids. Acta Metall. 1985,
33, 1087–1095. [CrossRef]
109. Noell, P.J.; Carroll, J.D.; Boyce, B.L. The Mechanisms of Ductile Rupture. Acta Mater. 2018, 161, 83–98. [CrossRef]
110. Dzioba, I.; Lipiec, S.; Pała, R.; Furmańczyk, P. On Characteristics of Ferritic Steel Determined during the Uniaxial Tensile Test.
Materials 2021, 14, 3117. [CrossRef]
111. Babout, L.; Maire, E.; Buffière, J.Y.; Fougères, R. Characterization by X-Ray Computed Tomography of Decohesion, Porosity
Growth and Coalescence in Model Metal Matrix Composites. Acta Mater. 2001, 49, 2055–2063. [CrossRef]
112. Gammage, J.; Wilkinson, D.; Brechet, Y.; Embury, D. A Model for Damage Coalescence in Heterogeneous Multi-Phase Materials.
Acta Mater. 2004, 52, 5255–5263. [CrossRef]
113. Jia, S.; Povirk, G.L. Modeling the Effects of Hole Distribution in Perforated Aluminum Sheets II: Minimum Strength Failure Paths.
Int. J. Solids Struct. 2002, 39, 2533–2545. [CrossRef]
114. Nagaki, S.; Nakayama, Y.; Abe, T. Relation between Damage Due to Circular Holes and Local Deformation of Perforated Sheets.
Int. J. Mech. Sci. 1998, 40, 215–226. [CrossRef]
115. Cao, T.-S.; Mazière, M.; Danas, K.; Besson, J. A Model for Ductile Damage Prediction at Low Stress Triaxialities Incorporating
Void Shape Change and Void Rotation. Int. J. Solids Struct. 2015, 63, 240–263. [CrossRef]
116. Komori, K. Improvement and Validation of the Ellipsoidal Void Model for Predicting Ductile Fracture. Proc. Eng. 2017, 207,
2036–2041. [CrossRef]
117. Shen, Y.; Ma, T.; Li, J. A Damage Evolution Model Considering Void Shape Effect and New Damage Level Definition. Eng. Fract.
Mech. 2022, 267, 108461. [CrossRef]

96
Materials 2022, 15, 6473

118. Achouri, M.; Germain, G.; Dal Santo, P.; Saidane, D. Experimental Characterization and Numerical Modeling of Micromechanical
Damage under Different Stress States. Mater. Des. 2013, 50, 207–222. [CrossRef]
119. Azghandi, S.H.M.; Weiss, M.; Arhatari, B.D.; Adrien, J.; Maire, E.; Barnett, M.R. A Rationale for the Influence of Grain Size on
Failure of Magnesium Alloy AZ31: An in Situ X-Ray Microtomography Study. Acta Mater. 2020, 200, 619–631. [CrossRef]
120. Croom, B.P.; Jin, H.; Noell, P.J.; Boyce, B.L.; Li, X. Collaborative Ductile Rupture Mechanisms of High-Purity Copper Identified by
in Situ X-Ray Computed Tomography. Acta Mater. 2019, 181, 377–384. [CrossRef]
121. Fabrègue, D.; Landron, C.; Bouaziz, O.; Maire, E. Damage Evolution in TWIP and Standard Austenitic Steel by Means of 3D X
Ray Tomography. Mater. Sci. Eng. A 2013, 579, 92–98. [CrossRef]
122. Nemcko, M.J.; Wilkinson, D.S. On the Damage and Fracture of Commercially Pure Magnesium Using X-Ray Microtomography.
Mater. Sci. Eng. A 2016, 676, 146–155. [CrossRef]
123. Song, Q.-Y.; Heidarpour, A.; Zhao, X.-L.; Han, L.-H. Experimental and Numerical Investigation of Ductile Fracture of Carbon
Steel Structural Components. J. Constr. Steel Res. 2018, 145, 425–437. [CrossRef]
124. Xie, J.; Zhang, R.; Liu, T.; Zhou, C.; Jia, L.J. Effect of Initial Void Shape on Void Growth of Structural Steels Based on Micromechan-
ical RVE Models. J. Mater. Civil Eng. 2022, 34, 04022010. [CrossRef]
125. Benzerga, A.; Leblond, J.-B. Ductile Fracture by Void Growth to Coalescence. Adv. Appl. Mech. 2010, 44, 169–305. [CrossRef]
126. Usman, M.; Waheed, S.; Mubashar, A. Effect of Shape on Void Growth: A Coupled Extended Finite Element Method (XFEM) and
Discrete Dislocation Plasticity (DDP) Study. Eur. J. Mech.-A Solids 2022, 92, 104471. [CrossRef]
127. Romero, I.; Segurado, J.; LLorca, J. Dislocation Dynamics in Non-Convex Domains Using Finite Elements with Embedded
Discontinuities. Model. Simul. Mater. Sci. Eng. 2008, 16, 035008. [CrossRef]
128. Liang, S.; Zhu, Y.; Huang, M.; Li, Z. Simulation on Crack Propagation vs. Crack-Tip Dislocation Emission by XFEM-Based DDD
Scheme. Int. J. Plast. 2019, 114, 87–105. [CrossRef]
129. Landron, C.; Maire, E.; Bouaziz, O.; Adrien, J.; Lecarme, L.; Bareggi, A. Validation of Void Growth Models Using X-Ray
Microtomography Characterization of Damage in Dual Phase Steels. Acta Mater. 2011, 59, 7564–7573. [CrossRef]
130. Pardoen, T.; Hutchinson, J.W. Micromechanics-Based Model for Trends in Toughness of Ductile Metals. Acta Mater. 2003, 51,
133–148. [CrossRef]
131. Shen, Y.; Morgeneyer, T.F.; Garnier, J.; Allais, L.; Helfen, L.; Crépin, J. Three-Dimensional Quantitative in Situ Study of Crack
Initiation and Propagation in AA6061 Aluminum Alloy Sheets via Synchrotron Laminography and Finite-Element Simulations.
Acta Mater. 2013, 61, 2571–2582. [CrossRef]
132. Helfen, L.; Myagotin, A.; Rack, A.; Pernot, P.; Mikulík, P.; Di Michiel, M.; Baumbach, T. Synchrotron-radiation Computed
Laminography for High-resolution Three-dimensional Imaging of Flat Devices. Phys. Status Solidi A 2007, 204, 2760–2765.
[CrossRef]
133. Bull, D.J.; Helfen, L.; Sinclair, I.; Spearing, S.M.; Baumbach, T. A Comparison of Multi-Scale 3D X-Ray Tomographic Inspection
Techniques for Assessing Carbon Fibre Composite Impact Damage. Compos. Sci. Technol. 2013, 75, 55–61. [CrossRef]
134. Gao, B.; Xiang, Q.; Guo, T.; Guo, X.; Tang, S.; Huang, X.X. In Situ TEM Investigation on Void Coalescence in Metallic Materials.
Mater. Sci. Eng. A 2018, 734, 260–268. [CrossRef]
135. Bensaada, R.; Kanit, T.; Imad, A.; Almansba, M.; Saouab, A. Void-Growth Computational Analysis in Elastic-Plastic Porous
Materials. Int. J. Mech. Sci. 2022, 217, 107021. [CrossRef]
136. Wierzbicki, T.; Bao, Y.; Lee, Y.-W.; Bai, Y. Calibration of Seven Fracture Models. Int. J. Mech.Sci. 2005, 47, 719–743. [CrossRef]
137. Bai, Y.; Wierzbicki, T. A New Model of Metal Plasticity and Fracture with Pressure and Lode Dependence. Int. J. Plast. 2008, 24,
1071–1096. [CrossRef]
138. Bai, Y.; Wierzbicki, T. Application of Extended Mohr–Coulomb Criterion to Ductile Fracture. Int. J. Fract. 2010, 161, 1. [CrossRef]
139. Bao, Y.; Wierzbicki, T. On Fracture Locus in the Equivalent Strain and Stress Triaxiality Space. Int. J. Mech. Sci. 2004, 46, 81–98.
[CrossRef]
140. Zhang, K.S.; Bai, J.B.; François, D. Numerical Analysis of the Influence of the Lode Parameter on Void Growth. Int. J. Solids Struct.
2001, 38, 5847–5856. [CrossRef]
141. Liu, X.; Yan, S.; Rasmussen, K.J.R.; Deierlein, G.G. Experimental Investigation of the Effect of Lode Angle on Fracture Initiation of
Steels. Eng. Fract. Mech. 2022, 271, 108637. [CrossRef]
142. Danas, K.; Ponte Castañeda, P. Influence of the Lode Parameter and the Stress Triaxiality on the Failure of Elasto-Plastic Porous
Materials. Int. J. Solids Struct. 2012, 49, 1325–1342. [CrossRef]
143. Dunand, M.; Mohr, D. Effect of Lode Parameter on Plastic Flow Localization after Proportional Loading at Low Stress Triaxialities.
J. Mech. Phys. Solids 2014, 66, 133–153. [CrossRef]
144. Bonora, N.; Testa, G. Plasticity Damage Self-Consistent Model Incorporating Stress Triaxiality and Shear Controlled Fracture
Mechanisms—Model Formulation. Eng. Fract. Mech. 2022, 271, 108634. [CrossRef]
145. Barsoum, I.; Faleskog, J. Micromechanical Analysis on the Influence of the Lode Parameter on Void Growth and Coalescence. Int.
J. Solids Struct. 2011, 48, 925–938. [CrossRef]

97
materials
Article
Mechanical Properties Analysis of Explosive Welded Sheet of
AA2519-Ti6Al4V with Interlayer of AA1050 Subjected
to Heat-Treatment
Ireneusz Szachogłuchowicz *, Lucjan Śnieżek and Tomasz Śl˛ezak

Faculty of Mechanical Engineering, Institute of Robots & Machine Design, Military University of Technology,
ul. gen. S. Kaliskiego 2, 00-908 Warsaw, Poland; [email protected] (L.Ś.); [email protected] (T.Ś.)
* Correspondence: [email protected]

Abstract: The paper presents results of investigations of welding sheets of AA2519-Ti6Al4V, a difficult-
to-joint components materials, produced by explosive welding with a thin technological interlayer
of AA1050. The joining process leads to the formation of intermetalics in the vicinity of joint and
generates significant residual stresses. In the next step the laminate was subjected to a heat treatment
process in order to improve the mechanical properties by precipitation hardening. This treatment
should not be carried out before welding because of negative influence on a ductility of the aluminum
alloy. Material in this state was subjected to the tests of chemical composition, microstructure, and
microhardness. A tensile test was carried out with accompanying strain analysis by the digital image
correlation (DIC) method. Moreover, the residual stresses were determined which were measured
by using two methods, the X-ray diffraction and the hole drilling. This approach made it possible
to measure the residual stresses both in the plane parallel to the surface and in the cross section of
the laminate.

Citation: Szachogłuchowicz, I.;


Keywords: explosive welding; laminate; AA2519; Ti6Al4V; heat treatment; residual stress measurement
Śnieżek, L.; Śl˛ezak, T. Mechanical
Properties Analysis of Explosive
Welded Sheet of AA2519-Ti6Al4V
with Interlayer of AA1050 Subjected 1. Introduction
to Heat-Treatment. Materials 2022, 15, Explosive welding technology enable to join the materials difficult-to-join by conven-
4023. https://doi.org/10.3390/ tional methods and it is possible to be realized on large plates also. This process is realized
ma15114023 by explosively generated pressure wave that imparts very high velocities to the solids
Academic Editor: Michele Bacciocchi (welded materials). During collision the pressure reaches up to 2 × 104 MPa. Such level of
pressure makes it possible to obtain physical states unattainable under conventional types
Received: 12 May 2022
of loads [1,2]. As a result, significant residual stresses appear, deviating from the primary
Accepted: 2 June 2022
distribution in the combined materials [3,4].
Published: 6 June 2022
AA2519 alloys andTi6Al4V alloys are characterized by increased ballistic resistance.
Publisher’s Note: MDPI stays neutral These alloys are used in aerospace industry. The combination of these materials using
with regard to jurisdictional claims in an explosive welding method should increase ballistic resistance by changing the density
published maps and institutional affil- centers of the material and the appearance of intermetallics in the joint layer. In the
iations. welding zone of Al-Ti explosively produced laminates, the intermetallic phases Al3 Ti and
Ti3 Al were revealed [5–7] which may promote delamination of the produced material
and the decrease of strength properties. Additionally, significant residual stresses are
present. Therefore, proper heat treatment should be carried out to decrease the negative
Copyright: © 2022 by the authors.
influence of pointed factors, especially that, the heat treatment is recommended to reduce
Licensee MDPI, Basel, Switzerland.
the adverse effects of the residual stresses formed in the material as a result of the explosive
This article is an open access article
welding process [8]. The most important in such a case is the selection of appropriate
distributed under the terms and
processing parameters [9]. In considered laminate, among all components the AA2519
conditions of the Creative Commons
Attribution (CC BY) license (https://
alloy is characterized by the lowest melting temperature, which amounts 821 K. Carefully
creativecommons.org/licenses/by/
selected heat treatment parameters, due to the addition of Cu, should provide the effect
4.0/). of precipitation hardening. Additionally, the alloy addition of Zr increases the resistance

Materials 2022, 15, 4023. https://doi.org/10.3390/ma15114023 99 https://www.mdpi.com/journal/materials


Materials 2022, 15, 4023

to secondary recrystallization and increases the tensile strength. It was revealed [10] that
the observed intermediate layer is formed by the deposition of aluminum on a titanium
substrate due to the interaction of high temperature and pressure. The thickness of this layer
depends primarily on the reaction time of the materials at temperature above 873 K. The
observed interlayer is formed by mutual diffusion until the formation of Al3 Ti intermetallic
phases. In case of realization of the fusion at the temperature reaching 1073 K, initially
Ti3 Al and TiAl phases are formed. Further maintaining the temperature at 1073 K results
in TiAl phases gradually transforming into Ti3 Al. These transformations lead to the
formation of continuous homogeneous Ti3 Al layer on the titanium surface. In case of Ti3 Al
intermetallic phases, the microhardness can increase up to two times. While the research
on microstructure and phase transitions are well described, this study has been undertaken
on the influence of heat treatment on the global state of the material.
The purpose of this research is to investigate the influence of hardening heat treatment
on the mechanical properties and state residual stresses of explosive welded sheet of
AA2519-Ti6Al4V with interlayer of AA1050.

2. Materials and Equipment


The subject of the study was an explosive-welded sheet of AA2519-Ti6Al4V with inter-
layer of AA1050. The AA2519 aluminum alloy is characterized by good fracture toughness
because of the copper content of 5.3–6.4%. An important advantage is its increased ballistic
resistance and the appropriate weldability. This alloy is strengthened by precipitations
through the process of supersaturation and aging. Investigated variety of aluminum alloy
was produced by adding the alloying elements Zr and Sc in the amount of 0.1–0.25% for
both. The addition of Sc has the effect of fragmenting the structure of the AA2519 alloy,
which improves the strength properties of the material. Precipitation strengthening was
obtained and the resistance to secondary recrystallization of microstructure increased. The
primary effect is a greater stability of mechanical properties at elevated temperature as a
result of homogeneous dispersion of Al3 Zr phase. After casting, the alloy was subjected
to hot rolling and then annealing at a temperature of 693 K in air condition for 2 h. A
fine-grained structure was obtained, which has a positive effect on the increase of plasticity
and fatigue strength. The strength properties and chemical composition of the modified
AA2519 alloy are given in Table 1.

Table 1. Strength properties and chemical composition of the AA2519 alloy.

Strength Properties Chemical Composition [wt%]


R0.2 [MPa] Rm [MPa] A [%] Si Fe Cu Mg Zn Ti V Zr Sc Al
312 335 6.5 0.06 0.08 5.77 0.18 0.01 0.04 0.12 0.2 0.36 rest

The second basic material of the laminate, Ti6Al4V alloy, is used in special construc-
tions which is caused by its high mechanical properties, corrosion resistance, and a melting
point reaching 1955 K. For this reason, it is used in the production of aircraft components,
including: the jet engine rotor blades or the wing caissons. The strength properties and
chemical composition of the Ti6Al4V alloy are presented in Table 2.

Table 2. Strength properties and chemical composition of the Ti6Al4V alloy.

Strength Properties Chemical Composition [wt%]


R0.2 [MPa] Rm [MPa] A [%] O V Al Fe H C N Ti
950 1020 14 <0.20 3.5 5.5 <0.30 <0.0015 <0.08 <0.05 rest

Explosive joining of materials with different mechanical properties requires the usage
of an interlayer. The AA1050 alloy was used, because it has a very good adhesive property.
The strength properties and chemical composition are shown in Table 3.

100
Materials 2022, 15, 4023

Table 3. Strength properties and chemical composition of the AA1050 alloy.

Scheme Chemical Composition [wt%]


R0.2 [MPa] Rm [MPa] A [%] Fe Si Zn Mg Ti Mn Cu Al
78 168 2.9 0.4% <0.25 <0.07 0.18 <0.05 <0.05 <0.05 rest

The layered material subjected to testing was manufactured in the Department of


High Energy Technologies EXPLOMET Galka Szulc general partnership. The functional
properties of explosives, including the detonation velocity, are derived from their chemical
composition, the physical form of their individual components, e.g., fragmentation, the
degree of pore surface development present in the structure of crystals of the ammonium
nitrate used, and the quality of their mixing. The assumed plan of plating tests included
the execution of joints in the AA2519-Ti6Al4V double-layer system with a separate spacer
made of AA1050 aluminum alloy (Figure 1) [11].

Figure 1. Schematic diagram of the collision course of explosion welded three-layer: β—angle of
collision of combined materials, β1 —angle of collision of base material with interlayer material,
β2 —angle of collision of overlay material with interlayer material Vc —velocity of collision point
with respect to contact point of welded plates, Vc1 —velocity of collision point of contact between
base material and interlayer material, Vc2 —velocity of the collision point of the contact point of
the base material and the interlayer, D—velocity of detonation of the explosive, H—height of the
explosive layer, g1 —thickness of the plate to be shot, g2 -thickness of the base plate, g3 —thickness of
the interlayer.

The final products were in the form of plates with dimensions of 200 × 150 mm and
thickness of 6.7 mm, approximately (Figure 2). The thickness can slightly vary because
of the dynamic nature of the joining process. The plates were rolled in order to give
them adequate flatness. Both described technological processes significantly deformed
the material and introduced elastic-plastic residual strains. In the next step the plates
were annealed at the temperature of 803–823 K for 2 h, cooled down in water at room
temperature and aged at 438 K for 10 h. The main aim of this treatment was to obtain an
effect of precipitation hardening. Besides the changes of microstructure and mechanical
properties this treatment has seriously influenced the residual stresses state.

Figure 2. Explosive welded sheet of AA2519-Ti6Al4V with interlayer of AA1050 after heat treatment.

101
Materials 2022, 15, 4023

Monotonic tensile tests were carried out on specimens made of manufactured laminate
after heat treatment. All the specimens subjected to axial tension had the same geometry.
Five specimens were made of each material (Figure 3). Tensile testing of AA2519-Ti6Al4V
laminate with AA1050 interlayer in axial tension conditions was performed according to
PN-EN ISO 6892-1:2010 on an Instron 8802 monotonic and fatigue hydraulic pulser.

(a) (b)
Figure 3. Dimensions of samples for tensile testing of AA2519-Ti6Al4V laminate with inter-layer of
AA1050 (a) and prepared test specimens (b).

Microhardness tests were carried out using a fully automatic universal laboratory
hardness tester DURA SCAN 70. The microhardness test was conducted in accordance
with EN ISO 6507-1. The indenter load was F = 0.9807 N. The process time to load the
indenter to the nominal load was 3 s. It took 10 s to load the sample to 0.98 N.
Digital image correlation (DIC) allows the measurement of deformations and defor-
mation of the surface of the tested object (Figure 4). The digital image correlation method
is used in optical, non-contact, and three-dimensional systems for real-time measurement
of displacement and deformation. The DIC system allows speckle images to be captured
using CCD cameras, while further data analysis and processing are carried out using Istra
4D software.
The procedure for measuring residual stresses using the hole trepanation method is
standardized by ASTM Standard Test Method E 837 [12], and the technical procedure of
implementation is described in Tech Note TN-503 [13]. Micro-Measurement RS-200 milling
guide was used to conduct the hole drilling process, necessary in order to measure the resid-
ual stresses. This equipment enables precise drilling of holes in the center point of strain
gauge rosette which is the most important condition for making correct measurements.
Figure 5 shows the distribution of gauge rosettes.

102
Materials 2022, 15, 4023

Figure 4. Digital image correlation measurement system from Dantec.

(a) (b)
Figure 5. Arrangement of gauge rosettes on TiAl4V side 9 (a) AA2519 (b).

Furthermore, it enables to control the depth of hole and to measure its diameter. The
carbide milling cutters with a diameter of 1.6 mm were used during the tests and these
tools were driven by high-speed rotation air turbine supplied by compressor. Testing stand
is presented in Figure 6.

(a) (b)
Figure 6. The overall view of testing stand (a) and the milling guide of RS-200 (b).

Surface residual stresses in the two main, perpendicularly oriented directions (“σ1,
σ2”) based on sin2ψ diffractometric measurements were obtained using an X-ray diffrac-
tometer -Bruker D8 Discover (Bruker Corporation, 40 Manning Road, Billerica, Mas-
sachusetts, United States of America) with a Euler wheel and a sample positioning system
along the three axes. Test samples for the research were prepared using electrical discharge
machining. Radiation and beam optics were characterized by CoKα filtration. Phase
analysis was performed in the Crystal Impact Match software with an ICDD PDF 4+ 2019

103
Materials 2022, 15, 4023

crystal-lographic database. TARSIuS (Texture-Aided Residual Stress Investigation System)


pack-age is a software developed by the Institute of Metallurgy and Materials Sciences of
the Polish Academy of Sciences in Krakow and it is used to visualize the residual stresses
in materials. Because this method is quite new, we decided to implement it in a short
explanation of its basics with examples. The areas were near to the connection of the plates
and the detailed locations are shown in the Figure 7.

(a) (b)
Figure 7. The surface area of the sample (a) and Bruker D8 Discover diffractometer with Euler
wheel (b).

3. Results
In the Figure 8 the view of plates after heat treatment and the metallographic cross-
section of the composite are shown.

Figure 8. Explosive welded sheet of AA2519–Ti6Al4V with interlayer of AA1050) a metallographic


cross-section of the composite.

The picture in the Figure 1 has shown that the plates after heat treatment are signifi-
cantly distorted. This deformation is caused by the difference of the value of the coefficient
of thermal expansion and the characteristic temperatures of phase transitions.
The microstructure of the heat treated of explosive welded sheet of AA2519-Ti6Al4V
with interlayer of AA1050 in the vicinity of joint is shown in the photographs (Figure 9).

104
Materials 2022, 15, 4023

(a) (b) (c) (d)


Figure 9. Metallographic cross-section of the produced explosive welded sheet of AA2519-Ti6Al4V
with interlayer of AA1050 after heat treatment: (a) sector I–Ti6Al4V alloy; (b) sector II–AA1050;
(c) sector III–AA1050; (d) sector IV–AA2519 alloy.

Heat treatment has not affected the structure of titanium alloy (Figure 9a). The AA1050
interlayer by the boundary with titanium is characterized by the grain growth with uniform
distribution (Figure 9b). Near the AA2519 alloy, the grains of the AA1050 alloy structure
have an equiaxial character (Figure 9c). For AA2519 alloy, the heat treatment process
improved the solubility of copper-rich precipitates in the aluminum matrix. The effect
of this process is the fine-grained structure and steady distribution of precipitates in the
matrix. At the interface between AA1050 aluminum and AA2519 alloy, a sublayer of
about 4–6 μm width was observed, which was formed as a result of heat treatment. It
is characterized by a fragmented structure and numerous voids with diameters of about
0.3–0.5 μm (Figure 9d). The slight voids were probably produced accidentally during the
supersaturation and annealing process.

3.1. Strength Properties


Tensile tests of AA2519-Ti6Al4V laminate with interlayer of AA1050 were conducted
according to [14]. The tests were carried out using five samples and obtained results were
repeatable. Samples were cut along the direction of explosive welding. The results obtained
during the tests are presented in the form of graphs in Figure 10.

Figure 10. Stress–strain curves of AA2519-Ti6Al4V laminate with interlayer of AA1050 in production
state and after heat treatment.

The heat treatment has resulted in an increase of the ultimate strength Rm from
657 MPa up to 704 MPa, which is about 7%, whereas the yield strength R0.2 has increased
from 436 MPa to 498 MPa and the increase is about 14%. On the other hand, the heat treat-

105
Materials 2022, 15, 4023

ment has negative influence on the elongation, A, which has decreased from 6.5% to 5.8%.
Based on these results, it can be stated that the ductile of heat-treated laminate decreased.
During tensile tests, an analysis of strain distribution using was applied using digital
image correlation (DIC). In case of the laminate subjected to additional heat treatment,
the strain distribution for the AA2519 alloy, up to the moment of a sample rupture, is
characterized by high uniformity (Figure 11).

Figure 11. Strain distribution using DIC method for selected points of the stress-strain curve obtained
for AA2519-Ti6Al4V laminate with interlayer of AA1050 after heat treatment: 1—The yield strength
R0.2; 2—the plastic strain of 2%; 3—the ultimate strength Rm; 4—the fracture of the specimen.

In case of the Ti6Al4V alloy, after exceeding the yield point, the strain maps illustrate
the occurrence of horizontal bands. DIC analysis made on the lateral surface has not
revealed local or banded strain inhomogeneity, which makes it impossible to predict the
location of cracking even under high loadings.

3.2. Microhardness
The measurement of microhardness HV0.1 was conducted according to the stan-
dard [15]. The study was performed in the immediate vicinity of the center layer. Obtained
results are presented in the Figure 12, where the AA1050 intermediate layer is in grey.
The microhardness of the AA2519 alloy after heat-treatment was below 100 HV. In case
of the interlayer, which is composed of the AA1050 alloy, there is noticeable a strengthening
effect near the boundary with the AA2519 due to the formation of intermetallic precipitates.
In case of Ti6Al4V alloy, the microhardness after heat treatment was about 350 HV and has
not changed compared to the state before heat treatment [16].

106
Materials 2022, 15, 4023

Figure 12. The results of microhardness measured in AA2519–Ti6Al4V laminate with interlayer
of AA1050.

3.3. Measurements of Residual Stresses


The applied processes of explosive joining and following rolling significantly influ-
enced the stress state of the material. It is important to define this state, therefore the
residual stresses were measured in two perpendicular planes, namely in the plane parallel
to the surface and in the cross section. The first measurements were conducted using
hole-drilling method and the second by X-ray diffraction.

3.3.1. Hole-Drilling Method


This method was used to determine the state of surface residual stresses and their
change with the thickness. The measurements were made at three randomly placed points
on both sides of the plate, from the aluminum and titanium. In the presence of residual
stresses, even a small hole into the material causes stress relaxation because each normal
stress component perpendicular to the free surface (e.g., the hole surface) is zero. Moreover,
this relaxation is related to the change in strain field in the direct vicinity of the hole
according to Hook’s law. This method is often referred to as a “semi-destructive” technique
because in many cases a small hole does not significantly damage the structural integrity
of the test object (the dimensions of the hole are usually 0.8–4.8 mm-diameter and depth).
After measuring large objects, it is possible to remove the hole by careful welding and
smoothing the surface with a grinder.
The obtained output voltage was transmitted from the rosette to the channels of the
ESAM Traveler Plus strain gauge bridge, where it was amplified. An exemplary graph
presenting the changes of voltage recorded during the test is presented in Figure 13a.

(a) (b)
Figure 13. The overall view of testing stand (a) and the milling guide of RS–200 (b).

107
Materials 2022, 15, 4023

The calculation of the strain values was made up to the depth of 1.0 mm employing
Equation (1).
E = 4 × UOUT / (U0 × N × K × A) × 106 [με = 10−6 ] (1)
where:
UOUT —output voltage [V];
U0 —supply voltage [V];
N—coefficient depending on the bridge type—for quarter-bridge N = 1;
K—strain gauge constant;
A—coefficient of output signal amplification.
As a result of the measurements and calculations carried out, the characteristics of
changes in relative strain values as a function of the depth of the hole drilled were obtained.
To determine the values of principal stresses and their direction, the following relations
were applied.

ε1 + ε2 1
σmax = − (ε 3 − ε 1 )2 + (ε 3 + ε 1 − 2ε 2 )2 [MPa] (2)
4· A 4· B

ε1 + ε2 1
σmax = + (ε 3 − ε 1 )2 + (ε 3 + ε 1 − 2ε 2 )2 [MPa] (3)
4· A 4· B
1 ε − 2ε 2 + ε 2
α = arctg 1 [MPa] (4)
2 ε2 − ε1
where:
σmax , σmin —principal stresses;
ε1 , ε2 , ε3 —strains measured on strain gauges number 1, 2, and 3;
A, B—coefficients depending on material properties and geometry of rosette and hole;
α—angle between strain gauge no. 1 and the direction of the nearest principal stress.
The result of performed analysis is presented in Figure 13b. Calculated changes
of strains (dashed lines) correspond to the measured data (diamonds). This accordance
was gained under the assumption that the residual stresses vary linearly with the depth.
Determination of the stress value and its character was accomplished using software H-
Drill. Obtained results of measured principal residual stresses and the angle from the
principal axis to defined direction, are presented in Table 4. The minus value of the angle
indicates that the angular deviation is counterclockwise.
The character of stresses was analyzed up to the depth of 1.0 mm and in all cases a
linear character of the stress changes was revealed. This assumption was justified by the
values of statistical parameter RMS misfit which vary from 1.9 me to 2.8 me for AA2519 and
from 4.6 me to 6.0 me for Ti6Al4V. If the uniform state of stresses was assumed, the RMS
misfit would have vary from 6.8 me to 11.6 me and from 9.1 me to 16.2 me, respectively.
Determined values of principal stresses in aluminum alloy AA2519 vary with the
depth within the ranges: σmin from −102 MPa up to +90 MPa and σmax from −27 MPa
up to +146 MPa. On the surface, the values of σmin and σmax are −96.0 ± 5.9 MPa and
−20.7 ± 7.6 MPa, respectively. The stresses are compressive and at different depths change
to tensile stresses. Comparing the principal directions of residual stresses with the geometry
of investigated plate it can be stated that σmin is oriented perpendicular to the distortion,
namely parallelly to the maximum gradient, and σmax opposite. The values of principle
stresses in titanium layer change with the ranges: σmin from +8 MPa up to +376 MPa and
σmax from +132 MPa up to +509 MPa and they were determined at different depths. The
residual stresses are tensile and increase linearly for the surface values of σmin and σmax
equal to +36.0 ± 21.7 MPa and +141.3 ± 11.8 MPa, respectively. It must be pointed that
the maximum values determined at greater depths can be subjected to high uncertainty.
When drilling a hole in the titanium alloy, high cutting resistance was observed which may
influence the closest vicinity of the hole, thus the near-surface values of residual stresses
are most reliable. The directions of principal stresses are oriented similarly to previously
described but σmin and σmax are swapped because on this side of the laminate the surface

108
Materials 2022, 15, 4023

is convex (on the side of aluminum is concave) which changes the compressive stresses
to tensile.
Table 4. The results of residual stresses measurements of the heat treated laminate AA2519-Ti6Al4V
laminate with interlayer of AA1050.

AA2519
Number of Measuring Point
#1 #2 #3
depth [mm] 0.0 0.5 1.0 0.0 0.5 1.0 0.0 0.5 1.0
residual stresses s [MPa] and angular orientation a [◦ ]
smax −10 49 108 −27 59 146 −25 56 136
smin −102 −66 −31 −98 −26 45 −88 1 90
a −81 −77 −74 −88 −82 −77 −86 −86 −87
Ti6Al4V
Number of measuring point
#1 #2 #3
depth [mm] 0.0 0.5 1.0 0.0 0.5 1.0 0.0 0.5 1.0
residual stresses s [MPa] and angular orientation a [◦ ]
smax 134 285 436 132 166 201 158 329 509
smin 8 120 232 39 85 128 61 223 376
a −32 −29 −28 −19 −25 −31 −6 7 17

3.3.2. X-ray Diffraction Method


Based on the gained data, the two-dimensional maps of residual stresses were devel-
oped whose visualization enable easy interpretation of the stress state (Figure 14).
Determined values of residual stresses measured in the vicinity of the joint show that
the directions of principal stresses are oriented approximately parallelly/perpendicularly
to this joint. It is indicated by the orientation of the angular marks in the circles placed at
the measuring points (Figure 14). Residual stresses are both compressive and tensile in
AA2519, but in Ti6Al4V are only compressive. The measurement results presented in form
of maps are the most reliable for the titanium alloy layer. The maximum residual is in its
central part, where they reach the value of −1250 MPa. The lower values of compressive
stresses in the horizontal direction may be related to the change of shape of the plate after
heat treatment, which is slightly distorted.
The values of stresses in the layer of AA2519 alloy should be treated as speculative,
because the values significantly exceeded the yield point of this material. This is most likely
due to the formation of intermetallic phases in the aluminum and grain growth.

109
Materials 2022, 15, 4023

(a)

(b)
Figure 14. The maps of residual stresses measured in two perpendicular directions X and Y in the
(a) AA2519 alloy and (b) Ti6Al4V; description in text.

4. Summary and Conclusions


The results of measurements the residual stresses obtained by two different methods
seem to be inconsistent with each other thus this issue needs wider explanation in order
to clarify doubts. First difference relates to the orientation of the measuring plane—the
hole-drilling method (HDM) was applied in plane parallel to the surface contrary to X-ray
method (XRM), which was applied in perpendicular plane. Second difference is related
with the place of measurements and state of the material. HDM has been adopted to
determine the stresses on the surface and at shallow depth of the material in the form of
full-scale plate after heat treatment without any additional mechanical processing. Whereas
XRM allowed to measure the stress state near the interlayer, but the small sample must be
cut from the plate for proper preparation. This process can influence the global stress state.
The last difference relates to the nature of the measurement. HDM allows to determine
only the value of the first-order residual stresses because the resultant strains are measured
in some distance from the measurement point. XRM is based on the measurement of the
distance between the crystallographic planes thus the values can be connected with the
microscale second- or even third-order residual stresses and the microscale mechanical
properties. The influence of this problem was observed during the measurement of stresses
in aluminum alloy when the determined values significantly overcame the macroscale
tensile strength. The above-described differences between HDM and XRM indicate that
comparisons of results should be made with caution.
The investigation was performed to analyze the influence of the heat treatment on the
structure and the mechanical properties of the AA2519-Ti6Al4V laminate with interlayer of

110
Materials 2022, 15, 4023

AA1050. The main purpose of the heat treatment was to obtain an effect of precipitation
hardening in aluminum alloy, which was successfully gained. Nevertheless, this process has
seriously influenced the residual stresses state. The laminate was distorted because of the
differences in thermophysical properties thus there were performed wide investigation on
the state of residual stresses. On the basis of the obtained results the following conclusions
can be stated:
1. A heat treatment affects the laminate resulting in an increase of the ultimate strength
Rm about 7% to the value of 704 MPa and the yield strength R0.2 about 14% to the
value of 498 MPa, but it has also negative influence on the ductility decreasing an
elongation from 6.5% to 5.8%;
2. As a result of heat treatment, the microhardness of titanium alloy remains unchanged
but in AA2519 it increased from app. 95 HV0.1 [13] to the value of 150 HV0.1 ;
3. The heat treatment in AA2519 caused fine-grained structure with steady distribution
of precipitates in the matrix and additionally a slight sublayer of about 4–6 μm width
with some voids observed at the interface between this material and interlayer;
4. The surface residual stresses were determined by the hole drilling method and are
closely related to distortion. They are compressive on the side of aluminum alloy (to
−102 MPa) and tensile on the side of titanium alloy (up to +158 MPa);
5. The residual stress state in cross-section was determined by X-ray diffraction method.
They are compressive reaching −1250 MPa in Ti6Al4V. The results for AA2519 are
very dubious.

Author Contributions: Conceptualization, T.Ś. and I.S.; methodology, L.Ś. and I.S.; validation, T.Ś.,
L.Ś. and I.S.; formal analysis, I.S. and T.Ś.; investigation, I.S.; resources, I.S.; data curation, T.Ś.;
writing—original draft preparation, I.S. and T.Ś.; visualization, T.Ś.; supervision, L.Ś.; All authors
have read and agreed to the published version of the manuscript.
Funding: This research was supported by the Polish Ministry of National Defence [grant number
PBG/13-998]. The project is carried out under Project PBS2/A5/35/2013 funded by the National
Research and Development Centre.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Informed consent was obtained from all subjects involved in the study.
Data Availability Statement: Not applicable.
Acknowledgments: We would like to thank the Explomet company from Opole, which performed
explosive bending of titanium and aluminum alloys.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Saravanan, S.; Raghukandan, K.; Hokamoto, K. Improved microstructure and mechanical properties of dissimilar explosive
cladding by means of interlayer technique. Arch. Civ. Mech. Eng. 2016, 16, 563–568. [CrossRef]
2. Loureiro, A.; Carvalho, G.H.S.F.L.; Galvão, I.; Leal, R.M.; Mendes, R. Chapter 6–Explosive Welding. In Advanced Joining Processes;
da Silva, L., El-Zein, M., Martins, P., Eds.; Elsevier: Amsterdam, The Netherlands, 2021; pp. 207–237. [CrossRef]
3. Sniezek, L.; Szachogluchowicz, I.; Gocman, K. The Mechanical Properties of Composites AA2519-Ti6Al4V Obtained by Detonation
Method. In Proceedings of the 8th International Conference on Intelligent Technologies in Logistics and Mechatronics Systems,
ITELMS 2013, Panevėžys, Lithuania, 23–24 May 2013; pp. 214–219.
4. Szachogłuchowicz, I.; Hutsaylyuk, V.; Sniezek, L. Low cycle fatigue properties of AA2519-Ti6Al4V. Procedia Eng. 2015, 114, 26–33.
[CrossRef]
5. Bataev, I.A.; Bataev, A.A.; Mali, V.I.; Pavliukova, D.V. Structural and mechanical properties of metallic–intermetallic laminate
composites produced by explosive welding and annealing. Mater. Des. 2012, 35, 225–234. [CrossRef]
6. Rohatgi, A.; Harach, D.J.; Vecchio, K.S.; Harvey, K.P. Resistance-curve and fracture behavior of Ti–Al3 Ti metallic–intermetallic
laminate (MIL) composites. Acta Mater. 2003, 51, 2933–2957. [CrossRef]
7. Jiang, S.-Y.; Li, S.-C.; Zhang, L. Microstructure evolution of Al-Ti liquid-solid interface. Trans. Nonferrous Met. Soc. China 2013, 23,
3545–3552. [CrossRef]

111
Materials 2022, 15, 4023

8. Luo, J.G.; Acoff, V.L. Using cold roll bonding and annealing to process Ti/Al multi-layered composites from elemental foils.
Mater. Sci. Eng. A 2004, 379, 164–172. [CrossRef]
9. Peng, L.M.; Li, H.; Wang, J.H. Processing and mechanical behavior of laminated titanium–titanium tri-aluminide (Ti–Al3 Ti)
composites. Mater. Sci. Eng. A 2005, 406, 309–318. [CrossRef]
10. Peng, L.; Wang, J.; Li, H.; Zhao, J.; He, L. Synthesis and microstructural characterization of Ti–Al3 Ti metal–intermetallic laminate
(MIL) composites. Scr. Mater. 2005, 52, 243–248. [CrossRef]
11. Walczak, W. Zgrzewanie wybuchowe metali i jego zastosowania. Zesz. Nauk. 2002, 51, 15–22.
12. ASTM Standard E 837; Determining Residual Stress by the Hole-Drilling Strain-Gage Method. ASTM International: West
Conshohocken, PA, USA, 2020.
13. Tech Note TN-503-6; Measurement of Residual Stresses by the Hole-Drilling Strain Gage Method. Micro-Measurements. Vishay
Precision Group: Malvern, PA, USA, 2010.
14. ISO 6892-1:2016; Metallic Materials—Tensile Testing—Part 1: Method of Test at Room Temperature. National Standards Authority
of Ireland: Dublin, Ireland, 2016.
15. ISO 6507-1:2007; Metallic Materials—Vickers Hardness Test—Part 1: Test Method. European Committee for Standardization:
Brussels, Belgium, 2018.
16. Szachogluchowicz, I.; Śnieżek, L.; Śl˛ezak, T.; Kluczyński, J.; Grzelak, K.; Torzewski, J.; Fras, T. Mechanical Properties Analysis of
the AA2519-AA1050-Ti6Al4V Explosive Welded Laminate. Materials 2020, 13, 4348. [CrossRef] [PubMed]

112
materials
Article
Effect of Steel-Cutting Technology on Fatigue Strength of Steel
Structures: Tests and Analyses
Sławomir Rowiński

Faculty of Civil Engineering, Wrocław University of Science and Technology, 50-370 Wrocław, Poland;
[email protected]; Tel.: +48-71-320-2905

Abstract: This paper presents the results of comparative fatigue tests carried out on steel S355J2N
specimens cut out using different cutting methods, i.e., plasma cutting, water jet cutting, and
oxyacetylene cutting. All the specimens were subjected to cyclic loading from which appropriate S-N
curves were obtained. Furthermore, face-of-cut hardness and roughness measurements were carried
out to determine the effect of the cutting method on the fatigue strength of the tested steel. The fatigue
strength results were compared with the standard S-N fatigue curves. The fatigue strength of the
specimens cut out with oxyacetylene was found to be higher than that of the specimens cut out with
plasma even though the surface roughness after cutting with plasma was smaller than in the case of
the other cutting technology. This was due to the significant effect of material hardening in the heat-
affected zones. The test results indicate that, in comparison with the effect of the cutting technology,
the surface condition of the specimens has a relatively small effect on their fatigue strength.

Keywords: steel structures; fatigue strength of steel; hardness; roughness; plasma cutting; water jet
cutting; gas cutting; composite dowel

Citation: Rowiński, S. Effect of


Steel-Cutting Technology on Fatigue
Strength of Steel Structures: Tests and
1. Introduction
Analyses. Materials 2021, 14, 6097.
https://doi.org/10.3390/ma14206097
Most of the steel frame structures currently built and used in, i.a., the construction
industry, the marine industry, and the manufacturing industry have their parts joined
Academic Editors:
together and properly shaped by, e.g., welding and cutting out from larger steel elements.
Jaroslaw Galkiewicz and Currently, oxyacetylene cutting, plasma cutting, and water jet cutting are the predominant
Lucjan Śnieżek cutting technologies. In the case of oxyacetylene cutting, computer-controlled devices
usually execute the cutting line with an accuracy of 0.8–1.6 mm. The width of the slit
Received: 23 August 2021 depends on the cutting parameters, i.e., the diameter and shape of the oxygen nozzle, the
Accepted: 12 October 2021 cutting oxygen and inflammable gas pressure, and the cutting speed. After oxygen cutting,
Published: 15 October 2021 the cut heat-affected zone (CHAZ) is relatively wide and depends on the alloying element
content in the material. In the case of low-carbon steel plates, the width of CHAZ amounts
Publisher’s Note: MDPI stays neutral to less than 0.8 mm at the thickness of 12.5 mm and to about 3 mm at the thickness of
with regard to jurisdictional claims in 150 mm [1]. Plasma cutting consists in melting metal and ejecting it from the slit with a
published maps and institutional affil- strongly concentrated electric arc flowing between a nonconsumable electrode and the
iations. workpiece. The width of CHAZ is inversely proportional to the cutting speed and depends
on the composition (conductivity) of the material being cut. In the case of 25 mm thick
18-8-type austenitic steels, CHAZ is 0.08–0.13 mm wide at the cutting speed of 1.2 m/min.
Water jet cutting consists in using a strongly compressed water jet formed by passing
Copyright: © 2021 by the author. water through a small-diameter nozzle. The water jet removes the cut material from the
Licensee MDPI, Basel, Switzerland. cutting slit through erosion and cutting fatigue under high local stresses and, additionally,
This article is an open access article through micromachining when abrasive powder with a (gamet, olivine, or silica) grain
distributed under the terms and size of 0.3–0.4 mm is used. The temperature of the cut edges does not exceed 100 ◦ C (cold
conditions of the Creative Commons cutting). The cut material thickness depends on the water jet cutting parameters, i.a., the
Attribution (CC BY) license (https:// cutting speed, the water pressure, the kind and grain size of the powder, and the powder
creativecommons.org/licenses/by/ feed rate [1].
4.0/).

Materials 2021, 14, 6097. https://doi.org/10.3390/ma14206097 113 https://www.mdpi.com/journal/materials


Materials 2021, 14, 6097

Thanks to the significant differences between the above-mentioned cutting technolo-


gies, they find application in the manufacture of all kinds of steel frame structures. The
choice of a cutting technology depends on the quality requirements and the manufacturer’s
technical and financial capabilities as well as on the requirements specified by the design
engineer who takes into consideration the effect of the particular technologies of cutting
out a steel frame structure on the latter’s ultimate and fatigue strength.
The adverse effect of the technology of cutting out a structure on the latter’s fatigue
strength was observed during extensive experimental studies on innovative connectors of
the composite dowel type used in composite steel and concrete bridges [2]. Such connectors
are created by appropriately joining together steel structural members and concrete. The
innovative composite dowel joint is based on the idea of cutting the rolled beam’s steel web
in two along a specifically shaped line (Figure 1) so that the dowels obtained in this way in
each of the two parts when embedded in concrete will constitute a mechanical connector
carrying the delamination forces between the steel and the concrete [3–5].

Figure 1. Cut-out MCL: (a) steel connectors with marked parts to be removed, (b) connector before separation of two parts
of cut beam, (c) component parts of the innovative joint with MCL dowels.

It was only after the PreCo-Beam [2] had been completed when the effect of the cutting
technology on the load-bearing capacity of the composite dowel joint was given some
thought. It was found that the roughness of the dowel’s front face after oxyacetylene
cutting could be the cause of fatigue cracking in beams under cyclic loading [6–8].
In the literature, one can find the results of experimental studies of the effect of the
technology of cutting steel on the latter’s strength. However, one should bear in mind that
such studies do not take into account the complexity of this problem (particularly in the
construction industry) as the experiments are conducted not on full-size models of the
structures but on specimens. Moreover, the results are for the element’s particular shape,
thickness, and steel grade. The effect of cutting technologies and the obtained specimen
surface characteristics on the fatigue strength is described in the work of [9]. Specimens
6 and 12 mm thick with a steel strength of 240–900 MPa were tested. It was confirmed
that the fatigue strength increases with the tensile strength of the steel and depending
on the surface roughness. The fatigue strength of the specimens increases when their
surface roughness is reduced by additional surface treatments, such as sandblasting [10].
The oxygen cut specimens have the highest fatigue strength, followed by the laser and
plasma cut specimens. The gas-cut specimens have the highest surface roughness but
also the highest compressive residual stress state. The plasma cut specimens have the
lowest roughness, but their residual stresses are practically zero in comparison with the
oxygen and laser-cut specimens [11]. The fatigue strength of plasma cut surfaces can
be significantly improved with a post-cutting treatment applicable. The improvement
is achieved by introducing compressive residual stress and reducing surface roughness
height through grinding [12].
For steel S690Q and steel S355M, it is observed that when straight edges are cut with
plasma and a laser, these cutting technologies improve, in comparison with gas cutting, the

114
Materials 2021, 14, 6097

fatigue strength [13,14]. Laser-cut steel S890Q was found to have higher fatigue strength
than when cut with gas or plasma [15].
It emerges from the tests carried out on specimens that the fatigue strength of steel
frame structures is a complex problem sensitive to many factors. In construction, this
problem is further compounded by the fact that it is not possible to directly observe the
initiation and propagation of cracks (the connector is embedded in concrete) and also by
the complicated interactions in the joint (delamination forces change from dowel to dowel
and are transmitted through the direct pressure of the concrete against the front faces of the
dowels and via the adhesive forces between the steel beam’s flat surfaces and the concrete).
Therefore in order to assess the effect of the cutting technology on the fatigue strength of
the material, comparative tests were carried out on dumbbell-shaped specimens cut out
using different cutting technologies.

2. Experiment
2.1. Test Specimens and Test Plan
The specific specimen shape and dimensions (thickness and fillet radius) (Figure 2)
were adopted so that the test results could apply to connectors of the composite dowel type
(Figure 1). The specimens were cut out of 10 mm thick plate S355J2N along the plate rolling
direction. Three series of specimens were cut out using different cutting technologies
designated as: A—water jet, B—oxyacetylene, and C—plasma.

(a) (b) (c) (d)


Figure 2. Specimen for comparative cyclic tests: (a) geometry (mm), (b) series A (water jet cut)
specimen, (c) series B (oxyacetylene cut) specimen, (d) series C (plasma cut) specimen.

The tests were carried out on a 100 kN testing machine. The specimens were subjected
to uniaxial tension-compression (R = −1) in the high cycle fatigue range until fracture. The
load spectrum was sinusoidal with a frequency f = 10 Hz. Through trials, such a frequency
was selected that the temperature of the specimens during tests would not exceed 60 ◦ C.
The tests were conducted in a cyclic testing machine in set stress ranges Δσi (Table 1) on one
to four specimens for each level Δσi . The number of cycles N = 5 m was set as the lifespan
limit. The force signal and the total deformation signal were registered during the tests.
Deformation was measured along the gauge length of 25 mm by means of an extensometer.

115
Materials 2021, 14, 6097

Table 1. Test stress ranges Δσi for cutting technologies.

Δσ Δσ Δσ
Specimen Type Specimen Type Specimen Type
(MPa) (MPa) (MPa)
150 200 125
175 225 150
B
A 200 250 C 175
Oxyacetylene
Water jet cutting 300 300 Plasma cutting 200
cutting
325 - 225
350 - -

The specimens would most often fail due to rupture in the fillet (geometric notch)
area. Photographs of selected specimens after failure for each of the cutting technologies
are shown below (Figure 3).

(a) (b) (c)

Figure 3. Failed specimens cut out with: (a) water, (b) oxyacetylene, (c) plasma.

2.2. Results
The test results are presented in Tables 2–4.

Table 2. Test results for specimens A.

Specimen No. (-) F (kN) Δσ (MPa) Nf (Cycles)


12A 5,402,341
13A 30 150 5,467,028
18A 5,034,480
11A 779,609
16A 35 175 1,209,227
17A 2,744,983
14A 536,918
40 200
15A 336,773
1A 7619
2A 60 300 9174
3A 8339
7A 3722
8A 65 325 3330
9A 3339
5A 2175
6A 70 350 2225
4A 70 350 2129
22A 1225

116
Materials 2021, 14, 6097

Table 3. Test results for specimens B.

Specimen No. (-) F (kN) Δσ (MPa) Nf (Cycles)


13B 5,476,832
40 200
14B 3,463,974
8B 5,249,668
9B 3,475,725
10B 45 225 3,568,974
11B 5,882,781
12B 2,870,595
3B 338,745
50 250
4B 1,030,958
5B 13,463
6B 60 300 32,353
7B 15,718

Table 4. Test results for specimens C.

Specimen No. (-) F (kN) Δσ (MPa) Nf (Cycles)


12C 5,000,000
13C 25 125 5,000,000
14C 5,000,000
7C 5,000,000
10C 30 150 1,714,866
11C 430,863
3C 1,143,120
5C 35 175 444,094
6C 716,537
9C 40 200 196,487
2C 141,694
4C 45 225 70,782
8C 108,665

Figure 4 shows the test results as logarithmic stress amplitude versus the logarithmic
number of cycles for the three cutting technologies. Test results regression curves for each
of the technologies were calculated. The standard curves for the fatigue categories of
80 MPa and 125 MPa according to standard [16] were included for comparison.
From the regression curves for the respective cutting technologies, Δσc stress values at
N = 2 million cycles, i.e., the fatigue categories, and values of fatigue curve slope cotangent
m were calculated. The results are presented in Table 5.

117
Materials 2021, 14, 6097

Figure 4. Results of tests carried out on specimens. (oxyacetylene, water, plasma).

Table 5. Values of Δσc and m and regression curves for considered cutting technologies.

Regression Fatigue Category Fatigue Curve


Cutting Technology
Curve Equation Δσ c (MPa) Slope m
water A Δσ = 764.6·(N)−0.105 167 10
oxyacetylene B Δσ = 531.6·(N)−0.059 226 17
plasma C Δσ = 990.6·(N)−0.132 146 8

3. Cut Edge Conditions


3.1. Macroscopic and Microscopic Examinations of Fatigue Fractures
Detailed macroscopic and microscopic analyses of selected specimens cut out with:
water (2A, 14A, 17A), oxyacetylene (4B, 10B, 14B), and plasma (6C, 8C, 10C) were car-
ried out. The specimens’ faces of cut and fatigue fracture and crack surfaces (Figure 5)
were examined.

Figure 5. Test specimen: L—face of cut, N—fatigue fracture, K—area from which samples were taken
to make metallurgical polished sections, Z-Z’—cross-sections on which polished sections were made,
g—metallurgical polished section hardness measuring length.

The macroscopic examinations were performed under a stereoscopic light microscope,


while the microscopic examinations were carried out using a confocal laser scanning
microscope and a scanning electron microscope. Samples for preparing metallurgical

118
Materials 2021, 14, 6097

polished sections were cut out using a precision cutter and mounted in conductive resin.
The mounted samples were ground and polished on a polishing machine and subjected to
etching with 5% HNO3 solution.
The tested specimens’ faces of cut showed numerous furrows resulting from cutting,
along which fatigue cracks propagated (Figures 6 and 7). This is particularly visible for
water jet cutting, in which case the face of the cut is distinctly varied, showing an area of
the entry of the water jet with an abrasive and an exit area (Figure 8).

(a) (b) (c)


Figure 6. Fracture of specimen 6C: (a) face of cut made with plasma, (b) numerically denoted fracture
sides no. 1, (c) numerically denoted fracture sides no. 2.

(a) (b) (c)

Figure 7. Fracture of specimen 14B: (a) face of cut made with gas, (b) numerically denoted fracture
sides no. 1, (c) numerically denoted fracture sides no. 2.

(a) (b) (c)


Figure 8. Fracture of specimen 17A: (a) face of cut made with water jet, (b) numerically denoted
fracture sides no. 1, (c) numerically denoted fracture sides no. 2.

The examinations of the fractured surfaces of the specimens clearly corroborated the
fatigued character of the fractures, as visible in Figure 9.

119
Materials 2021, 14, 6097

Figure 9. Fractured surface of specimen 14B: 1—brittle fracture area, 2—ductile fracture area, 3—
granular area (end phase of fracture).

3.2. Investigations of Face-of-Cut Roughness


Roughness, i.e., the arithmetic mean deviation of the profile from the mean line, on
the two faces of cut denoted in Figure 10 was investigated for each specimen. The results
are presented in Table 6.

Figure 10. Denotations of surfaces subjected to roughness measurements.

Table 6. Results of roughness measurements: Ra —roughness, Ra,mean —mean roughness.

Specimen No. (-) Ra (μm) Ra,mean (μm)


L P
2A 3.358 3.529
14A 3.889 3.311 3.496
17A 3.715 3.172
4B 1.531 1.719
10B 1.158 1.166 1.444
14CB 1.530 1.561
6C 0.296 0.369
8C 0.314 0.269 0.208
10C 0.355 0.329

120
Materials 2021, 14, 6097

3.3. Microscopic Examinations of Metallurgical Polished Sections


For microscopic examinations, a sample was cut out from the measurement area of
the specimens consistently with the Z-Z’ plane, parallel to the fracture surface (Figure 5).
After etching with 5% HNO3 solution, a nonequilibrium ferritic-pearlitic structure became
visible (Figure 11).

(a) (b) (c)

(d) (e) (f)

(g) (h) (i)


Figure 11. Microstructures of specimens: (a–c) cut out with water, (d–f) cut out with oxyacetylene, (g–i) cut out with plasma.

3.4. Investigations of Cut Heat-Affected Zone


The specimens cut out using the gas technology, and the plasma technology showed a
distinctly changed structure at the cut edge (Figures 12 and 13) due to the local heating of
the material. The cut heat-affected zone (CHAZ presented in Table 7) was characterized
by a martensitic structure with varied carbon content. This structure, unlike the ferritic-
pearlitic structure, was more brittle and susceptible to cracking (Figure 14). No significant
changes in microstructure were observed in the case of the water cut specimens.

3.5. Hardness Tests


Hardness was measured using the Vickers method in accordance with the stan-
dard [17]. The measurements were performed under the load of 10 kg acting over the time
of 10 s.
The hardness measurements were carried out on metallurgical polished sections along
segment g (Figure 5) consistent with the pearlite-ferrite banding. The distance between the
cut edge and the first impression amounted to 0.3 mm. The next impressions were spaced
at every 1.5 mm. In total, 10 impressions were made for each specimen.
The hardness of the tested specimen materials ranged from 146 to 352 HV10 (Figure 15).
Hardness values above 150 HV10 were measured exclusively at the edges of the specimens

121
Materials 2021, 14, 6097

cut with oxyacetylene and plasma. The average hardness in these places amounted to
244 HV10 and 310 HV10 for the type B specimens and the type C specimens, respectively.

(a) (b) (c)

Figure 12. Structures in cut heat-affected zone in specimens: (a) 4B, (b) 10B, (c) 14B—oxyacetylene cutting technology.

(a) (b) (c)


Figure 13. Structures in cut heat-affected zone in specimens: (a) 6C, (b) 8C, (c) 10C—plasma cutting technology.

Table 7. Depth of cut heat-affected zone (CHAZ).

Specimen No. (-) CHAZ (μm)


4B 450
10B 550
14B 550
6C 505
8C 517
10C 475

122
Materials 2021, 14, 6097

(a) (b) (c) (d)

Figure 14. Structures: (a) acicular structure in CHAZ in specimens cut with plasma, (b) acicular structure in CHAZ in
specimens cut with oxyacetylene, (c) pearlitic-ferritic structure in specimens cut with plasma, (d) pearlitic-ferritic structure
in specimens cut with oxyacetylene.

(a) (b) (c)

Figure 15. Hardness of type: (a) A specimens (water), (b) B specimens (oxyacetylene), (c) C specimens (plasma).

4. Analysis of Results
The specimens cut out using the oxyacetylene technology showed a longer fatigue
lifespan than the other specimens cut out with plasma and water, as presented in Figure 6.
Fatigue categories were calculated from the regression curves of the cutting technolo-
gies. The fatigue categories amounted to: Δσc ,B = 226 MPa for oxyacetylene cutting,
Δσc ,A = 167 MPa for water jet cutting and Δσc ,C = 146 MPa for plasma cutting. The results
are above fatigue category Δσc = 125 MPa according to the work of [16], which is safely and
most frequently adopted by building designers and applies to gas-cut metal plates with
removed edge discontinuities. Unfortunately, in the case of series-produced connectors of
the composite dowel type, each of the production operations, including the machining of
the face of cut, is thought to add to the cost, and so efforts are made to reduce the latter by,
i.a., limiting the additional treatments of the face of the cut. The current design guidelines
according to the work of [16] do not take into account the effect of the cutting technology
and the quality of the cut surface on the fatigue strength of the structure. Therefore it is
necessary to clarify and specify more precisely the fatigue category for other cutting tech-
nologies, including water jet cutting and plasma cutting, which should have a beneficial
effect on the design of steel frame structures.
In all the considered cases, the fatigue curve slope cotangents m were larger (mA = 10,
mB = 17, mC = 8) in comparison with the standard curves according to the work of [16],
for which m = 3. The slope values provide information about the speed of fracture of the
specimens under variable load. In the case of oxyacetylene cutting, the specimens would
fracture slowest (mB = 17), as opposed to the specimens cut out with plasma, which would
fracture fastest (mC = 8). It should be noted that the results of this test depend on, i.a.,
the number and shape of the specimens, the character of the fatigue load, and the yield
point of the material. Therefore it is difficult to directly compare the obtained results. A

123
Materials 2021, 14, 6097

convergence between the results is considered to be a satisfactory outcome. In the case


of steel S355, one gets curve slope m = 7 for oxygen cut specimens and m = 13 for plasma
cut specimens [11], or m = 5.2 for plasma, m = 5.8 for oxygen and m = 16.8 for water, as
described in the work of [18]. The ongoing research confirms the conservative standard
recommendations for which m = 3 [16]. It should be added that the obtained moderate
conservatism can be proper considering that small-scale specimens, in general, ensure
greater fatigue reliability than large-scale beam specimens [19].
The material tests corroborated the relatively smaller surface roughness for plasma
cutting (Ra,mean = 0.208 μm) and oxyacetylene cutting (Ra,mean = 1.444 μm) than for water
jet cutting (Ra,mean = 3.496 μm). The numerous furrows in the water jet entry zones in the
water cut surfaces were micronotches in which the initiation of fatigue fractures would
take place. It is noteworthy that the choice of cutting parameters and the thickness of the
metal plate being cut have a bearing on the quality of the cut-out specimen’s surface [18].
The effect of the kind of machining and the surface layer condition on the fatigue
strength is expressed by surface condition coefficient βp as a ratio of the fatigue strength
of an unnotched (polished) specimen to the latter’s strength after machining. The higher
the surface condition coefficient, the lower the specimen’s fatigue strength due to surface
irregularities. Figure 16 shows the results of experiments [20] in which the effect of the
kind of machining (grinding, fine rolling, coarse rolling) on the value of coefficient βp ,
depending on the tensile strength, was studied. As one can see, the surface condition
coefficient increases with surface roughness. Additionally, the mean surface roughness
values of the faces of cut of in-house specimens of type A, B, and C for steel S355J2N with
tensile strength fu = 510 MPa were included in the figure.

Figure 16. Effect of kind of machining on the value of surface condition coefficient βp for tension or
bending, depending on tensile strength of steel and kind of machining for: 0—polished, 1—ground,
2—fine rolled, 3—coarse rolled, and 4—sharply ring-notched (for comparison) specimens. Adapted
with permission from Ref. [20], 2021, Wydawnictwo Naukowe PWN.

Table 8 contains the measured values of βp (according to Figure 16) and the calculated
fatigue categories Δσc for each of the cutting technologies and the percentage differences
relative to the results for the specimens cut out with plasma. Judging by the differences,
the condition of the surface of the specimens has a relatively small (up to 12%) effect on
their fatigue strength in comparison with the technologies used to cut them (55%).

124
Materials 2021, 14, 6097

Table 8. Values of βp and Δσc and differences between results for tested specimens.

Cutting Difference between Difference between


βp (-) Δσ c (MPa)
Technology βp Results (%) Δσ c Results (%)
Water (A) 1.13 12 167 14
Oxyacetylene (B) 1.05 4 226 55
Plasma (C) 1.01 0 146 0

The fatigue strength of the specimens cut out with oxyacetylene (Δσc = 226 MPa) is
higher than that of the specimens cut out with plasma (Δσc = 146 MPa) even though the
surface roughness after cutting with plasma is smaller than in the case of the other cutting
technology. This is due to the significant effect of material hardening in the heat-affected
zones. In both cases, acicular structures (Figure 12) with comparable heat-affected zone
depths measured from the surface of the cut were obtained. However, in the case of plasma
cutting, the hardness measured at the cut edge (310 HV10) was 27% greater than for the
specimens cut out using the gas cutting technology (244 HV10). A similar agreement
between the results was obtained in the tests described in the work of [18], where the
specimens cut out with plasma were characterized by the highest hardness (280 HV10) and
lower fatigue strength (Δσc = 239 MPa) in comparison with the oxygen cut specimens for
which hardness amounted to 190 HV10 and fatigue strength to 264 MPa.
The results of the comparative tests indicate that the gas cutting technology used so far
to cut out connectors for the innovative composite dowel joint is more advantageous than
the plasma cutting technology or the water cutting technology. Furthermore, oxyacetylene
cutting is the cheapest and most available cutting technology in prefabrication plants.

5. Conclusions
From the results of the fatigue tests carried out on steel S355J2N specimens cut out
using different cutting methods, i.e., plasma cutting, water jet cutting, and oxyacetylene
cutting, the following conclusions, providing a basis for further analyses leading to the
development of design guidelines for steel connectors of the composite dowel type, can
be drawn:
1. The technology of cutting out dowels of the composite dowel type has a bearing on
their fatigue strength. Connectors cut out using oxyacetylene cutting can have higher
fatigue strength than the ones cut out using plasma cutting or water jet cutting;
2. The effect of the technology used to cut out steel connectors of the composite dowel
type can be greater than that of the condition of the face of the cut;
3. The slopes of the fatigue strength curves determined for the cut-out specimens:
mA = 10 for water jet cutting, mB = 17 for oxygen cutting, and mC = 8 for plasma
cutting, corroborate the conservative standard recommendation m = 3 according to
the work of [16];
4. The FAT125 fatigue curve according to the work of [16] can be appropriate for the
design of composite dowel connectors to be cut out using oxygen cutting, plasma
cutting, and water jet cutting. Nevertheless, further experimental studies (the S-N
curve method) need to be carried out on beam specimens of composite structures in
order to verify the fatigue curve for the composite dowel connector.

Funding: This research received no external funding.


Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Data sharing not applicable.
Conflicts of Interest: The authors declare no conflict of interest.

125
Materials 2021, 14, 6097

Abbreviations
The following symbols and notations, in the order in which they appear in the text, are used in
this paper:

CHAZ Cut heat-affected zone


R Stress ratio
Δσ Nominal stress range
F Force
Nf Number of cycles to failure
logA Intercept of mean S-N curve
m Slope of fatigue strength curve
Ra Surface roughness
βp Coefficient of surface condition

The other symbols used in this paper are explained when they appear in the text for the first time.

References
1. Klimpel, A. Welding, Autogenous Welding and Cutting of Metals; Wydawnictwo Naukowo–Techniczne: Warszawa, Poland, 1999;
pp. 633–654. (In Polish)
2. Seidl, G.; Popa, N.; Zanon, R.; Lorenc, W.; Kożuch, M.; Rowiński, S.; Franssen, J.-M.; Fohn, T.; Hermosilla, C.; Farhang,
A.; et al. RFCS Dissemination Knowledge Project PRECO+: Prefabricated Enduring Composite Beams Based on Innovative Shear
Transmission, RFS2-CT-2011-00026 “PRECO+” Design Guide. Available online: http://www.stb.rwth-aachen.de/projekte/2005
/INTAB/downloadPreco.php (accessed on 21 August 2021).
3. Lorenc, W.; Kożuch, M.; Rowiński, S. The behaviour of puzzle-shaped composite dowels. Pt. 1, experimental study. J. Constr.
Steel Res. 2014, 101, 482–499. [CrossRef]
4. Lorenc, W.; Kożuch, M.; Rowiński, S. The behaviour of puzzle-shaped composite dowels. Pt. 2, Theoretical investigations.
J. Constr. Steel Res. 2014, 101, 500–518. [CrossRef]
5. Kożuch, M.; Rowinski, S. Elastic behaviour of the steel part of a shear connection with MCL composite dowels. Steel Constr. 2016,
9, 107–114. [CrossRef]
6. Dudziński, W.; P˛ekalski, G.; Harnatkiewicz, P.; Kopczyński, A.; Lorenc, W.; Kożuch, M.; Rowiński, S. Study on fatigue cracks in
steel-concrete shear connection with composite dowels. Arch. Civ. Mech. Eng. 2011, 11, 839–858. [CrossRef]
7. Rowiński, S. Fatigue Strength of Steel Dowels in Innovative Shear Connection of Steel–Concrete Composite Beam. Ph.D. Thesis,
Wrocław University of Technology, Wrocław, Poland, 2012; pp. 11–72. (In Polish).
8. Rykaluk, K.; Marcinczak, K.; Rowiński, S. Fatigue hazards in welded plate crane runway girders–Locations, zauses and calculation.
Arch. Civ. Mech. Eng. 2017, 18, 69–82. [CrossRef]
9. Sperle, J.-O. Influence of parent metal strength on the fatigue strength of parent material with machined and thermally cut edges.
Weld. World 2008, 52, 79–92. [CrossRef]
10. Remes, H.; Korhonen, E.; Lehto, P.; Romanoff, J.; Niemelä, A.; Hiltunen, P.; Kontkanen, T. Influence of surface integrity on the
fatigue strength of high-strength steels. J. Constr. Steel Res. 2013, 89, 21–29. [CrossRef]
11. Stenberg, T.; Lindgren, E.; Barsoum, Z.; Barmicho, I. Fatigue assessment of cut edges in high strength steel-influence of surface
quality. Mater. Sci. Eng. Technol. 2017, 48, 556–569. [CrossRef]
12. Lillemäe, I.; Liinalampi, S.; Lehtimäki, E.; Remes, H.; Lehto, P.; Romanoff, J.; Ehlers, S.; Niemelä, A. Fatigue strength of high
strength steel after shipyard production process of plasma cutting, grinding and sandblasting. Weld. World 2018, 62, 1273–1284.
[CrossRef]
13. García, T.; Álvarez, J.A.; Cicero, S.; Carrascal, I.; Martin-Meizoso, A. Effect of thermal cutting methods on the fatigue life of high
strength structural steel S690Q. In Proceedings of the ASME Proceeding 2015 Pressure Vessel and Piping Conference, Anaheim,
CA, USA, 19–23 July 2015.
14. Garcia, T.; Cicero, S.; Carrascal, I.; Madrazo, V.; Alvarez, J.A. Effect of cutting method on fatigue crack initiation and fatigue life of
structural steel S355M. In Proceedings of the ASME Proceeding 2014 Pressure Vessel and Piping Conference, Anaheim, CA, USA,
20–24 July 2014; Volume 3.
15. Cicero, S.; García, T.; Álvarez, J.A.; Bannister, A.; Klimpel, A.; Martín-Meizoso, A.; Aldazabal, J. Fatigue behavior of high strength
steel S890Q containing thermally cut straight edges. Procedia Eng. 2016, 160, 246–253. [CrossRef]
16. Eurocode 3: Design of Steel Structures. Part 1–9: Fatigue. 2005. Available online: https://www.phd.eng.br/wp-content/uploads/
2015/12/en.1993.1.9.2005-1.pdf (accessed on 20 September 2021).
17. PN-EN ISO 6507-1:1999. Metals. Vickers hardness measurement. In Test Method; ISO: Geneva, Switzerland, 1999. (In Polish)
18. Diekhoff, P.; Hensel, J.; Nitschke-Pagel, T.; Dilger, K. Investigation on fatigue strength of cut edges produced by various cutting
methods for high-strength steel. Weld. World 2020, 64, 545–561. [CrossRef]

126
Materials 2021, 14, 6097

19. Keating, P.B.; Fisher, J.W. Evaluation of fatigue tests and design criteria on welded details. In National Cooperative Highway Research
Program (NCHRP), Report 286; Transportation Research Board: Washington, DC, USA, 1986.
20. Dietrich, M. Fundamentals of Machine Construction; Wydawnictwo Naukowo–Techniczne: Warszawa, Poland, 1999; Volume 1,
pp. 319–485. (In Polish)

127
materials
Article
The Influence of Heat Treatment on the Mechanical Properties
and Corrosion Resistance of the Ultrafine-Grained AA7075
Obtained by Hydrostatic Extrusion
Marta Orłowska 1, *, Ewa Ura-Bińczyk 2 , Lucjan Śnieżek 1 , Paweł Skudniewski 2 , Mariusz Kulczyk 3 ,
Bogusława Adamczyk-Cieślak 2 and Kamil Majchrowicz 2

1 Faculty of Mechanical Engineering, Military University of Technology, Gen. S. Kaliskiego 2,


00-908 Warsaw, Poland; [email protected]
2 Faculty of Materials Science and Engineering, Warsaw University of Technology, Wołoska 141,
02-507 Warsaw, Poland; [email protected] (E.U.-B.); [email protected] (P.S.);
[email protected] (B.A.-C.); [email protected] (K.M.)
3 Institute of High Pressure Physics, Polish Academy of Sciences, Sokolowska 29/37 St., 01-142 Warsaw, Poland;
[email protected]
* Correspondence: [email protected]; Tel.: +48-261-839-245

Abstract: In this paper, the corrosion resistance and mechanical properties of the 7075 aluminum
alloy are studied. The alloy was deformed by hydrostatic extrusion and then aged both naturally and
artificially. Results are compared with those of coarse-grained material subjected to T6 heat treatment.
The aim of the research is to find the optimal correlation between the mechanical properties and the
corrosion resistance of the alloy. To this end, static tensile tests with subsequent fractography, open
circuit potential, and potentiodynamic polarization tests in 0.05 M NaCl were conducted. Obtained
Citation: Orłowska, M.; Ura-Bińczyk, results show that a combination of precipitate hardening and a deformed microstructure leads to
E.; Śnieżek, L.; Skudniewski, P.; increased mechanical strength with high anisotropy due to the presence of fibrous grains. Plastic
Kulczyk, M.; Adamczyk-Cieślak, B.; deformation increases susceptibility to corrosion due to the increased number of grain boundaries,
Majchrowicz, K. The Influence of which act as paths along that corrosion propagates. However, further artificial aging incurs a positive
Heat Treatment on the Mechanical effect on corrosion resistance due to changes in the chemical composition of the matrix as a result of
Properties and Corrosion Resistance the precipitation process.
of the Ultrafine-Grained AA7075
Obtained by Hydrostatic Extrusion.
Keywords: 7075 aluminum alloy; hydrostatic extrusion; heat treatment; microstructure; mechanical
Materials 2022, 15, 4343. https://
properties; corrosion resistance
doi.org/10.3390/ma15124343

Academic Editor: Young Gun Ko

Received: 2 June 2022


1. Introduction
Accepted: 18 June 2022
Published: 20 June 2022 The 7XXX series of Al–Zn–Mg(–Cu) aluminum alloys (AAs) are highly important
industrial materials, primarily due to their light weight, high strength, and good corrosion
Publisher’s Note: MDPI stays neutral
resistance. Such materials are commonly used in the construction of aircraft, among other
with regard to jurisdictional claims in
uses. As a group of materials, the 7XXX alloys are particularly interesting because of
published maps and institutional affil-
the process of precipitation hardening, which causes a complex microstructure. Such a
iations.
microstructure is composed of fine hardening precipitates and coarse intermetallic phases
within the Al matrix [1]. For the 7XXX family, this includes Al3 Fe, Al2 CuMg, Al2 Cu,
Al3 (FeCu), Al7 Cu2 Fe, or Mg2 Si particles. The hardening precipitates are formed and dis-
Copyright: © 2022 by the authors. tributed throughout the matrix via proper heat treatment: supersaturated solid solutioning
Licensee MDPI, Basel, Switzerland. followed by aging. The principal precipitation sequence that dominates hardening in most
This article is an open access article commercially used 7XXX alloys is presented in Equation (1):
distributed under the terms and
conditions of the Creative Commons SSSS∝ → GP zones → η → η (1)
Attribution (CC BY) license (https://
creativecommons.org/licenses/by/ where SSSSα represents a supersaturated solid solution, GP zones are Guinier–Preston
4.0/). zones, η is a metastable phase (with a Mg:Zn ratio in the range from 1:1 to 1:1.15), and η is

Materials 2022, 15, 4343. https://doi.org/10.3390/ma15124343 129 https://www.mdpi.com/journal/materials


Materials 2022, 15, 4343

a stable MgZn2 phase. MgZn2 plays the most significant role in the precipitate hardening
process [2].
The mechanical properties of Al–Zn–Mg–Cu quaternary alloys are determined pri-
marily by precipitation hardening, where the nanometer-scale precipitates act as pinning
centers to prevent dislocation motion [3]. Hardening is achieved via the presence of
microstructural obstacles that inhibit dislocation movement and improve the overall me-
chanical properties of the alloy [4]. Such obstacles are also grain boundaries. Reducing
the grain size increases the fraction of grain boundaries within the material and corre-
spondingly increases the mechanical strength of the material. The application of severe
plastic deformation (SPD) [5,6] leads to grain formation and achieves an ultrafine-grained
(UFG) microstructure, which substantially increases mechanical strength. Following SPD,
7XXX alloys exhibited an increase in tensile strength from 228 to 434 MPa [7]. However,
the greatest increases in strength are obtained when SPD is combined with precipitate
strengthening. The application of hydrostatic extrusion followed by aging increases the
tensile strength of AA 7475 to 700 MPa [8]. As such, the optimal microstructure to improve
the mechanical strength of 7XXX alloys has a fine grain size with a uniform dispersion of
small hard particles to inhibit dislocation motion.
Along with mechanical properties, corrosion resistance is also affected by both pre-
cipitates and grain boundaries. Nevertheless, there is no clear correlation, as corrosion
resistance is influenced by various microstructural factors and also the environment in
which the examination takes place. Therefore, determining the influence of one microstruc-
tural component on the corrosion resistance is challenging. Nevertheless, there are works
devoted to these topics. In the case of grain size, susceptibility to corrosion may decrease [9];
in the case of AAs, it reduces cathodic kinetics, improves the results of mass-loss testing,
and in some cases improves resistance to stress corrosion cracking [10]. Studies that sug-
gest that corrosion resistance increases with decreasing grain size generally find that the
refined microstructures are attributed to an increased readiness to passivate compared to
coarse-grained (CG) materials due to a higher grain boundary density [11]. Further to grain
refinement, SPD physically breaks down second-phase or intermetallic particles below
a critical size, which subsequently prevents these second-phase particles from operating
as efficient local cathodes and sites for the initiation of localized attack [9,12]. However,
some works present an opposing view: corrosion rate may also increase as AA grain size
increases [9,13]. In [14], friction stir processing was used to refine the microstructure of
AA 6063. The smaller grains produced by such an approach were more susceptible to
corrosion in 1 M HCl. Furthermore, Osorio et al. suggested that, when exposed to NaCl
electrolytes, CG AAs showed better corrosion resistance than that of fine-grained AAs, as
they had fewer corrosion initiation sites [15].
The purpose of this work is to correlate the changes in the microstructure caused by
plastic deformation and subsequent heat treatment with the mechanical properties and
corrosion resistance of the AA7075. However, in opposition to the previous works devoted
to this topic, the approach of maintaining the size and distribution of intermetallic particles
was undertaken. This is crucial in terms of corrosion resistance, as such particles are the
places where corrosion attack initiates due to their electrochemical properties; therefore,
their influence is the most significant. In the present study, due to preserving the size
of the intermetallic particles, other structural components were investigated, i.e., grain
boundaries and second-phase particles. As 7XXX AAs have a complex microstructure
that is determined by thermomechanical processing, the primary goal of this study is to
determine the optimal heat treatment parameters of ultrafine-grained AA 7075, obtained
by hydrostatic extrusion. The aim of the study is to increase the mechanical strength of
AA 7075 while maintaining its high corrosion resistance.

2. Materials and Methods


The research subject is commercially available AA 7075, with its chemical composition
given in Table 1, delivered in the form of rods with a diameter of ∅20. The manufacturer of

130
Materials 2022, 15, 4343

the alloy was Kamensk Uralsky Metallurgical Works. For the purpose of this study, the
samples underwent different combinations of heat treatment and SPD. The SPD method
used was hydrostatic extrusion [16]. In this approach, the billet is located in the container
and surrounded with a pressure transmitting medium. The piston compresses the medium
until the billet starts to extrude through the die. The negligible friction between the billet
and the die allows for the use of small die angles that ensure that the high deformation is
homogeneous. The unique features of hydrostatic extrusion are three-axial compressive
stresses within the billet and high strain rates greater than 104 s−1 . In comparison to
standard SPD methods, as a means of grain refinement, this process requires significantly
smaller total strain. In this study, hydrostatic extrusion was performed in a single step with
a reduction in the sample diameter from ∅20 to ∅10. The equivalent strain of hydrostatic
extrusion is given by Equation (2):


d
ε = 2ln (2)
d

where d is the initial diameter of the rod, and d” is the exit diameter. The equivalent strain
following the single-step hydrostatic extrusion process that we used was approximately 1.4.
The following list of samples were examined:
• CG, a coarse-grained precipitation-strengthened sample used as a reference material,
solution heat-treated with T6 aging following hydrostatic extrusion to maintain the
size and location of intermetallic inclusions.
• HE, an ultrafine-grained sample naturally aged for 180 days following hydrostatic
extrusion.
• HT1, an ultrafine-grained sample precipitation strengthened following hydrostatic
extrusion (artificially aged at 100 ◦ C for 24 h).
• HT2, an ultrafine-grained sample precipitation strengthened following hydrostatic
extrusion (artificially aged at 120 ◦ C for 24 h).
For the aging process, an SUP-30G furnace from WAMED Co. was used, which has
very high certainty of maintaining the temperature, i.e., ±0.2 ◦ C.

Table 1. Chemical composition of AA 7075.

Element Zn Mg Cu Cr Ti Si Fe Mn Al
Content (wt. %) 5.70 2.40 1.50 0.19 0.04 0.09 0.23 0.06 balanced

The microstructure and postcorrosion morphology of the samples were investigated


using a Hitachi Su70 scanning electron microscope (SEM). The surface corrosion attack was
observed with samples polarized up to 0.1 mA/cm2 , and the postcorrosion morphology
(the number and shape of pits) of the sample cross-sections was examined. The depth and
morphology of the pits were evaluated on the longitudinal cross-section of the sample,
oriented along the grains in the direction of hydrostatic extrusion). Energy-dispersive X-ray
spectroscopy (EDS) was used to perform the elemental analysis and chemical character-
ization of the intermetallic particles present in the microstructure. Electron backscatter
diffraction (EBSD) was used to examine the crystallographic orientation of grain and grain
boundary characteristics. We determined the crystallographic orientations of the grains,
the grain size, the fraction of high angle grain boundaries (HAGBs), and the fraction of low
angle grain boundaries (LAGBs). The data collection step size was 250 μm.
For detailed microstructure characterization, a JEOL JEM-1200 transmission electron
microscope (TEM) operating at 120 kV was used. TEM samples were sectioned along the
transverse and longitudinal axes of the rods. Thin TEM foils were prepared by grinding
slices down from a thickness of 1 mm into a thickness of 150 μm, and then electropol-
ishing them using a solution of Struers electrolyte A2 at 5 ◦ C and 25 V. To complement
the TEM investigations, we also used X-ray diffraction (XRD). To correctly identify the

131
Materials 2022, 15, 4343

visible precipitates on the TEM samples, XRD measurements were taken using a Bruker
D8 Advance diffractometer with Cu Kα radiation. Specimens were rotated at 10 rpm
during the diffraction recording. Diffraction lines were matched to the pattern using
EVA V3.0 software.
The mechanical properties of the processed samples were determined by tensile tests.
For each sample, five measurements were taken. Static tensile tests were carried out on flat
minisamples [17,18] with cross sections of 0.6 × 0.8 mm and gauge lengths of 5 mm. The
tests were performed at room temperature at an initial strain rate of 1 × 10−3 s−1 , using
digital image correlation for noncontact strain measurements. We determined quantitative
average values for ultimate tensile strength (UTS), 0.2% offset yield strength (YS), and
elongation at break (Eb ). Due to the size of the examined tensile test samples, measurements
were taken from two directions, longitudinal and transverse, with respect to the extrusion
direction. This approach allowed for the anisotropy of the mechanical properties to be
examined. After the tensile tests, fractography was performed on an SEM.
Corrosion resistance was investigated via electrochemical experiments. Samples were
subjected to cyclic polarization in potentiodynamic (PP) mode. Before each test, each
sample was immersed in the prepared electrolytes for 10 min to stabilize the rest potential.
Additionally, the change in open circuit potential (EOCP ) was observed for 48 h to determine
the behavior of the material in the corrosive environment over an extended period of time.
To produce homogeneous, comparable surfaces and ensure the presence of a repeatable
passivation layer, all samples were ground by silicon carbide water grinding abrasive paper,
with grades of up to #4000, and then placed within a desiccator for 24 h prior to corrosion
testing. All tests were performed at room temperature, with a corrosive medium of Cl–
ions within an aqueous environment. To this end, a solution of 0.05 M NaCl using distilled
water was prepared. A NOVA AutoLab PGSTAT302N potentiostat was used to control the
potential between the working electrode (sample) and the Ag/AgCl reference electrode
while measuring the current between the working electrode and the platinum sheet counter
electrode. The surface of each sample was sealed in plastic, with a 20 mm2 area left exposed.
Corrosion tests were conducted on the surfaces perpendicular to the hydrostatic extrusion
direction. The cyclic PP with a 1 mV/s scan rate started from −0.05 V relative to the EOCP ,
and this was reversed when the current had reached a value of 0.3 mA. These measurements
were repeated three times to ensure the reproducibility of the results. Average corrosion
parameters, including corrosion potential (Ecorr ), corrosion current density (icorr ), and
repassivation potential (Erep ), were obtained from the curves.

3. Results
3.1. Microstructure
3.1.1. SEM/EBSD
Maps of the grain boundaries of CG and HE samples are presented in Figure 1. Both
the transverse and longitudinal planes are shown with respect to the extrusion direction.
For the latter, the extrusion direction is indicated by black arrows. There are noticeable
differences between the samples. In the transverse plane, the CG sample contained a large
amount of HAGBs, with a much smaller number of LAGBs distributed throughout. The
grains were equiaxial with an average grain size of 17.1 μm. The largest grain exceeded
40 μm in size. The hydrostatic extrusion process caused significant grain refinement,
resulting in an average grain size of 1.3 μm (including subgrains) for the HE sample. This
difference was primarily caused by a considerable increase in the density of LAGBs: the
CG sample contained 5.56% LAGBs, while the HE sample contained 76%. LAGBs are
predominantly located in the grain interiors, creating a network of subgrains surrounded
by the HAGBs which delineate and separate each grain.
In the longitudinal plane, HAGBs lay parallel with one another in accordance with
the extrusion direction, creating strongly elongated grains. The heat treatment of the CG
sample resulted in a larger average grain size of 17.6 μm. Nevertheless, the elongated shape
was preserved. Within the grains, a small number of LAGBs were observed. For the HE

132
Materials 2022, 15, 4343

sample, a network of LAGBs existed within the interiors of significantly elongated grains,
which were separated by the HAGBs. LAGBs propagated both perpendicular and parallel
to the hydrostatic extrusion direction. The average grain size of the HE sample within the
longitudinal plane was 1.7 μm. The length of the lamellas defined by the HAGBs exceeded
75 μm, while their maximal thickness did not exceed 10 μm.

(a) (b)

(c) (d)

Figure 1. Grain boundary maps, with LAGBs shown in red, and HAGBs shown in black: (a) CG
transverse plane, (b) CG longitudinal plane, (c) HE transverse plane, and (d) HE longitudinal plane;
black arrows indicate the direction of the hydrostatic extrusion.

The crystallographic orientation of the grains was also analyzed. Figure 2 shows
grain orientation maps of the transverse sample planes. The inverse pole figures for each
sample are shown as insets. A comparison of these maps shows that a strong texture
developed following hydrostatic extrusion. For the CG sample, the grain orientation
was scattered, with preferred orientations of <111>, <112>, and <012>. However, the
maximal measured intensity was only 1.68. For the HE sample, two distinct dominant
crystallographic orientations could be identified: <001> and <111>. Moreover, the maximal
intensity of 5.48 was larger than that of the CG sample.

133
Materials 2022, 15, 4343

(a) (b)
Figure 2. Orientation maps of (a) CG and (b) HE samples.

3.1.2. XRD
Samples were analyzed using XRD to identify the phases. The corresponding XRD
patterns are presented in Figure 3. XRD peaks could be identified mainly as ∝ Al solid
solution (labeled as Al), η-precipitates (labeled as MgZn2 ), and also some peaks as MgCuAl2
and Al3 Cu2 . In general, the XRD peaks were slightly stronger for the CG sample than
those for the other samples. However, the XRD diffraction peaks originating from the η
precipitates were weak for all samples. Following hydrostatic extrusion, a strong texture
caused the incomplete detection of peaks originating from the Al solid solution. The
nonequilibrium phase (η ) was also detected, but the extremely small size of the η -phase
precipitates resulted in diminished XRD peaks and significantly complicated their proper
description. Furthermore, for samples that had undergone hydrostatic extrusion, XRD
graphs were very similar, with the same identified constituents.

Figure 3. XRD patterns of examined samples (PDF no. of identified phases: PDF 00-004-0787—Al,
PDF 01-073-5874—MgCuAl2 , PDF 00-034-0457—MgZn2 , PDF 01-071-5716—Al3 Cu2 ).

134
Materials 2022, 15, 4343

3.1.3. TEM
TEM bright-field images of the CG sample are presented in Figure 4a,b. The microstruc-
ture of the CG sample consisted of equiaxed grains with finely dispersed η-precipitates
(MgZn2 ). The majority of such precipitates were located within the grains; a small fraction
were located at grain boundaries. Much smaller nonequilibrium precipitates (η ) were
homogeneously distributed in the grain interiors. These precipitates are the blurry disc- or
rod-shaped dots in the background of the images. Figure 4c,d show that the microstructure
of the naturally aged HE sample was significantly deformed. The dark grain areas likely
indicate the presence of a high density of dislocations. This suggests that the microstructure
obtained via the hydrostatic extrusion process was highly deformed, with significant stress
applied to the material that correspondingly resulted in the generation of dislocations.
Compared to the CG sample, the number and size of η precipitates visible within the grains
were significantly reduced for the HE sample. However, those that were observed were
homogeneously distributed within the microstructure. Moreover, the precipitates were also
present at the grain boundaries. Fine metastable precipitates (η ) could also be observed
within the grains.

(a) (b)

(c) (d)

Figure 4. TEM micrographs of the samples: (a) transverse plane of CG, (b) longitudinal plane of CG,
(c) transverse plane of HE, and (d) longitudinal plane of HE.

TEM micrographs of samples HT1 and HT2 are presented in Figure 5. Regarding
the HE sample, a deformed microstructure with a high density of dislocations could be
observed. Elongated grains were visible within the longitudinal plane, with precipitates
formed both within the grains and on the grain boundaries. The grains were elongated
along the extrusion direction, which is marked by white arrows. The η precipitates of the
HT1 and HT2 samples were much smaller than those of the CG sample, but larger than
those of the HE sample, indicating that an increased aging temperature increases the size
of η precipitates within the microstructure following hydrostatic extrusion. Moreover, the

135
Materials 2022, 15, 4343

numbers of precipitates in the HT1 and HT2 samples were increased in comparison to
those in the HE sample. As the aging temperature increased, the size of the precipitates
at the grain boundaries increased. The HT1 sample had slightly larger η precipitates than
those of the HT2 sample. For both samples, η precipitates were also visible.

(a) (b)

(c) (d)

Figure 5. TEM micrographs of the samples: (a) HT1 transverse plane, (b) HT2 transverse plane,
(c) HT1 longitudinal plane, and (d) HT2 longitudinal plane; white arrows indicate the direction of the
hydrostatic extrusion.

3.1.4. SEM/EDS
The intermetallic particles were examined using SEM/EDS. Figure 6 presents the
EDS mapping results, showing the chemical composition of the intermetallic particles
(inclusions) within the microstructure. As the CG sample was heat-treated following
hydrostatic extrusion to obtain a coarse-grained microstructure, there was no noticeable
difference between each sample in the shape and size of the particles. Results demonstrate
that the majority of the analyzed particles consisted of Al, Cu, and Fe, with a smaller
amount of Mn and a very small amount of Cr. The particles were comparatively large
and irregularly shaped, with sizes of 1–10 μm. The particles were formed during alloy
solidification and did not dissolve during subsequent thermomechanical processing [19].
From the literature, the particles that occur most commonly in AA 7075 and correspond
with the compounds that we identified are Al7 Cu2 Fe, Al2 CuMg, Al2 Cu, and Al3 Fe [20,21].
Compounds that contain Cu and Fe are nobler than pure Al, resulting in the dissolution of
more active matrix and leading to the appearance of rings around a nearly intact particle
or particle colony [22]. This phenomenon was observed later, during the postcorrosion
damage evaluation.

136
Materials 2022, 15, 4343

Figure 6. EDS mapping results of inclusions within AA 7075.

3.2. Mechanical Properties


The representative stress–strain curves are shown in Figure 7, while the average results
of YS, UTS, and Eb tests are presented in Figure 8. The values of UTS, YS, and Eb are shown
for the tests performed in the longitudinal and transverse directions with respect to the
extrusion direction. The CG sample had a UTS of approximately 530 MPa and a YS of
460 MPa. Slightly higher values were obtained in the longitudinal direction. Moreover,
Eb was approximately 4% larger in the longitudinal direction than it was in the transverse
direction. Following both hydrostatic extrusion and heat treatment, mechanical strength
increases and elongation decreases. Noticeable differences were observed on the two
measured directions. The highest value of UTS was obtained for the HE sample in the
longitudinal direction; in the transverse direction, HE, HT1, and HT2 all displayed similar
values. The value of YS gradually increased as the aging temperature increased. The
opposite relationship existed between Eb and aging temperature. A noticeable difference in
results obtained for the two measured directions could be observed. The values of all three
parameters were larger in the longitudinal direction. For the CG sample, the differences
were relatively small—approximately 10 MPa. However, the remaining samples displayed
much larger differences, in the range of 80–140 MPa for YS and 60–100 MPa for UTS. The
highest anisotropy of mechanical strength was observed for HE, while it was reduced after
subsequent heat treatment. The work hardening capacity, which is defined as the ratio
of UTS to YS, was the largest for the HE sample and gradually decreased for a higher
aging temperature.
The fracture surfaces of the samples are shown in Figure 9. Images were taken from
the tensile specimens cut from the transverse direction in relation to the extrusion direction.
In each case, the fracture was ductile and with a transcrystalline character. The main
observed feature was the presence of numerous equiaxial dimples. The CG sample was
larger than the samples after hydrostatic extrusion. In the case of the aging process, there
was no correlation. This indicates that the size and subsequently the number of dimples
were connected with grain size and dislocation density because these were the sites of the
nucleation of the cavities, which may result in the formation of dimples. In [23], a decrease
in grain size also led to a decrease in dimple size.

137
Materials 2022, 15, 4343

Figure 7. Representative stress–strain curves of the samples.

(a)

(b)

Figure 8. Values of UTS, YS, and Eb for each sample, measured in the (a) longitudinal and (b) trans-
verse directions with respect to the extrusion direction.

3.3. Corrosion
3.3.1. Electrochemical Properties
Figure 10 shows the EOCP over time for each sample during their 48 h immersion in
0.05 M NaCl. The values of EOCP sharply increased, followed by a steady decrease over
time. The only exception was the HE sample, which maintained a very unstable EOCP level
that oscillated between −0.70 and −0.67 V/Vref, producing an irregular line on the graph.
For the HT2 sample, the EOCP value stabilized at approximately −0.75 V/Vref, following
14 h of a sharp decrease. For the HT1 sample, the EOCP value decreased steadily over almost
35 h before stabilizing at −0.72 V/Vref. Similarly, the EOCP of the CG sample stabilized
at −0.70 V/Vref following a gradual decrease. After 48 h, the HT2 sample displayed

138
Materials 2022, 15, 4343

the lowest EOCP value of approximately −0.75 V/Vref, and the HE sample displayed the
highest value of approximately −0.67 V/Vref.

(a) (b)

(c) (d)

Figure 9. Fractures of the tensile samples cut from the transverse direction: (a) CG, (b) HE, (c) HT1
and (d) HT2.

Figure 10. EOCP over time for the CG, HE, HT1, and HT2 AA 7075 samples during 48 h immersion in
0.05 M NaCl.

139
Materials 2022, 15, 4343

Figure 11 shows representative cyclic PP curves during 600 s of EOCP stabilization


in 0.05 M NaCl solution. The electrochemical data obtained from the curves are listed in
Table 2. During the forward scan, the shape of each curve was similar: a rapid increase in
current density once corrosion potential (Ecorr ) had been reached. This indicates that the
anodic behavior was dominated by active dissolution. The cathodic current density during
the forward scan was the lowest for the HE sample, higher for the aged HT1 and HT2
samples, and highest for the CG sample. This may indicate the acceleration of the oxygen
reduction reaction. The b factor, which describes the inclination of the cathodic branch, was
the highest for the CG sample and decreased with increasing aging temperature. The lowest
Ecorr was observed for the HE sample. This shifted to more anodic values following aging.
Moreover, artificial aging lowers the corrosion current density (icorr ), which was the lowest
for the HT2 sample. This indicates that artificial aging at elevated temperatures reduces
susceptibility to localized attacks. During the reverse scan, the HE curve significantly
differed in shape. The hysteresis loops and repassivation potentials were somewhat similar
for the aged and recrystallized samples. A large hysteresis and much lower repassivation
potential (−780 mV/Vref) were recorded for the HE sample. This suggests differences in
repassivation, and consequently in the morphology of the corrosion attack [24].

Figure 11. PP curves of the CG, HE, HT1, and HT2 AA 7075 samples following 600 s of EOCP
stabilization in 0.05 M NaCl solution.

Table 2. Average values of corrosion potential (Ecorr ), corrosion current density (icorr ), and repassiva-
tion potential (Erep ), b factor calculated from the potentiodynamic polarization curves.

Sample Ecorr , mV/Ref icorr , μA/cm2 Erep , mV b Factor, V/dec


CG −590 ± 5 0.72 ± 0.08 −599 ± 3 0.048
HE −615 ± 4 0.78 ± 0.13 −780 ± 3 0.039
HT1 −604 ± 4 0.66± 0.21 −617 ± 6 0.024
HT2 −592 ± 3 0.50 ± 0.05 −600 ± 1 0.024

A comparison of the results of PP to the literature data for AA7075 shows that a
similar course of the curves was observed as that in [25]; however, due to the lower pH of
the 3.5% NaCl solution, the observed PP curves were shifted to less noble values, and a
breakage potential was also observed. Similar findings in the case of Ecorr were observed
for AA7075 examined in 3.5% NaCl [26], but the values of icorr were in the same range as
that in the present study. An abrupt increase in a current density was observed here after
reaching Ecorr , which indicated rapid dissolution. AA7075 was also investigated in [27],

140
Materials 2022, 15, 4343

where 0.001 M NaCl was used for corrosion tests. The obtained average values of Ecorr
were higher for about 50–100 mV, but the scattering of the results was more pronounced
in comparison to the results obtained in the present study. Similar values of icorr were
also observed.

3.3.2. Postcorrosion Morphology


Figure 12 presents the postcorrosion morphology following the PP tests. Corrosion
products were detected on the surface of several samples. The most substantial differences
between the samples were the depths and shapes of the pits rather than the number of pits
present. Images of the cross-sections are shown in Figure 13. CG and HE samples had the
deepest pits of approximately 100 and 130–150 μm, respectively. The HT1 and HT2 samples
had the shallowest pits, of depth 50–70 μm. Although the pits were relatively narrow, for
samples that had undergone hydrostatic extrusion were located in close proximity to one
another, creating a network of pits and creating a larger region of degradation. The pit
locations appeared to be correlated with grain shape and elongation in addition to the
locations of the intermetallic particles. The hydrostatic extrusion process tends to align
constituent particles into bands within the alloy, but no significant difference in their sizes
was noted. At multiple locations, the pitting and initial corrosion damage was located close
to the intermetallic inclusions, and propagated along the intermetallic particles. Moreover,
the intermetallic particles were frequently observed to be intact, with corrosion occurring at
the particle–matrix interface, resulting in the dissolution of the adjacent material (matrix).

(a) (b)

(c) (d)
Figure 12. Sample surfaces following potentiodynamic polarization in 0.05 M NaCl: (a) CG, (b) HE,
(c) HT1, and (d) HT2.

141
Materials 2022, 15, 4343

(a) (b)

(c) (d)

Figure 13. Sample cross-sections following potentiodynamic polarization in 0.05 M NaCl: (a) CG,
(b) HE, (c) HT1, and (d) HT2.

4. Discussion
4.1. Microstructural Evolution through Hydrostatic Extrusion with Subsequent Aging
Microscopic observations of the material following hydrostatic extrusion reveal that
the microstructure was severely deformed, with the grains significantly elongated (fibrous)
along the extrusion direction. Microstructural evolution in Al during hydrostatic extrusion
had been demonstrated [28] with three presented mechanisms of HAGB formation. The first
mechanism is based on the formation of new grains as a result of dynamic recrystallization,
the second describes continuous grain rotation as a result of mobile dislocation absorption
into grain boundaries, and the third concerns neighbor switching due to slip banding. In
general, the fraction of HAGBs increases with increasing strain. In this study, an applied
equivalent strain was estimated at ε = 1.4. This relatively small strain resulted in the
formation of a high density of dislocations. Higher strains would reduce the dislocation
density and increase the misorientation angles of the grain boundaries. EBSD analysis
showed that the samples had reached an average grain size of approximately 1.3 μm on
the transverse plane following hydrostatic extrusion. In comparison, the reference sample
displayed an average grain size of 17 μm. This shows that hydrostatic extrusion is an
efficient means of grain refinement and, at this strain level, the formation of subgrain
structures. The EBSD investigation further revealed that hydrostatic extrusion significantly
changed the characteristics of the grain boundaries, increasing the fraction of LAGBs within
the microstructure to 76%. The high fraction of LAGBs was caused by relatively low
accumulated deformation. Previous work showed that applying hydrostatic extrusion to
an Al–Mg–Si alloy at a similar level as that in the present study could further accommodate
plastic strain via the dislocation of cells [29]. Further deformation would increase the
misorientation angle between cells to above 15◦ , causing the formation of HAGBs.
Grain orientation analysis provided information about the favorable crystallographic
orientations within the microstructure following hydrostatic extrusion. The grains aggre-

142
Materials 2022, 15, 4343

gated into two dominant orientations: <001> and <111>. The <111> orientation is typical
for a fiber texture, whereas the <001> orientation may result from dynamic recrystallization
caused by the temperature increase during hydrostatic extrusion [28]. As was shown,
the grain orientation stimulates the deformation substructure: for grains with the <111>
orientation, dense dislocation walls are formed; for grains with the <001> orientation,
stable cell dislocation structures can be observed [30]. In contrast, within the CG sample,
no clear texture or preferred crystallographic orientation was identified. This may have
been caused by the solution annealing performed after the hydrostatic extrusion, leading
to grain recrystallization following plastic deformation.
In addition to grain size and texture, precipitates were also investigated. XRD studies
showed that MgZn2 precipitates formed within the structure as a result of the aging process.
In addition, η’ phases were observed by the TEM. An increase in aging temperature in-
creased the size of the MgZn2 precipitates within the microstructure. Hydrostatic extrusion
also influenced the precipitation phenomena. The work investigating an Al–Mg–Si alloy
showed that a heterogeneous microstructure following hydrostatic extrusion results in
differences in the precipitate sequence [30]. This is caused by thermal shocks during the
deformation process, which may induce the formation of clusters before the precipitate
forms. For pure Al, adiabatic heating during hydrostatic extrusion with a strain of ε = 1.4
should be approximately 120–130 ◦ C, and may influence or initiate the precipitation process
in the case of age-hardenable alloys [31]. For AA 7475, the hydrostatic extrusion process
changed the precipitation kinetics and phase stability induced by the grain boundaries [32].
Precipitation occurs along fast diffusion paths, such as grain boundaries and dislocations.
Therefore, a larger number of precipitates at grain boundaries was observed within a de-
formed microstructure, as in this study. However, due to the higher number of nucleation
sites, the precipitates were smaller than those in coarse-grained material. This effect was
most noticeable within the grain interiors. Nevertheless, the increase in temperature of
artificial aging led to an increase in the size of the precipitates.

4.2. Influence of Microstructural Evolution on Mechanical Properties


The evaluation of the mechanical properties revealed that hydrostatic extrusion caused
an increase in mechanical strength. The highest UTS (approximately 675 MPa) was obtained
for the HE sample in the longitudinal direction. In the transverse direction, the HE, HT1,
and HT2 samples demonstrated similar values (approximately 575 MPa) and greater than
those the CG sample. The increase in aging temperature caused an increase in YS. The
average tensile strength of the commercially available coarse-grained AA 7075 in a T6
state is approximately 570 MPa, with a YS of 500 MPa and an elongation at break of
approximately 9% [33]. This indicates that the combination of hydrostatic extrusion and T6
thermal treatment increased tensile strength by up to 18% when compared with the heat
treatment of CG AA 7075.
Hydrostatic extrusion significantly increases mechanical strength, which is caused
by the introduction of a considerable number of structural defects that inhibit the motion
of dislocations during deformation. In the case of pure metals, when these defects are
the only strengthening mechanism, the increase in mechanical strength is dependent on
the value of true strain [34]. Nevertheless, in the case of age-hardened alloys such as
AA 7075, strengthening from precipitates must also be considered. Following hydrostatic
extrusion, and with dislocation slip in the dominant deformation mechanism, the particle
strengthening of AA 7475 is limited by both enhanced grain boundary precipitation that
does not contribute to an increase in mechanical strength, and the smaller size of precipitates
within grain interiors, which are stronger barriers for moving dislocations [35]. Precipitates
within grain interiors were also larger for the CG sample (see Figure 4) than for samples
that had undergone hydrostatic extrusion. Therefore, precipitation strengthening is less
effective for UFG material, with a larger contribution by structural defects such as grain
boundaries and dislocations.

143
Materials 2022, 15, 4343

The anisotropy of mechanical properties was also investigated, as tensile tests were
performed on samples cut from two directions, transverse and longitudinal, with respect
to the extrusion direction. The microstructural anisotropy was significant as grains form
fibrous shapes along the extrusion direction. This caused long HAGBs with interior
LAGBs to create a network of subgrains. Regardless of the deformation ratio during
hydrostatic extrusion, the AA 6060 alloy was characterized by weaker strength in the
cross-section perpendicular to the extrusion direction [36], as was found in the present
study for AA 7075. The anisotropy of AA 6060 increased with an increase in ε UTS and
YS values indicated approximately 10% anisotropy, with higher values in the longitudinal
direction for ε = 2.28. The anisotropy of mechanical properties was caused by the strong
crystallographic macrotexture that develops in the material during hydrostatic extrusion.
As a result, the spacing between HAGBs was smaller in the longitudinal plane than that
in the transverse plane (see Figure 1). This caused a higher volume of HAGBs, which
are stronger barriers for moving dislocations [34,37]. As a result, the samples cut in the
longitudinal direction possessed higher mechanical strength.

4.3. Influence of Microstructural Evolution on Corrosion Resistance


The obtained results of the CG and HE samples show that the plastic deformation
increased the corrosion susceptibility of the material. This may have resulted from the
grain boundary characteristics (the fraction of LAGBs and HAGBs) of the samples being
modified by the HE process. EBSD examination confirmed that HE contained a much
larger fraction of LAGBs in comparison to CG. However, the absolute number and length
of HAGBs were also greater. The dissolution of the material preferably occurred along
HAGBs, since the HAGBs indicate regions with elevated levels of excess energy [38]. This
phenomenon may have been the principal reason for the significantly lower corrosion
resistance of the HE sample.
Postcorrosion analysis of the pit morphologies showed that, for CG, the pits were rela-
tively narrow; for the samples that had undergone hydrostatic extrusion, the pits tended to
develop inside the material, creating larger networks of dissolved material. Simultaneously,
the depth of the corrosion pits decreased for samples that had been subjected to artificial
aging. As the aging conditions and hydrostatic extrusion did not change the composition
and size of the Al–Cu–Fe particles, the obtained results must be connected with differences
in the matrix composition due to the growth of precipitates. In particular, the naturally
aged HE sample exhibited more active breakdown potential than that of the peak-aged
T6 HT2 sample, because the former had a higher potential difference between the inter-
metallic particles and the matrix and thus exhibited a stronger galvanic relationship [37].
The improvement of corrosion resistance following artificial aging can be attributed to the
number of precipitates, as artificial aging significantly accelerates the precipitation effect.
That is, the matrix composition of the material changes with aging due to precipitation,
rendering it nobler and improving the corrosion resistance of the material. This may explain
why the naturally aged HE was much more prone to dissolution and more susceptible to
corrosion when compared to the artificially aged samples. Furthermore, artificial aging at
higher temperatures leads to the coarsening of large precipitates at the expense of smaller
precipitates, which in turn reduces the number of the precipitates within the microstructure.
This effect may influence the overall corrosion behavior of the material, since corrosion
is centered on specific sites in the heterogeneous microstructure, such as precipitates and
intermetallic particles. Therefore, their limited number reduced the number of preferential
nucleation sites of a corrosion attack [39].

5. Conclusions
In this paper, the effect of combined hydrostatic extrusion processing and various
thermal treatments on the corrosion and mechanical properties of AA 7075 was investigated.
Due to preserving the size of intermetallic particles that significantly influence the corrosion

144
Materials 2022, 15, 4343

resistance, other microstructural components and their influence could be examined. The
main conclusions from the present study are as follows:
• Hydrostatic extrusion with an equivalent strain of ε = 1.4 resulted in grain refinement
from 17 to 1.3 μm. A microstructure consisting of fibrous grains oriented in the <001>
and <111> directions with a fraction of LAGBs of 76% was obtained.
• As a result of thermomechanical treatment, an increase in tensile strength of up to
100 MPa was obtained when compared to the CG T6 state. Hydrostatic extrusion
caused a noticeable anisotropy of the mechanical properties, as differences in ten-
sile strength between longitudinal and transverse directions were in the range of
60–100 MPa.
• Artificial aging caused an increase in the size of MgZn2 precipitates within the mi-
crostructure, and led to improved corrosion resistance due to changes in the chemical
composition of the matrix. The differences in mechanical strength in relation to the
aging temperature were not significant due to the location of precipitates at the grain
boundaries, which were not effective in strengthening the material.
• The optimal correlation of high mechanical strength and high corrosion resistance was
obtained for the sample subjected to hydrostatic extrusion and subsequently aged at
120 ◦ C for 24 h, corresponding to a T6 state.

Author Contributions: Conceptualization, M.O. and E.U.-B.; methodology, M.O., E.U.-B. and M.K.;
validation, E.U.-B., L.Ś. and M.K.; formal analysis, M.O.; investigation, M.O., E.U.-B., P.S., K.M., M.K.
and B.A.-C.; resources, E.U.-B.; data curation, M.O., P.S., K.M. and B.A.-C.; writing—original draft
preparation, M.O., E.U.-B. and L.Ś.; writing—review and editing, M.O. and E.U.-B.; visualization,
M.O.; supervision, E.U.-B. and L.Ś.; project administration, E.U.-B.; funding acquisition, E.U.-B. and
L.Ś. All authors have read and agreed to the published version of the manuscript.
Funding: This research was funded by POB “Technologie Materiałowe” of Warsaw University of
Technology (Poland) within the Excellence Initiative: Research University (IDUB) program under the
project titled “Optimization of the corrosion resistance and mechanical properties of precipitation
strengthened 7075 aluminum alloy subjected to hydrostatic extrusion”.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: The data presented in this study are available on request from the
corresponding author.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Meng, Q.; Frankel, G.S. Effect of Cu Content on Corrosion Behavior of 7xxx Series Aluminum Alloys. J. Electrochem. Soc. 2004,
151, B271. [CrossRef]
2. Polmear, I.P. Light Alloys—From Traditional Alloys to Nanocrystals; Butterworth-Heinemann: Oxford, UK, 2006; Volume 2.
3. Osamura, K.; Kubota, O.; Protit, P.; Okuda, H.; Ochiai, S.; Fujii, K.; Kusui, J.; Yokote, T.; Kubo, K. Development of high-strength
aluminum alloys by mesoscopic structure control. Metall. Mater. Trans. A 1995, 26, 1597–1599. [CrossRef]
4. Ralston, K.D.; Birbilis, N.; Weyland, M.; Hutchinson, C.R. The effect of precipitate size on the yield strength-pitting corrosion
correlation in Al-Cu-Mg alloys. Acta Mater. 2010, 58, 5941–5948. [CrossRef]
5. Valiev, R. Nanostructuring of metals by severe plastic deformation for advanced properties. Nat. Mater. 2004, 3, 511–516.
[CrossRef]
6. Majchrowicz, K.; Pakieła, Z.; Giżyński, M.; Karny, M.; Kulczyk, M. High-cycle fatigue strength of ultrafine-grained 5483 Al-Mg
alloy at low and elevated temperature in comparison to conventional coarse-grained Al alloys. Int. J. Fatigue 2018, 106, 81–91.
[CrossRef]
7. Darban, H.; Mohammadi, B.; Djavanroodi, F. Effect of equal channel angular pressing on fracture toughness of Al-7075. Eng. Fail.
Anal. 2016, 65, 1–10. [CrossRef]
8. Wawer, K.; Lewandowska, M.; Kurzydłowski, K.J. Improvement of mechanical properties of a nanoaluminium alloy by precipitate
strengthening. Arch. Metall. Mater. 2012, 57, 877–881. [CrossRef]
9. Miyamoto, H. Corrosion of ultrafine grained materials by severe plastic deformation, an overview. Mater. Trans. 2016, 57, 559–572.
[CrossRef]

145
Materials 2022, 15, 4343

10. Miyamoto, H.; Yuasa, M.; Rifai, M.; Fujiwara, H. Corrosion behavior of severely deformed pure and single-phase materials. Mater.
Trans. 2019, 60, 1243–1255. [CrossRef]
11. Song, D.; Ma, A.B.; Jiang, J.H.; Lin, P.H.; Shi, J. Improving corrosion resistance of pure Al through ECAP. Corros. Eng. Sci. Technol.
2011, 46, 505–512. [CrossRef]
12. Son, I.J.; Nakano, H.; Oue, S.; Kobayashi, S.; Fukushima, H.; Horita, Z. Pitting corrosion resistance of anodized aluminum-copper
alloy processed by severe plastic deformation. Mater. Trans. 2008, 49, 2648–2655. [CrossRef]
13. Nakano, H.; Yamaguchi, H.; Yamada, Y.; Oue, S.; Son, I.J.; Horita, Z.; Koga, H. Effects of high-pressure torsion on the pitting
corrosion resistance of aluminum-iron alloys. Nippon Kinzoku Gakkaishi/J. Jpn. Inst. Met. 2013, 77, 543–549. [CrossRef]
14. Mahmoud, T.S. Effect of friction stir processing on electrical conductivity and corrosion resistance of AA6063-T6 Al alloy. Proc.
Inst. Mech. Eng. Part C J. Mech. Eng. Sci. 2008, 222, 1117–1123. [CrossRef]
15. Osório, W.R.; Freire, C.M.; Garcia, A. The role of macrostructural morphology and grain size on the corrosion resistance of Zn
and Al castings. Mater. Sci. Eng. A 2005, 402, 22–32. [CrossRef]
16. Lewandowska, M.; Kurzydlowski, K.J. Recent development in grain refinement by hydrostatic extrusion. J. Mater. Sci. 2008, 43,
7299–7306. [CrossRef]
17. Molak, R.M.; Paradowski, K.; Brynk, T.; Ciupinski, L.; Pakiela, Z.; Kurzydlowski, K.J. Measurement of mechanical properties in a
316L stainless steel welded joint. Int. J. Press. Vessel. Pip. 2009, 86, 43–47. [CrossRef]
18. Majchrowicz, K.; Jóźwik, P.; Chromiński, W.; Adamczyk-Cieślak, B.; Pakieła, Z. Microstructure, texture and mechanical properties
of mg-6sn alloy processed by differential speed rolling. Materials 2021, 14, 83. [CrossRef]
19. Birbilis, N.; Buchheit, R.G. Investigation and Discussion of Characteristics for Intermetallic Phases Common to Aluminum Alloys
as a Function of Solution pH. J. Electrochem. Soc. 2008, 155, C117. [CrossRef]
20. Huo, W.; Hou, L.; Cui, H.; Zhuang, L.; Zhang, J. Fine-grained AA 7075 processed by different thermo-mechanical processings.
Mater. Sci. Eng. A 2014, 618, 244–253. [CrossRef]
21. Málek, P.; Cieslar, M. The influence of processing route on the plastic deformation of Al-Zn-Mg-Cu alloys. Mater. Sci. Eng. A 2002,
324, 90–95. [CrossRef]
22. Birbilis, N.; Cavanaugh, M.K.; Buchheit, R.G. Electrochemical behavior and localized corrosion associated with Al7Cu2Fe particles
in aluminum alloy 7075-T651. Corros. Sci. 2006, 48, 4202–4215. [CrossRef]
23. Abd El Aal, M.I.; Um, H.Y.; Yoon, E.Y.; Kim, H.S. Microstructure evolution and mechanical properties of pure aluminum deformed
by equal channel angular pressing and direct extrusion in one step through an integrated die. Mater. Sci. Eng. A 2015, 625,
252–263. [CrossRef]
24. Brunner, J.G.; May, J.; Höppel, H.W.; Göken, M.; Virtanen, S. Localized corrosion of ultrafine-grained Al-Mg model alloys.
Electrochim. Acta 2010, 55, 1966–1970. [CrossRef]
25. Andreatta, F.; Terryn, H.; de Wit, J.H.W. Corrosion behaviour of different tempers of AA7075 aluminium alloy. Electrochim. Acta
2004, 49, 2851–2862. [CrossRef]
26. Sun, Q.; Yang, M.; Jiang, Y.; Lei, L.; Zhang, Y. Achieving excellent corrosion resistance properties of 7075 Al alloy via ultrasonic
surface rolling treatment. J. Alloys Compd. 2022, 911, 165009. [CrossRef]
27. Tian, W.; Li, S.; Wang, B.; Liu, J.; Yu, M. Pitting corrosion of naturally aged AA 7075 aluminum alloys with bimodal grain size.
Corros. Sci. 2016, 113, 1–16. [CrossRef]
28. Lewandowska, M. Mechanism of grain refinement in aluminium in the process of hydrostatic extrusion. Solid State Phenom. 2006,
114, 109–116. [CrossRef]
29. Chrominski, W.; Majchrowicz, K.; Lewandowska, M. Microstructural response to compression deformation of ultrafine-grained
aluminum with various microstructures. Mater. Sci. Eng. A 2019, 763, 138184. [CrossRef]
30. Chrominski, W.; Lewandowska, M. Precipitation phenomena in ultrafine grained Al-Mg-Si alloy with heterogeneous microstruc-
ture. Acta Mater. 2016, 103, 547–557. [CrossRef]
31. Pachla, W.; Kulczyk, M.; Smalc-Koziorowska, J.; Wróblewska, M.; Skiba, J.; Przybysz, S.; Przybysz, M. Mechanical properties and
microstructure of ultrafine grained commercial purity aluminium prepared by cryo-hydrostatic extrusion. Mater. Sci. Eng. A 2017,
695, 178–192. [CrossRef]
32. Ura-Bińczyk, E. Improvement of Pitting-Corrosion Resistance of Ultrafine-Grained 7475 Al Alloy by Aging. Materials 2022, 15,
360. [CrossRef] [PubMed]
33. Available online: http://www.matweb.com/search/DataSheet.aspx?MatGUID=4f19a42be94546b686bbf43f79c51b7d&ckck=1
(accessed on 29 January 2022).
34. Hansen, N. Hall–Petch relation and boundary strengthening. Scr. Mater. 2004, 51, 801–806. [CrossRef]
35. Lewandowska, M.; Wawer, K.; Kozikowski, P.; Ohnuma, M.; Kurzydlowski, K.J. Precipitation in a nanograined 7475 aluminium
alloy—Processing, properties and nanoanalysis. Adv. Eng. Mater. 2014, 16, 482–485. [CrossRef]
36. Przybysz, S.; Kulczyk, M.; Pachla, W.; Skiba, J.; Wróblewska, M.; Mizera, J.; Moszczynska, D. Anisotropy of mechanical and
structural properties in aa 6060 aluminum alloy following hydrostatic extrusion process. Bull. Pol. Acad. Sci. Tech. Sci. 2019, 67,
709–717. [CrossRef]
37. Hansen, N. Boundary strengthening in undeformed and deformed polycrystals. Mater. Sci. Eng. A 2005, 409, 39–45. [CrossRef]

146
Materials 2022, 15, 4343

38. Orłowska, M.; Ura-Bińczyk, E.; Olejnik, L.; Lewandowska, M. The effect of grain size and grain boundary misorientation on the
corrosion resistance of commercially pure aluminium. Corros. Sci. 2019, 148, 57–70. [CrossRef]
39. Murer, N.; Oltra, R.; Vuillemin, B.; Néel, O. Numerical modelling of the galvanic coupling in aluminium alloys: A discussion on
the application of local probe techniques. Corros. Sci. 2010, 52, 130–139. [CrossRef]

147
materials
Article
Fatigue Fracture Analysis on 2524 Aluminum Alloy with the
Influence of Creep-Aging Forming Processes
Liyong Ma 1,† , Chi Liu 2,† , Minglei Ma 3 , Zhanying Wang 1,4, *, Donghao Wu 1 , Lijuan Liu 1 and Mingxing Song 1, *

1 School of Mechanical Engineering, Hebei University of Architecture, Zhangjiakou 075031, China;


[email protected] (L.M.); [email protected] (D.W.); [email protected] (L.L.)
2 School of Mechanical and Electrical Engineering, Changsha University, Changsha 410199, China;
[email protected]
3 Zhangjiakou Cigarette Factory Co., Ltd., Zhangjiakou 075001, China; [email protected]
4 Zhangjiakou Special Equipment Intelligent Monitoring Operation and Maintenance Technology Innovation
Center, Zhangjiakou 075031, China
* Correspondence: [email protected] (Z.W.); [email protected] (M.S.);
Tel.: +86-313-418-7755 (Z.W.); +86-313-418-7749 (M.S.)
† These authors contributed equally to this work.

Abstract: The different creep-aging forming processes of 2524 aluminum alloy were taken as the
research object, and the effects of creep-aging temperature and creep stress on the fatigue-crack
propagation properties of the alloy were studied. The research results showed the following under
the same sintering time of 9 h, at creep-aging temperatures of 100 ◦ C, 130 ◦ C, 160 ◦ C, and 180 ◦ C,
respectively, with an increase in creep-aging temperature: the fatigue-crack propagation rate was
promoted, the spacing of fatigue striations increased, and the sizes of dimples decreased while the
number was enlarged; this proves that the fatigue property of the alloy was weakened. Compared
with the specimens with creep deformation radii of 1000 mm and 1500 mm, the creep deformation
Citation: Ma, L.; Liu, C.; Ma, M.; stress was the smallest when the forming radius was 1800 mm, with a higher threshold value of
Wang, Z.; Wu, D.; Liu, L.; Song, M. fatigue-crack growth in the near-threshold region of fatigue-crack propagation (ΔK ≤ 8 MPa·m1/2 ).
Fatigue Fracture Analysis on 2524
Under the same fatigue cycle, the specimens under the action of larger creep stress endured a longer
Aluminum Alloy with the Influence
fatigue stable-propagation time and a faster fracture speed. Comparing the effect of creep-aging
of Creep-Aging Forming Processes.
temperature and creep stress, the creep-aging temperature plays a dominant role in the fatigue-crack
Materials 2022, 15, 3244. https://
propagation of creep-aged 2524 aluminum alloy.
doi.org/10.3390/ma15093244

Academic Editors: Daolun Chen, Keywords: creep-aging forming; creep-aging temperature; creep stress; 2524 aluminum alloy; fatigue-
Lucjan Śnieżek and Jaroslaw crack propagation rate; fatigue fracture morphology
Galkiewicz

Received: 31 March 2022


Accepted: 27 April 2022
Published: 30 April 2022 1. Introduction
Publisher’s Note: MDPI stays neutral
2524 aluminum alloy is a new type of high-strength aluminum alloy for aviation [1]
with regard to jurisdictional claims in that is mainly used for wing skins [2]. During service, it is subjected to loads in complex
published maps and institutional affil- environments, so its fatigue performance is particularly critical [3]. Compared with alu-
iations. minum alloys such as 2024, 2124 and 2224, 2524 aluminum alloy has higher strength and
better fatigue properties [4,5]. The fracture toughness is increased by 15~20%, the fatigue
resistance doubles, and the fatigue life is increased by 27~45% [6,7], which is well suited to
the requirements of modern aircraft design for material damage tolerance [8]. Meanwhile,
Copyright: © 2022 by the authors. with the maturity of the creep-aging forming process, combined with the good hot-working
Licensee MDPI, Basel, Switzerland. properties of 2524 aluminum alloy, the application prospects of 2524 aluminum alloy in the
This article is an open access article future aerospace industry are broad [9].
distributed under the terms and During the creep-aging forming process, the microstructure and evolution process of
conditions of the Creative Commons
the material are complex, and the factors that affect the fatigue properties of the creep-aging
Attribution (CC BY) license (https://
forming material also become complicated. Pitcher et al. [10] studied creep-aging-formed
creativecommons.org/licenses/by/
2024A, 8090, and 7449 aluminum alloys from the two aspects of springback and damage
4.0/).

Materials 2022, 15, 3244. https://doi.org/10.3390/ma15093244 149 https://www.mdpi.com/journal/materials


Materials 2022, 15, 3244

tolerance, and conducted a large number of fatigue-crack propagation rate experiments. The
result showed that 2024A and 8090 had sufficient damage resistance after creep deformation,
which can be useful for making lower wing skins. Rafiq A. Siddiqui et al. [11] studied the
effects of age temperature and duration on the fatigue properties of 6063 aluminum alloys.
The study found the vacancy diffusion mechanism played a significant role in the formation
of Guinier–Preston (GP) zones, in which the solid solution is precipitated due to uniform
nucleation during the dissolution process, and a fine quasi-stable phase is precipitated.
The structure of this stable phase was similar to that of the main matrix, and the two were
coherent with each other, thus disturbing the regularity of the lattice, leading to the increase
in the fatigue defects of the alloy. Brav G.H. et al. [12] studied the effects of different aging
processes on the crack propagation rate of the 2024-T351 alloy and Al–Cu–Mg–Li alloy. The
results showed that artificial aging increases the strength of the alloy, but it also reduces
the crack propagation resistance of the alloy. In addition, with the prolongation of artificial
aging duration, the crack propagation rate increased monotonically. The authors believed
that the reduction in crack propagation resistance was related to the precipitation of the T1
phase and S phase in the alloy; however, the influence mechanism of the aging process on
the precipitation phase was not discussed, nor was the effect of fine precipitation on the
crack-tip slip-mode. Sarioglu F. and Orhaner F.Ö. [13] researched the fatigue property of
the 2024-T3 aluminum alloy after aging at a low temperature of 130 ◦ C for 100 h/1000 h.
The results showed that long-term aging eliminated the difference in the crack propagation
rate of alloys in the sampling directions (L-T, T-L, 60◦ ). Burba et al. [14] found that the
minimum fatigue life of the alloy was mainly affected by the density of the θ precipitation
phase, but had little effect on the average fatigue life. Yang [15] conducted a lot of research
on the effect of the creep-aging forming process on the high-cycle fatigue properties of
Al-Zn-Mg series-7075 high-strength aluminum alloys. The study found that the increase
in creep-aging temperature and time has a positive effect on the improvement of alloy
fatigue properties. The effect of preloading stress will reduce the fatigue resistance of the
alloy. However, due to the differences in microstructure and chemical composition, the
conclusions of this study were not applicable to 2524 aluminum alloys.
Xu et al. [16] studied the effect of tensile pre-strain before creep-aging forming on
the mechanical properties of 2524 aluminum alloy using a constant-stress creep-aging test.
they found that the magnitude of creep strain was greatly increased with the increase in
pre-strain and that the pre-strain improved the formability, mechanical properties and
microstructures of 2524. In addition, Xu et al. [17] also conducted research on the tension
and compression creep-aging behaviors of Al-Cu-Mg alloy. They demonstrated that the
creep strains under tensile stresses were larger than the creep strains under compressive
stresses. Meanwhile, the formation of the S phase in the aluminum matrix was caused by the
compressive stress, inhibiting the precipitates in the grain boundary, which was beneficial
to the hardness of the compression creep-aged alloy. Liu et al. [18] studied the effects
of creep-aging and artificial aging on the fatigue-crack propagation of 2524 aluminum
alloy. They found that the prolonged aging time resulted in excessive precipitation of the
needle-like S phase, changed the dislocation slip mode, reduced the reversibility of the slip,
and accelerated the accumulation of fatigue damage. In addition, creep stress accelerated
the aging precipitation process of the alloy. Compared with artificial aging, under the
same aging time, the size of the precipitates in the creep-aged alloy was larger and the
yield strength and hardness were increased, but the fatigue resistance was decreased. In
our previous work [19], we studied the effects of creep time on the microstructural and
mechanical properties, as well as the fatigue-crack propagation, in 2524 aluminum alloy.
TEM was employed for the observation of precipitation and dislocation.
Most scholars have focused on the relationship between the creep/stress relaxation
mechanism and the material deformation. However, the relationship between creep-
aging temperature, creep stress, and the fatigue properties of materials has not been
comprehensively and systematically studied.

150
Materials 2022, 15, 3244

This paper will focus on analyzing the macro-microscopic characteristics of the fatigue-
crack fractures of creep-formed specimens, and analyze the fracture mode, crack propa-
gation path and fracture mechanism of creep-formed components. The fatigue property
of creep-formed alloys at different creep-aging temperatures and creep stresses will be
explored from the micro-morphological features.

2. Experimental Processes
2.1. Materials
In this study, a 2524 aluminum alloy sheet for aviation (the thickness was 3.5 mm) was
selected for research. After the alloy was solution-treated, it was cold-machined, and then
naturally aged to a stable state. The main chemical composition of 2524 aluminum alloy is
shown in Table 1 [9,16].

Table 1. Chemical composition of 2524 aluminum alloy (mass fraction: wt%).

Cu Mg Mn Fe Zn Si Ti Cr Al
4.62 1.32 0.57 0.035 0.004 0.025 0.02 0.001 Bal.

2.2. Creep-Aging Forming


The creep-aging forming experiment specimen and mold are shown in Figure 1a, and
the size of the experiment sheet was 360 mm × 220 mm. The 2524 aluminum alloy sheet
was placed in the center of the mold, then the mold and sheet were wrapped with air
felt and a vacuum bag; then, the bag was sealed with heat-resistant glue. The bag was
gradually evacuated to a vacuum, and a negative pressure of 0.1 MPa was maintained.
The sheet was elastically deformed under uniform load until it was completely fitted with
the mold surface. The deformed sheet and mold were put into an autoclave (Figure 1b),
with a heating rate of 1.5 ◦ C/min. In addition, the creep-aging treatment was carried out
according to the set creep-aging temperature, and the vacuum bag was kept in a sealed
state during the creep-forming process. After the set aging time was reached, the load and
temperature were removed, and the components were cooled in a furnace, obtaining the
final desired shape.

(a) (b)
Figure 1. Creep-aging forming mold and tooling: (a) mold and sheet; (b) autoclave.

(1) To study the effect of creep-aging temperature on fatigue property, creep deforma-
tion experiments were carried out at creep-aging temperatures of 100 ◦ C, 130 ◦ C, 160 ◦ C
and 180 ◦ C, respectively, and the creep-aging time of all specimens was 9 h.
(2) To study the effect of creep stress, the radii of curvature ρ were 1000 mm, 1500 mm
and 1800 mm, respectively. The larger the radius ρ, the smaller the creep stress. The
creep-aging temperature was 160 ◦ C, and the creep time was 9 h.

151
Materials 2022, 15, 3244

2.3. Fatigue Experiment


The fatigue-crack propagation experiment was carried out at room temperature. The
sample preparation and stress loading method of the fatigue-crack propagation specimen
were designed according to ASTM-E647 [20]. The experimental equipment was the MTS-810
fatigue experimenting machine made in the U.S. The fatigue-crack propagation experiment
utilized standard compact tensile specimens (CT specimens), and the specimens were cut
using a wire electric discharge. The size of the sample is shown in Figure 2. Before the
fatigue-crack propagation experiment, a crack of 2.75 mm was prefabricated using the
pre-crack module in the experiment system. The experiment adopted sine wave loading,
with the maximum loading of 2400 N. The stress ratio R of 0.1~0.5 (R = Pmin /Pmax ), the
loading frequency f of 10 Hz, and the crack length a were detected by the compliance
control method (COD).

Figure 2. Schematic diagram of CT specimen for fatigue-crack propagation experiment.

The fatigue-crack propagation data obtained in the experiment were processed using
the seven-point incremental polynomial method. For data point i and its front 3 points
and back 3 points—a total of 7 continuous data points—the quadratic polynomial was
employed to perform local fitting and derivation, and the fitting values of the fatigue-crack
propagation rate were obtained using Equation (1):

Ni − C1 N − C1 2
ai = b0 + b1 + b2 i (1)
C2 C2

where Ni is the number of cycles, ai is the fitting crack length value; b0 , b1 and b2 are
the regression parameters determined according to the minimum squared deviation be-
tween the observed value of the crack length and the fitting value; C1 = (Ni−3 − Ni+3 )/2;
C2 = (Ni+3 − Ni−3 )/2; and −1 ≤ (Ni−3 − C1 )/C2 ≤ 1. By derivation of Equation (1), the
crack propagation rate at Ni can be obtained:


da b 2b ( N − C )
= 1 + 2 i2 1 (2)
dN ai C2 C2

For CT specimens with type-I open cracks, the range of stress intensity factor ΔK at
the crack tip can be calculated using Equation (3) [19,20]:

ΔP (2 + α )
ΔK = √ · (0.886 + 4.64α − 13.32α2 + 14.72α3 − 5.6α4 ) (3)
B W (1 − α)3/2

152
Materials 2022, 15, 3244

where ΔP is the force value range; ΔP = Pmax − Pmin , Pmax is the maximum loading force;
Pmin is the minimum loading force; α = a/W, a is the crack length; B is the width of the
specimen (B = 5 mm); and W is the width of the specimen (W = 44 mm).

3. Result and Discussion


3.1. Effect of Creep-Aging Temperature on Fatigue-Crack Propagation
3.1.1. Effect of Creep-Aging Temperature on Fatigue-Crack Growth Rate
The creep-aging temperatures were 100 ◦ C, 130 ◦ C, 160 ◦ C and 180 ◦ C, respectively,
and the aging time was the same, i.e., 9 h. The fatigue-crack propagation rate experiment
was carried out. The da/dN − ΔK curve is shown in Figure 3. In Figure 3a, the crack
propagation curves of the creep-formed alloy at 100 ◦ C and 130 ◦ C coincide with those
of 2524 aluminum alloy without creep forming, indicating a fatigue property with no
significant change at lower creep-aging temperatures (<130 ◦ C), because of insufficient
effective deformation of the alloy at lower temperatures. Additionally, the da/dN − ΔK
curve of 2524-T3 aluminum alloy without creep forming was taken as the comparison data.
2524-T3 is a kind of alloy obtained through cold-working of 2524 aluminum alloy after
solution treatment, then stabilization by natural aging.

(a)

(b)
Figure 3. Fatigue-crack propagation rate da/dN-ΔK curves of 2524 aluminum alloy at different
creep-aging temperatures: (a) 100 ◦ C and 130 ◦ C; (b) 160 ◦ C and 180 ◦ C.

153
Materials 2022, 15, 3244

Nevertheless, under the age temperatures of 160 ◦ C and 180 ◦ C, the crack propa-
gation rate curves show significant differences (Figure 3b). In the near-threshold region
(ΔK ≤ 8 MPa·m1/2 ), the crack propagation resistance of the alloy is significantly reduced
under high-temperature creep-aging forming. Under the same ΔK of 6 MPa·m1/2 , when
the creep-aging temperature is 160 ◦ C, da/dN = 5.62 × 10−6 mm/cycles−1 , but when
the creep-aging temperature is 180 ◦ C, da/dN is increased to 9.11 × 10−6 mm/cycles−1 ,
showing that the crack propagation resistance of 2524 aluminum alloy is decreased to a
certain extent under high-temperature creep-aging forming. In addition, with the increase
in ΔK, the difference in crack propagation rate gradually decreases. According to the curve,
it is delineated in the Paris region of 10 MPa·m1/2 ≤ ΔK ≤ 30 MPa·m1/2 , and the curve
shows an obvious linear relationship.
The straight part of the da/dN − ΔK curve in the double logarithmic coordinate is
fitted by the Paris equation, and the corresponding fitting constants C, n and the fatigue-
crack propagation rate under the same ΔK are shown in Table 2. The error value is
2.16%. The values of n are close, ranging from 2.7 to 3.0, indicating that the creep-aging
temperature has little effect on the crack propagation rate of 2524 aluminum alloy in the
medium and high stress range. When ΔK exceeds 30 MPa·m1/2 , the curve of da/dN − ΔK
has an obvious turning point, and the crack propagation rate da/dN increases rapidly from
10−3 mm/cycles to 0.1 mm/cycles until an instability fracture occurs.

Table 2. Paris fitting parameters C and n at different creep-aging temperatures.

da/dN = CΔKn /(mm·cycle−1 )


Aging Status C n ΔK = 16
ΔK = 8 ΔK = 12
(MPa·m1/2 )
100 ◦ C/9 h 1.77 × 10−7 2.99 3.51 × 10−5 1.52 × 10−4 4.31 × 10−4
130 ◦ C/9 h 1.77 × 10−7 2.99 4.15 × 10−5 1.75 × 10−4 3.97 × 10−4
160 ◦ C/9 h 4.32 × 10−7 2.67 4.52 × 10− 5 2.25 × 10−4 4.78 × 10−4
180 ◦ C/9 h 2.81 × 10−7 2.80 5.02 × 10−5 2.13 × 10−4 5.74 × 10−4

3.1.2. Fracture Morphologies at Different Creep-Aging Temperatures


(1) Fatigue-crack propagation zone
The creep-forming temperature has a decisive effect on the solute precipitation in the
supersaturated state of 2524 aluminum alloy [21]. The increase in the age temperature
enhances the atomic activity in the alloy, resulting in a quick increase in the precipitation
rate of the precipitation phase [22]. Therefore, the nucleation, growth and enrichment of
the precipitates during the aging process will be affected by the age temperature, which
will lead to changes in the fatigue properties of the material. Figure 4 shows the fracture
morphology of 2524 aluminum alloy at crack length a = 5 mm after aging at different
creep-aging temperatures (100 ◦ C, 130 ◦ C, 160 ◦ C, 180 ◦ C) for 9 h. The corresponding ΔK at
this time is about 15 MPa·m1/2 , and the crack is in the stable propagating stage.
In Figure 4, the fatigue sections of the specimens are all flat, and the relative torsion of
the cleavage planes in adjacent grains makes the crack propagate along many transgranular
planes, forming smooth and flat sections, i.e., large fatigue platforms. These fatigue
platforms are connected by the tearing edge, indicated by mark 2 in the figure. The tearing
edge is deflected by an angle of 10◦ ~40◦ relative to the main crack propagation direction,
and the fracture shows the characteristics of a ductile transgranular fracture. There are
also many micro-pores (mark 1) distributed on the cross-section, which originate from the
tiny plastic deformations confined around the coarser second-phase particles during the
fatigue process. Meanwhile, secondary cracks (mark 4) approximately perpendicular to
the direction of the main crack-propagation plane are also observed, which propagate into
the material. There are more fracture cleavage steps in the fracture morphology under
high-temperature aging of 180 ◦ C—9 h, and the fracture surface is rough, with part of the
fracture morphology even showing slight brittle fracture characteristics. From Figure 4d, it

154
Materials 2022, 15, 3244

can be seen that the obvious small fatigue steps are connected by the shear edges at mark 3;
moreover, the heights are different, indicating that the aging precipitation and hardening
rate of 2524 aluminum alloy are promoted due to high-temperature aging. The alloy enters
the overaging state in advance, and the ductility decreases significantly, which adversely
affects the fatigue property and toughness of the alloy.




 

 
 

(a) (b)


  

    




(c) (d)
Figure 4. SEM images of stable fatigue-crack propagating zone at different creep-aging temperatures
(a = 5 mm): (a) 100 ◦ C—9 h; (b) 130 ◦ C—9 h; (c) 160 ◦ C—9 h; (d) 180 ◦ C—9 h. Marks in the figure:
1—micropore; 2—tear edge; 3—shear edge; 4—secondary crack.

The micro-morphology of the corresponding position in Figure 4 is magnified to


20,000 times for observation, and fatigue striations can be observed, as shown in Figure 5.
There are obvious differences in the fatigue striation spacing of the specimens under
different creep-aging temperatures. Since the spacing of only one fatigue striation is too
small to measure, and the measured value of only one fatigue striation often brings in error,
the spacing of five fatigue striations was measured for precision in mirroring the effect of
creep temperature and stress on the fatigue-crack propagation. Under the low-temperature
aging of 100 ◦ C—9 h and 130 ◦ C—9 h, the spacing of five fatigue striations are relatively
small, at 1.22 μm and 2.41 μm, respectively, and the striation morphology is very clear and
regular. With the increase in creep-aging temperature, under the same crack length, the
striation spacing increases, and the appearance is rough. Under the high temperature aging
of 160 ◦ C—9 h and 180 ◦ C—9 h, the average striation spacing is 3.52 μm and 3.86 μm. The
difference in fatigue striation spacing reflects the increase in the size and volume fraction
of the alloy precipitates in the peak or overaging state, which increases the strength of
the material, but reduces its elongation and increases its brittleness, which makes fatigue
streaks less likely to occur.

155
Materials 2022, 15, 3244

(a) (b)

ȝP

(c) (d)
Figure 5. SEM images of fatigue striations in the stable propagation zone at different creep-aging
temperatures (a = 5 mm): (a) 100 ◦ C—9 h; (b) 130 ◦ C—9 h; (c) 160 ◦ C—9 h; (d) 180 ◦ C—9 h.

At different creep-aging temperatures, when the precipitated strengthening phase


maintains a coherent or semi-coherent relationship with the matrix, it is generally believed
that dislocations can cut through the precipitated phase [23], and then plane slip occurs,
resulting in uneven deformation in local areas. As the slip plane continues to expand,
the crack propagation path may deflect, kink, and bifurcate, reducing the rate of crack
propagation. When the strengthening phase is incoherent with the matrix, the dislocations
bypass the precipitation phase, and the deformation in the local area is relatively uniform;
this reduces the possibility of deflection, kink and bifurcation during crack propagation,
thereby increasing the rate of crack propagation. Therefore, for samples with an age
temperature of 100 ◦ C and 130 ◦ C, the precipitates maintain a coherent or semi-coherent
relationship with the matrix, and the fatigue-crack propagation rate is slow. However,
when the age temperature is 180 ◦ C, the precipitation phase of the material grows up and
breaks away from the semi-coherent relationship with the matrix, so the crack propagation
rate of the sample is at its fastest, as shown in Figure 3b.
(2) Fatigue fracture zone
Figure 6 shows the fracture morphologies of 2524 aluminum alloy in the Fatigue
fracture zone (a = 25 mm) after aging at different creep-aging temperatures (100 ◦ C, 130 ◦ C,
160 ◦ C, 180 ◦ C) for 9 h. From the fracture surface of the specimen with the creep-aging
temperature of 130 ◦ C (see Figure 6b), the large pits caused by the debonding between the
precipitation phase and the interface of the aluminum matrix are surrounded by smaller
pits. Meanwhile, the large pits are not connected to each other during the fracture process,
and there are obvious tearing edges along the small pits, which indicates that the specimen
also has good plastic deformation ability. the higher the creep-aging temperature, the larger
the dimple size. However, there is no obvious tearing edge, and the fracture morphology

156
Materials 2022, 15, 3244

is also flatter. The high-temperature creep accelerates the nucleation and growth of the
precipitation phase at 180 ◦ C. During the tearing fracture process, the precipitation phase
will hinder the dislocation slip due to the incompatibility of the precipitation phase with
the aluminum matrix [24], resulting in stress concentration. When the stress concentration
exceeds the critical value that the material can withstand, the interface between the pre-
cipitate and the aluminum matrix is debonded [25], or the precipitate is fractured, which
reduces the plastic deformation ability of the aluminum alloy. With the continuous increase
in plastic strain, the interface debonding between the precipitation phase and the aluminum
matrix causes the aggregation of large pits. When the effective bearing area is reduced to a
critical value, the aluminum matrix will break rapidly.

Small
dimples

Tearing edge
Large
dimples

(a) (b)

(c) (d)
Figure 6. SEM images of fatigue-crack transient region at different creep-aging temperatures:
(a) 100 ◦ C—9 h; (b) 130 ◦ C—9 h; (c) 160 ◦ C—9 h; (d) 180 ◦ C—9 h.

Under the impact of the creep-aging temperature and the critical size, the enrichment
speed of the precipitates are different, and the type, density and length of the precipitates
also change [19,26]. In addition, 2524 aluminum alloys under different creep-aging temper-
ature have differences in microstructures and mechanical properties such as yield strength,
hardness, and elongation [26–28]. Therefore, by controlling the creep-aging temperature, a
2524 aluminum alloy with good plasticity and fatigue properties can be obtained.

3.2. Effect of Creep-Aging Stress on Fatigue-Crack Propagation


3.2.1. Effect of Creep Stress on Fatigue-Crack Propagation Rate
The fatigue-crack propagation rate curves of 2524 aluminum alloy after forming under
different creep stresses are shown in Figure 7. In the near-threshold region of fatigue-crack
propagation (ΔK ≤ 8 MPa·m1/2 ), the fatigue-crack propagation threshold of 2524 aluminum
alloy with a forming radius of 1800 mm is higher than that at 1500 mm and 1000 mm,

157
Materials 2022, 15, 3244

and its propagation rate is also lower than alloys in the other two states. However, the
propagation rate with the forming radius of 1000 mm and 1500 mm are relatively close.
This shows that the increase in creep stress also reduces the crack propagation resistance of
the alloy, but the effect is not as significant as creep-aging temperature. When 8 MPa·m1/2
≤ ΔK ≤ 20 MPa·m1/2 , the crack propagation enters the stable expansion stage, and the
curve shows an obvious linear relationship. When ΔK exceeds about 18 MPa·m1/2 , the
propagation rate goes up sharply and da/dN rapidly expands until fracture occurs.

Figure 7. Fatigue-crack propagation rate curves of 2524 aluminum alloy under different creep stresses.

The linear part in Figure 7 is fitted according to the Paris equation, and the correspond-
ing fitting constants C, n and the fatigue-crack propagation rate under the same ΔK are
shown in Table 3. The error value is 4.65%. From the fitting results, the values of the expo-
nents n are very close, ranging from 2.7 to 3.0. In the medium stress region and high stress
region, the propagation rate of the formed specimens at a radius of 1000 mm is the fastest,
indicating that the increase in creep stress reduces the fatigue-crack propagation resistance.

Table 3. Paris fitting parameters C and n of 2524 aluminum alloy under different creep stresses.

da/dN = CΔKn /(mm·cycle−1 )


ρ C n ΔK = 21
ΔK = 7 ΔK = 12 ΔK = 16
(MPa·m1/2 )
1.77 × 2.90 × 2.25 × 4.75 × 1.12 ×
1000 mm 2.99
10−7 10−5 10−4 10−4 10− 3
4.32 × 2.24 × 2.23 × 4.78 × 9.67 ×
1500 mm 2.67
10−7 10− 5 10−4 10−4 10−4
2.81 × 1.21 × 2.95 × 4.10 × 6.83 ×
1800 mm 2.80
10−7 10−5 10−4 10−4 10−4

On the one hand, this is because the alloy formed under the radius of 1000 mm
has the largest bending deformation and the highest dislocation density contained in the
crystal [29]. The large-scale dislocation accumulation and entanglement lead to the work-
hardening of the alloy itself. On the other hand, these dislocations provide a large number
of nucleation sites, and meanwhile, facilitate the short-circuit expansion of solute atoms in
the alloy matrix, promote the precipitation and coarsening of the precipitation phase [30],
and indirectly improve the yield strength. In addition, the effect of alloy yield strength on
the fatigue-crack propagation rate is mainly reflected in the size of the plastic zone at the

158
Materials 2022, 15, 3244

crack tip. The plastic zone at the crack tip has the following relationship with the yield
strength [31]:


ΔK2 3
rp ( α ) = 2 2
sin2 α + (1 − 2μ)2 (1 + cos α) (4)
4π (σ0.2 )
where rp is the radius vector; ΔK is the stress intensity factor; σ0.2 is the yield stress; α
is the polar angle; and μ is the Poisson’s ratio of 2524 aluminum alloy. Under the same
stress intensity factor amplitude ΔK, the size rp of the plastic zone at the crack tip is
inversely proportional to the yield strength σ0.2 . The larger the rp , the more energy is
absorbed under each cyclic load, and the better the fatigue damage resistance of the alloy
(i.e., the increase in the yield strength reduces the crack propagation resistance of the alloy).
Therefore, the increase in creep stress has an adverse effect on the fatigue performance of
2524 aluminum alloy.
In different creep-aging specimens, the larger the aging stress, the lower the apparent
activation energy, and the lower the resistance encountered by the movement of disloca-
tions [12]. Meanwhile, a large number of dislocations brought by pre-deformation provide
a large number of mobile dislocations for creep, and also promote the nucleation of the
second phase [32]. Therefore, when the creep specimen is creep formed under larger stress,
the precipitation and growth rate of the precipitates are also faster, but the density of the
precipitates declines. The enhancement of the pinning effect of precipitation relative to
dislocations improves the plastic deformation resistance, and the damage caused by the
same fatigue cyclic load is smaller.
The increase in creep stress also reduces the crack propagation resistance relative to
unstressed aging. However, compared with the effect of creep-aging temperature, the effect
of creep stress is not as significant as that of creep temperature.

3.2.2. Fracture Morphologies at Different Creep Stresses


(1) Fatigue-crack propagation zone
Figure 8 shows the fatigue striation spacing under the action of different creep cur-
vature radii of 1000 mm, 1500 mm and 1800 mm, respectively. The widths of the five
fatigue striations under different creep curvature radii are significantly different, at 2.27 μm,
2.66 μm and 2.93 μm, respectively. Therefore, in the stable propagation stage with the stress
intensity factor ΔK ranging from 10 to 25 MPa MPa·m1/2 , the fatigue striation spacing of
the specimen with a smaller creep radius of curvature (i.e., a larger creep stress) is larger
than that of the specimen with a larger radius of curvature; moreover, the crack propagation
rate is faster, and the fatigue resistance is lower.
For 2524 aluminum alloy under stress aging, the preferential growth orientation of
the precipitates is sensitive to the applied stress. The applied stress field changes the
degree of mismatch between the matrix and the precipitates, which causes the elastic
distortion field and the elastic energy of the coherent precipitates to change, thus affecting
the precipitation and evolution of the precipitates [33]. Larger aging stress results in lower
apparent activation energy, less resistance encountered by the movement of dislocations,
and a large number of dislocations caused by pre-deformation; these provide a large
number of movable dislocations for creep, and promote the nucleation of the second
phase [34]. Therefore, when the creep specimen creeps under the larger stress, the damage
caused by the same fatigue cyclic load to the specimen is smaller; however, after the alloy
enters the over-aging stage, the fatigue property declines significantly. This is because the
precipitation and propagation rate of the precipitation phase goes up with the increase
in stress [19], while the density of the precipitation phase goes down with an increase in
stress, which makes the plasticity of the aluminum alloy stronger. While the fatigue life of
the specimen decreases, the fatigue striation spacing becomes wider.

159
Materials 2022, 15, 3244

(a) (b)

2.93ΐm

(c)

Figure 8. SEM images of fatigue striations under different curvature radii ρ: (a) ρ = 1000 mm;
(b) ρ = 1500 mm; (c) ρ = 1800 mm.

(2) Fatigue fracture zone


Figure 9 illustrates the SEM images of the fracture zone under different creep stresses,
and the curvature radii are 1000 mm, 1500 mm and 1800 mm, respectively. With the
increase in creep stress, the fracture morphologies are slightly different; dimples and
fractured second-phase particles are clearly seen. In the sample with a curvature radius
of 1000 mm, the fracture contains dimples and cleavage fracture morphology. Comparing
the fatigue properties of the specimens treated by creep-aging under different curvature
radii, under the same fatigue cycle, the specimen with smaller curvature radius has a
larger area of the fatigue-crack stable-growth zone, and a smaller area of fracture zone; this
demonstrates that the specimen under the action of larger creep stress has longer fatigue
stability propagation time and faster instantaneous fracture speed.
The applied creep stress changes the precipitation process of the precipitation phase.
Under the action of high stress, the S phase is more likely to precipitate and grow. During
creep, dislocation density increases and becomes entangled, affecting the dislocation slip.
As the creep deformation increases, the dislocations in the grains are rearranged, and the
entangled dislocations drive the formation of subgrains [35]. The dislocations entangled
with high density form the unit cell-walls of the subgrains, and the dislocation density in
the subgrains is low. In addition, the creep stress breaks the balance of the precipitates and
changes the precipitation process, and the growth of the precipitates hinders the movement
of grain boundaries and dislocations, improving the properties of the aluminum alloy.

160
Materials 2022, 15, 3244

Figure 9. SEM images of the instantaneous break area under different curvature radii ρ:
(a) ρ = 1000 mm; (b) ρ = 1500 mm; (c) ρ = 1800 mm.

4. Conclusions
(1) With the same ΔK, the crack propagation rate increases with the increase in creep-
aging temperature. With the increase in ΔK, the difference in crack propagation rate
gradually decreases.
(2) Under the same crack length, with the increase in creep-aging temperature, the
spacing of fatigue striations increases and the size of dimples decreases, while the number
of dimples increases, and the fatigue resistance of the alloy decreases.
(3) In the near-threshold region of fatigue-crack propagation (ΔK ≤ 8 MPa·m1/2 ),
the fatigue-crack propagation threshold of 2524 aluminum alloy with a forming radius of
1800 mm is higher than that of alloys at 1500 mm and 1000 mm.
(4) Under the same fatigue cycle, the specimens under the action of larger creep
stress have longer fatigue stable-propagation time and a faster transient fracture speed.
However, compared with the effect of creep-aging temperature, the effect of creep stress is
not as significant as that of creep temperature. Hence, the creep-aging temperature plays a
dominant role in the fatigue-crack propagation of creep-aged 2524 aluminum alloy.

Author Contributions: Conceptualization, L.M., C.L. and M.S.; methodology, L.M. and M.S.; soft-
ware, C.L. and M.M.; validation, L.M., C.L. and M.M.; formal analysis, D.W. and L.L.; investigation,
L.M., C.L., M.M., Z.W., D.W. and L.L.; resources, L.M. and C.L.; data curation, Z.W.; writing—original
draft preparation, L.M., C.L., M.M. and Z.W.; writing—review and editing, M.M., Z.W., and L.L.; su-
pervision, Z.W. and M.S.; project administration, L.M., C.L., Z.W. and L.L.; funding acquisition, L.M.,
C.L., Z.W., and L.L. All authors have read and agreed to the published version of the manuscript.
Funding: This research was funded by the Hebei Province Technology Innovation Guidance Program
Project: Science and Technology Winter Olympics Special, grant number 20475501D; the Basic

161
Materials 2022, 15, 3244

Scientific Research Business Project of Hebei University of Architecture, grant numbers 2022QNJS02
and 2021QNJS08; and the Changsha Municipal Natural Science Foundation, grant number kq2007085.
Data Availability Statement: The data presented in this study are available on request from the
corresponding authors.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Li, S.B.; Li, X.; Liang, W.; Liu, Y.L.; Yan, H.Z.; Liu, C. Effects of laser shock peening on fatigue crack growth rate and fracture
properties of AA2524 aluminum alloy. J. Cent. South Univ. 2022, 29, 848–859. [CrossRef]
2. Song, M.; Wu, L.; Liu, J.; Hu, Y. Effects of laser cladding on crack resistance improvement for aluminum alloy used in aircraft skin.
Opt. Laser Technol. 2021, 133, 106531. [CrossRef]
3. Mason, C.J.T.; Avery, D.Z.; Phillips, B.J.; Jordon, J.B.; Allison, P.G. Strain Rate Dependent Plasticity Model for Precipitate Hardened
Aerospace Aluminum Alloy Produced with Solid-State Additive Manufacturing. J. Dyn. Behav. Mater. 2021. [CrossRef]
4. Song, B.; Liu, Z.; Gu, Y.; Zhou, X.; Zeng, S. Microstructures and fatigue fracture behavior of an Al–Cu–Mg–Ag alloy with a low
Cu/Mg ratio. Mater. Sci. Eng. A 2011, 530, 473–480.
5. Hu, Y.-J.; Sun, Y.-P.; Zhou, S.-P.; He, J.-M.; Yang, C.-Y. Effect of a cooling method on the structural and mechanical properties of
friction stir spot welding with a 2524 aluminum alloy. Mater. Res. Express 2021, 8, 026517. [CrossRef]
6. Srivatsan, T.S.; Kolar, D.; Magnusen, P. Influence of temperature on cyclic stress response, strain resistance, and fracture behavior
of aluminum alloy 2524. Mater. Sci. Eng. A 2001, 314, 118–130. [CrossRef]
7. Zheng, Z.Q.; Cai, B.; Zhai, T.; Li, S.C. The behavior of fatigue crack initiation and propagation in AA2524-T34 alloy. Mater. Sci.
Eng. A 2011, 528, 2017–2022. [CrossRef]
8. Srivatsan, T.S.; Kolar, D.; Magnusen, P. The cyclic fatigue and final fracture behavior of aluminum alloy 2524. Key Eng. Mater.
2008, 378–379, 207–230. [CrossRef]
9. Liu, C.; Ma, L.; Zhang, Z.; Fu, Z.; Liu, L. Research on the Corrosion Fatigue Property of 2524-T3 Aluminum Alloy. Metals 2021, 11,
1754. [CrossRef]
10. Pitcher, P.D.; Styles, C.M. Creep Age Forming of 2024A, 8090 and 7449 Alloys. Mater. Sci. Forum 2000, 331–337, 455–460. [CrossRef]
11. Siddiqui, R.A.; Abdullah, H.A.; Al-Belushi, K.R. Influence of aging parameters on the mechanical properties of 6063 aluminium
alloy. J. Mater. Processing Technol. 2000, 102, 234–240. [CrossRef]
12. Bray, G.H.; Glazov, M.; Rioj, R.J.; Lib, D.; Gangloffb, R.P. Effect of artificial aging on the fatigue crack propagation resistance of
2000 series aluminum alloys. Int. J. Fatigue 2001, 23, 265–276. [CrossRef]
13. Sarioğlu, F.; Orhaner, F.Ö. Effect of prolonged heating at 130◦ C on fatigue crack propagation of 2024 Al alloy in three orientations.
Mater. Sci. Eng. A 1998, 248, 115–119. [CrossRef]
14. Burba, M.E.; Caton, M.J.; Jha, S.K.; Szczepanski, C.J. Effect of Aging Treatment on Fatigue Behavior of an Al-Cu-Mg-Ag Alloy.
Metall. Mater. Trans. A 2013, 44, 4954–4967. [CrossRef]
15. Songbai, L.I.; Liyong, M.A.; Chi, L.; Jiuhuo, Y.I. Effects of Aging Temperature on Microstructure and High Cycle Fatigue
Performance of 7075 Aluminum Alloy. J. Wuhan Univ. Technol. (Mater. Sci.) 2017, 32, 677–684.
16. Xu, Y.; Zhan, L.; Li, W. Effect of pre-strain on creep aging behavior of 2524 aluminum alloy. J. Alloys Compd. 2017, 691,
564–571. [CrossRef]
17. Xu, Y.; Zhan, L.; Xu, L.; Huang, M. Experimental research on creep aging behavior of Al-Cu-Mg alloy with tensile and compressive
stresses. Mater. Sci. Eng. A 2017, 682, 54–62. [CrossRef]
18. Liu, Y.L.; Wang, Q.; Liu, C.; Song-Bai, L.I.; Jun, H.; Zhao, X.Q. Effect of creep and artificial aging on fatigue crack growth
performance of 2524 aluminum alloy. J. Jilin Univ. (Eng. Ed.) 2019, 49, 1636–1643.
19. Liu, C.; Liu, Y.; Li, S.; Ma, L.; Zhao, X.; Wang, Q. Effect of creep aging forming on the fatigue crack growth of an AA2524 alloy.
Mater. Sci. Eng. A 2018, 725, 375–381. [CrossRef]
20. ASTM E647-08. Standard Test Method for Measurement of Fatigue Crack Growth Rates. ASTM International: West Conshohocken,
PA, USA, 2008.
21. Khan, I.N.; Starink, M.J.; Yan, J.L. A model for precipitation kinetics and strengthening in Al–Cu–Mg alloys. Mater. Sci. Eng. A
2008, 472, 66–74. [CrossRef]
22. Quan, L.W.; Zhao, G.; Tian, N.; Huang, M.L. Effect of stress on microstructures of creep-aged 2524 alloy. Chin. J. Nonferrous Met.
2013, 23, 2209–2214. [CrossRef]
23. Zhang, H.; Qiu, X.; Xu, D.; Liu, Y.; Zhao, X. Effect of precipitated phase on dislocation activity under high-frequency impacting
and rolling. Micro Nano Lett. 2018, 13, 1542–1544. [CrossRef]
24. Benachour, M.; Hadjoui, A.; Benguediab, M.; Benachour, N. Stress Ratio Effect on Fatigue Behavior of Aircraft Aluminum Alloy
2024 T351. MRS Proc. 2011, 1276. [CrossRef]
25. Masoudi Nejad, R.; Berto, F.; Tohidi, M.; Jalayerian Darbandi, A.; Sina, N. An investigation on fatigue behavior of AA2024
aluminum alloy sheets in fuselage lap joints. Eng. Fail. Anal. 2021, 126, 105457. [CrossRef]
26. Xu, Y.; Yang, L.; Zhan, L.; Yu, H.; Huang, M.J.M. Creep Mechanisms of an Al–Cu–Mg Alloy at the Macro- and Micro-Scale: Effect
of the S /S Precipitate. Materials 2019, 12, 2907. [CrossRef]

162
Materials 2022, 15, 3244

27. Xu, Y.; Zhan, L.; Ma, Z.; Huang, M.; Wang, K.; Sun, Z. Effect of heating rate on creep aging behavior of Al-Cu-Mg alloy. Mater. Sci.
Eng. A 2017, 688, 488–497. [CrossRef]
28. Zhan, L.H.; Tan, S.G.; Yang, Y.L.; Huang, M.H.; Shen, W.Q.; Xing, Z. A Research on the Creep Age Forming of 2524 Aluminum
Alloy: Springback, Mechanical Properties, and Microstructures. Adv. Mech. Eng. 2014, 6, 707628. [CrossRef]
29. Lin, Y.C.; Jiang, Y.Q.; Zhang, X.C.; Deng, J.; Chen, X.M. Effect of creep-aging processing on corrosion resistance of an Al–Zn–Mg–
Cu alloy. Mater. Des. 2014, 61, 228–238. [CrossRef]
30. Gouma, P.I.; Lloyd, D.J.; Mills, M.J. Precipitation processes in Al–Mg–Cu alloys. Mater. Sci. Eng. A 2001, 319–321,
439–442. [CrossRef]
31. Kurguzov, V.D.; Kornev, V.M.; Moskvichev, V.V.; Kozlov, A.A. Influence of periodic change in the yield strength in a plate on the
development of plastic zones near a crack tip. J. Appl. Mech. Tech. Phys. 2014, 55, 1037–1044. [CrossRef]
32. Chen, Y.Q.; Pan, S.P.; Zhou, M.Z.; Yi, D.Q.; Xu, D.Z.; Xu, Y.F. Effects of inclusions, grain boundaries and grain orientations on the
fatigue crack initiation and propagation behavior of 2524-T3 Al alloy. Mater. Sci. Eng. A 2013, 580, 150–158. [CrossRef]
33. Liu, C.; Liu, Y.; Ma, L.; Li, S.; Zhao, X.; Wang, Q. Precipitate Evolution and Fatigue Crack Growth in Creep and Artificially Aged
Aluminum Alloy. Metals 2018, 8, 1039. [CrossRef]
34. Lin, Y.C.; Xia, Y.C.; Jiang, Y.Q.; Zhou, H.M.; Li, L.T. Precipitation hardening of 2024-T3 aluminum alloy during creep aging. Mater.
Sci. Eng. A 2013, 565, 420–429. [CrossRef]
35. Ungár, T.; Victoria, M.; Marmy, P.; Hanák, P.; Szenes, G. A new procedure of X-ray line profile analysis applied to study the
dislocation structure and subgrain size-distributions in fatigued MANET steel. J. Nucl. Mater. 2000, 276, 278–282. [CrossRef]

163
materials
Article
An Investigation of the Contact Fatigue Characteristics of an RV
Reducer Crankshaft, Considering the Hardness Gradients and
Initial Residual Stress
Xin Li, Wen Shao *, Jinyuan Tang, Han Ding and Weihua Zhou

State Key Laboratory of High Performance Complex Manufacturing, College of Mechanical and Electrical
Engineering, Central South University, Changsha 410083, China
* Correspondence: [email protected] or [email protected]

Abstract: The crankshaft is one of the core components of a Rotate Vector (RV) reducer. The fatigue
life of the RV reducer is severely hindered by fatigue failure on the eccentric cylindrical surface of the
crankshaft. The hardness gradients and residual stress in the crankshaft, associated with machining
operations, exert an enormous impact on the rolling contact fatigue (RCF). In this work, a finite ele-
ment method (FEM)-based three-dimensional elasto-plastic contact model is established to calculate
the stress–strain field by taking hardness gradients and initial residual stress into account. The RCF
characteristics of an RV reducer crankshaft is investigated by applying modified Fatemi–Socie (FS)
multiaxial fatigue criterion. The results indicate that initial residual stress plays an influential role in
the fatigue damage by altering the distribution of the maximum normal stress near the contact surface.
The modified FS fatigue criterion could better consider the effect of initial residual stress and the shear
stress, which significantly improves the prediction accuracy of the contact fatigue life model. The
contact fatigue performance could be considerably improved by designing appropriate shot peening
Citation: Li, X.; Shao, W.; Tang, J.;
Ding, H.; Zhou, W. An Investigation
parameters to obtain optimized residual stress distribution. Therefore, the technique presented may
of the Contact Fatigue Characteristics serve as an important guideline for the anti-fatigue design of an RV reducer crankshaft.
of an RV Reducer Crankshaft,
Considering the Hardness Gradients Keywords: RV reducer crankshaft; initial residual stress; hardness gradients; rolling contact fatigue;
and Initial Residual Stress. Materials modified Fatemi–Socie criterion
2022, 15, 7850. https://doi.org/
10.3390/ma15217850

Academic Editors: Lucjan Śnieżek,


1. Introduction
Jaroslaw Galkiewicz and Sebastian
Lipiec The RV (Rotate Vector) reducer is extensively used in industry robots owing to its
peculiar and fascinating properties, such as compact structure, small size, light weight,
Received: 31 August 2022 high reduction ratio, high transmission accuracy and efficiency, high torsional rigidity,
Accepted: 3 November 2022 etc. [1]. The crankshaft and cylindrical roller bearing are the core components of the RV
Published: 7 November 2022
reducer. Due to the limited available space of the RV reducer, the cylindrical roller bearing
Publisher’s Note: MDPI stays neutral is usually without the inner ring and outer ring. That is, the inner and outer rings of the
with regard to jurisdictional claims in bearing are directly composed of the eccentric cylindrical surface of the crankshaft and
published maps and institutional affil- the bearing inner hole on the cycloid wheel. The transmission errors and fatigue life of
iations. the RV reducer is severely hindered by failure modes, such as pitting and spalling on the
eccentric cylindrical surface caused by the long-term high cyclic contact stress between the
eccentric cylindrical surface of the crankshaft and the roller bearing [2]. With increasingly
higher performance requirements, such as high transmission accuracy, high load-carrying
Copyright: © 2022 by the authors.
capacity, and long service life of the RV reducer, the RCF of the crankshaft has also become
Licensee MDPI, Basel, Switzerland.
the limiting factor affecting the reliability of the RV reducer.
This article is an open access article
Several attempts have been made towards improving the transmission performance
distributed under the terms and
of RV reducers. Zhang et al. [3] established a mixed lubrication analysis model for RV
conditions of the Creative Commons
reducers. The contact load, surface roughness, and geometry of the cylindrical roller
Attribution (CC BY) license (https://
creativecommons.org/licenses/by/
bearings were innovatively included in their model. Xu et al. [4,5] developed a dynamic
4.0/).
model for the transmission systems of an RV reducer that took into account the cylindrical

Materials 2022, 15, 7850. https://doi.org/10.3390/ma15217850 165 https://www.mdpi.com/journal/materials


Materials 2022, 15, 7850

roller bearing’s radial clearance. Wang et al. [6] presented a contact force and transmission
error analysis of an RV Reducer. Their results revealed that the modified model, based
on contact force curves, could improve the transmission performance of the RV reducer
as the load increased. Deng et al. [7] calculated the rated life of the RV reducer’s angular
contact ball bearing. They also discovered that bearing life had a significant impact on
the total life of the RV reducer under heavy load conditions. The present investigations
on RV reducers have been mostly focused on the meshing characteristics of the cycloid
pinwheel, the dynamic characteristics of the RV reducer, and the life estimation of the
bearing. Few studies have been reported on the RCF performances of the crankshaft of the
RV reducer due to the complexity of the contact load between the RV reducer crankshaft
and the cylindrical roller bearing, although it exerts an enormous impact on the reliability
and fatigue life of the RV reducer.
Concerted efforts were directed towards the fatigue life and damage prediction meth-
ods for gears and rolling bearings. Li et al. [8] presented a multiaxial fatigue model
considering mixed lubrication for crack initiation life prediction of spur gears. Liu et al. [9]
proposed an improved multiaxial fatigue life model with higher life prediction accuracy
compared with classical models. Vijay et al. [10] presented a novel model to simulate
the crack initiation and propagation in bearing steels, considering the anisotropy of crys-
tal. Their results indicated that the Fatemi–Socie (FS) criterion could be used to estimate
the RCF life. Continuum damage mechanics (CDM) and the elasto-plastic model with
damage-coupling were widely used to investigate the spalling initiation and propagation
behaviors of cylindrical rolling bearings [11,12]. Despite the efforts of these earlier studies,
prediction of the fatigue life and damage from these investigations may inevitably suffer
from the disadvantage of neglecting the effect of mechanical property that is introduced in
the manufacturing processes, such as heat treatment, grinding, shot peening, etc.
In recent years, considerable progress has been achieved for the evaluation of the
contact fatigue performance by taking surface integrity such as surface topography, residual
stress, hardness gradient, and microstructure into account. Numerous researchers have
shown that surface integrity significantly affects the fatigue life in rolling contact [13].
Choi et al. [14] pointed out that the prediction accuracy of fatigue life increased more
than 40% when the residual stress was taken into account. The results also showed that
increments of more than 12 and 8 times could be reached for the crack initiation propaga-
tion lives if residual stress was considered. The critical plane approach was recognized
as an effective method to solve multiaxial fatigue problems [9]. The Dang–Van fatigue
criterion, the FS criterion, and Zaretsky fatigue life model were applied to study the contact
fatigue performance of a carburized gear under heavy loading by incorporating residual
stress [15,16]. Mahdavi et al. [17] investigated the effect of superposed residual stresses on
micro plasticity around inclusions in bearing steel. Ooi et al. [18] experimentally studied
the impact of restrained austenite and residual stress on the fatigue life of carburized AISI
8620 steel. Although residual stress was considered in [17,18], the distribution of residual
stress was not well represented in these studies for the estimation of RCF. Guan et al. [19]
investigated the influence of compressive residual stress (CRS) induced by shot peening on
fatigue risks and found that appropriate CRS distribution could decrease the rate of damage
accumulation in bearing steel containing carbide. Walvekar et al. [20] studied the combined
impact of hardness gradient and residual stress curves on the RCF lives of bearing steel
materials. It was found that optimized carburizing depth could prolong the fatigue life of
bearing steel materials to a large extent. Furthermore, the relationship between associated
parameters such as effective case hardening depth (CHD), surface hardness, and hardness
curve shape and fatigue performance was also demonstrated [21,22].
The above studies confirm that it is of theoretical value and engineering significance to
consider the mechanical property gradient and initial residual stress when evaluating the
contact fatigue performance of carburized steel materials or components. The researchers
above also provided a theoretical basis for the contact fatigue characteristic analysis of
the crankshaft. Moreover, the successful prediction of the fatigue life of the crankshaft

166
Materials 2022, 15, 7850

incorporating residual stress and hardness gradient through numerical modelling provides
the superiorities of yielding numerous results in a short period and sometimes offers
beneficial insight into different states of RCF processes, which can be also used to optimize
the machining parameters. Therefore, it is of great importance and necessity to consider
the initial state of the crankshaft in the contact fatigue life prediction model.
In this work, focusing on fatigue life assessment of an RV reducer crankshaft, an FEM-
based three-dimensional elasto-plastic contact model is established. The hardness gradients
and initial residual stress are obtained by Vickers hardness tests and X-ray diffraction
method. The collected hardness and residual stress data are incorporated into the elasto-
plastic contact model and then the stress–strain histories are obtained. Contact fatigue life
assessment is performed by the critical plane method and multiaxial fatigue criterion. The
influence of friction factor and initial residual stress on fatigue damage and fatigue life is
investigated. Moreover, the effect of residual stress induced by shot peening on improving
fatigue life is also analyzed quantitatively.

2. Failure Analysis of Crankshaft


The reliability and fatigue life are important performance indicators of the RV reducer,
so it is necessary to carry out fatigue life tests of the RV reducer. In addition, the fatigue
life test is not only conducive to analyzing the performance degradation law and failure
mechanism of the reducer, but also provides guidance for the design and manufacture of
key parts of the RV reducer.
Figure 1 shows fatigue life test system of an RV reducer. Because the RV reducer is a
piece of high-precision transmission equipment, transmission efficiency can be selected as
the failure judgment criterion, and comprehensive judgment can be made in combination
with vibration, temperature rise, and noise. One RV reducer testing machine is selected for
fatigue test. The test is stopped when the transmission efficiency is detected to be lower
than 85% of the threshold, and a service life of approximately 2150 h was obtained.

ȱ
Figure 1. Fatigue life test system of RV Reducer.

The reducer was disassembled after the fatigue test. Comparing the new crankshaft
parts, it was found that the cylindrical surface of the needle roller bearing in contact with
the crankshaft was seriously worn, as shown in Figure 2.

Figure 2. Macroscopic morphology of crankshaft: (a) the new part; (b) the failed part.

167
Materials 2022, 15, 7850

The SEM micrographs of the eccentric cylindrical surface of the crankshaft before and
after failure are shown in Figure 3. Regular grinding grooves are left on the new crankshaft
surface (as shown in Figure 3a). There are thin strip-like scratches along the rolling direction,
and numerous pit-like cracks appeared on the surface of the failed crankshaft (Figure 3b,c).

ȱ
Figure 3. The SEM micrographs of the eccentric cylindrical surface of the crankshaft: (a) surface of
new parts; (b,c) surface after failure.

Because the crank shaft is the core part of the power input and output of the reducer,
it bears periodic radial load and dynamic load. Moreover, the reducer often experiences
impact load, continuous start and stop, and other working conditions, which leads to
complex load changes on the crankshaft. Therefore, the eccentric cylindrical surface of
the crankshaft is also the most prone to failure in engineering practice. Fatigue pitting
occurs on the eccentric cylindrical surface of the crankshaft under the long-term high cyclic
contact stress, and a series of pits are formed. The crankshaft is one of the weakest links of
the RV reducer, which severely restricts the fatigue life of the RV reducer.

3. Methodology
3.1. Stress Analysis of RV Reducer
The crankshaft used in this study is taken from the RV reducer of an industrial robot.
The structure of the RV reducer is shown in Figure 4. The critical parameters of the rolling
contact pair between the crankshaft and roller are given in Table 1. The crankshaft material
is 20CrNi2MoA. The composition of it is listed in Table 2. The crankshaft has undergone
several manufacturing processes, such as carburizing, quenching, tempering, and finally
precision grinding. The detailed thermal treatment process is shown in Figure 5. The
crankshaft sample is carburized and diffused for 6 h after the temperature soared to 930 ◦ C,
then the sample is quickly quenched in oil, followed by low temperature tempering at
230 ◦ C for 2 h. After heat treatment, the lath martensite structure is finally obtained. Lath
martensite can better resist impact and crack propagation, so that the material has higher
hardness, good wear resistance, and higher contact fatigue properties.
Through the motion and force analysis of the cycloidal gear and cylindrical roller
bearing (Figure 6), the resultant force acting on the cylindrical roller bearing is obtained.
The cylindrical roller bearing bears the force from the cycloidal gear and the crankshaft.
Fj1 , Fj2 , and Fj3 are the components of force, F, which acts on the cycloidal gear via the
cylindrical roller bearing (as shown in Figure 6a). F can also be decomposed into the normal
force, Fr , and tangential force, Ft (as shown in Figure 6b).

168
Materials 2022, 15, 7850

Figure 4. Transmission schematic diagram of RV Reducer.

Table 1. Crankshaft-bearing rolling pair parameters.

Parameters Value Parameters Value


Crankshaft rotation speed (r/min) ns = 585 Rated output torque (N·m) Tout = 800
Radius of the rolling element (mm) R1 = 4.0 Rolling element material GCr15
Radius of the eccentric cylindrical surface (mm) R2 = 16.6 Crankshaft material 20CrNi2MoA
Length of the rolling element (mm) l1 = 12.0 Young’s modulus (GPa) E1 = 219, E2 = 210
Radius of needle tooth distribution circle (mm) Rz = 82 Poisson ratio v1 = 0.3, v2 = 0.275
Radius of crankshaft distribution circle (mm) ro = 46.77 Eccentricity (mm) e = 1.5
Number of teeth of needle wheel z4 = 39 Number of crankshafts n=3
Number of teeth of cycloid gear z5 = 40 Short amplitude coefficient k = 0.7317

Table 2. The composition of 20CrNi2MoA.

Element C Mn Cr Ni Si Mo Cu S P Fe
Wt.% 0.21 0.63 0.57 1.8 0.33 0.25 0.30 0.015 0.02 Bal.

T
+HDWLQJDQG
&DUEXUL]LQJDW 'LIIXVLQJDW
LQVXODWLRQDW
ć ć 3UHFRROLQJ
ć
WRć
2LO
FRROLQJ

7HPSHULQJ
C  C  C  DWć
KaK KaK K $LU
K FRROLQJ
t
Figure 5. The curve of the thermal treatment process.

Based on the structural parameters of the RV reducer and the force equilibrium, F can
be expressed as follows, and the detailed calculation process can be seen in [23].

Tout
F= Ft2 + Fr2 = (ez5 )2 + ro2 + k2y ro2 + 2ez5 ro cos ϕ − 2k y ez5 ro sin ϕ (1)
2nez5 ro

Combining Table 1 and Equation (1), the resultant force distribution of the cycloid
gear acting on the cylindrical roller is obtained, as depicted in Figure 7. Periodic sinusoids
are found for the forces acting on the cylindrical rollers. The load fluctuation of three

169
Materials 2022, 15, 7850

cylindrical roller bearings in the same cycloid gear are found to have a phase difference of
120◦ . According to Harris et al. [24], the equivalent load of cylindrical roller bearings can
be calculated as [23]:
⎛    2 ⎞1/4
4 2 2
Tout ⎜ (ez5 ) + 4 1 + k y (ez5 ) ro + 1 + k y ro ⎟
2 2 4
Fm = ⎝ ⎠ (2)
2 n4 (ez5 )4 ro4

D y E y
&\FORLGDOJHDU

¦Py 5ROOHUEHDULQJ
P ¦Px &UDQNVKDIW y
Fj Fj x
Fj2 Fj2 or Ft
F
M M Fr
o Fj3 x Fj3 x
o
o o

5ROOHUEHDULQJ
5ROOHUEHDULQJ
ȱ
Figure 6. Schematic diagram: (a) force on cycloid gear; (b) force on cylindrical roller bearing.






7KHF\OLQGULFDOUROOHUEHDULQJ
7KHF\OLQGULFDOUROOHUEHDULQJ
)RUFH 1

 7KHF\OLQGULFDOUROOHUEHDULQJ








         
7KHURWDWLRQDQJOHRIWKHFUDQNVKDIW ƒ ȱ
Figure 7. The variation of forces acting on the cylindrical roller bearings.

Power is transmitted between the crankshaft and the cylindrical rollers. Dynamic
loads and relatively large radial forces act on the cylindrical rollers. The force analysis of
each roller on the cylindrical roller bearing is carried out after obtaining the resultant force
on the cylindrical roller bearing. The Newton–Raphson algorithm is employed to calculate
the load distribution of each rolling element acting on the crankshaft. The crankshaft is
subjected to periodic load, hence the equivalent contact load acting on the crankshaft by
the cylindrical roller bearing will be used as input in the following section.

3.2. Modeling of Residual Stress and Hardness Gradient


The x, y, and z directions in the crankshaft coordinates act along the rolling, depth,
and axial directions, respectively. Several empirical methods for fitting the hardness
distribution profile introduced by carburizing have been proposed by Lang and Kernen [25]

170
Materials 2022, 15, 7850

and Thomas [26]. Among them, the Thomas method is the closest to the measured results
and has been widely used in ISO standards. The empirical formula can be expressed as [26]:


⎪ a ·y2 + ba ·y + c a (0 ≤ y < CHD )
⎨ a
HV (y) = ab ·y2 + bb ·y + cb (CHD ≤ y < ycor ) (3)


⎩ HV (y ≤ y)
cor cor
⎧ 550− HVsur

⎪ a a = CHD2 − ; b = −2· a a ·y HV,max ; c a = HVsur ;
2·y HV,max ·CHD a



H (CHD )
⎪ ab = 2·(CHD−y ) ; bb = −2· ab ·ycor ; (4)


cor

cb = 550 − ab ·CHD2 − bb ·CHD; H  (CHD ) = 2· a a ·CHD + ba ;
where HVsur is the hardness on the surface, HVcor is the hardness in the core, and ycor and
yHV,max denote the depths with hardness equal to HVcor and the maximum hardness, respec-
tively. In this study, yHV,max is equal to zero. The depth with a hardness of 550 HV is defined
as CHD (the case hardening depth). The CHD, HVsur , and HVcor are designed as 1.0 mm,
670 HV, and 450 HV, respectively, based on the engineering practice of the crankshaft.
The crankshaft sample is cut along the cross-section by a wire cutter, inlaid and
polished, and then the cross-sectional microhardness is measured by a Vickers hardness
tester. A pyramid diamond indenter is selected for the hardness test, with a load of 0.5 kgf
and duration of 15 s. Different positions at the same depth were measured at least two times,
and the average hardness was obtained. Figure 8 shows the hardness data (the green circle)
measured by the hardness tester machine. The red and black dashed lines are used for
determine the value of CHD. The empirical hardness gradient curve (the blue solid line)
based on the Thomas method is also shown in Figure 8. It can be seen that the empirical
curve correlates well with the experimental data.

 0HDVXUHGGDWD
+DUGQHVVEDVHGRQ
 7KRPDVPHWKRG




+9





 &+' PP



      


'HSWKIURPVXUIDFH PP
Figure 8. The measured hardness data and the empirical hardness curve.

The tensile strength and yield strength and residual stress distribution will be altered
by the variation of martensite, retained austenite, and other structural components during
the carburizing-quenching process. A linear relationship existed between Vickers hardness
and tensile strength and yield strength, which was given by [27,28].

HV (y)
σys (y) = ·(0.1)m−2 (5)
3

HV (y) 12.5(m − 2) (m−2)
σb (y) = [1 − (m − 2)] (6)
3 1 − ( m − 2)

171
Materials 2022, 15, 7850

where σys (y), HV(y), and σb (y) are the yield strength, Vickers hardness, and the tensile
strength, respectively, and m is the Meyer hardness coefficient, which can be taken as
2.19 for high-strength steel materials.
The generally recognized X-ray diffraction test with electrolytic polisher and empirical
formula can be used to determine the residual stress at the contact surface or near-surface
areas of the crankshaft. In the present work, the electropolishing method (as shown in
Figure 9b) is employed to remove the material layer by layer, and the corresponding
residual stress distribution is examined through a Proto iXRD system with Cr-Kα radiation
(as shown in Figure 9a). The magnitudes of the residual stress along the x and z axis (σr,x
and σr,z ) are found to be equivalent at the same depth, thus the residual stress is expressed
as symbol σr in the subsequent section. The residual stress along the depth direction, σr,y ,
is always negligible compared with the other two components, and is thus ignored in
this study.

Figure 9. Residual stress measurement: (a) X-ray diffractometer; (b) Electrolytic polisher.

Empirical methods have also been provided to characterize the linear relationship
between the hardness curve and the residual stress distribution [29,30]. Hertter’s empirical
formula employed in this study is given by [30]:

−1.25·( HV (y) − HVcore ) ( HV (y) − HVcore ≤ 300)
σr (y) = (7)
0.2857·( HV (y) − HVcore ) − 460 ( HV (y) − HVcore > 300)

Figure 10 shows the residual stress data measured by X-ray diffraction (the green
circle), and the empirical curve (the yellow solid line) is obtained by the Hertter’s method.
It can be observed that the measured residual stresses are, in general, consistent with those
fitted by the empirical formula. Therefore, the empirical residual stress curve is employed
in the numerical model.
Moreover, the influence of tensile residual stress caused by improper carburizing or
grinding burns during the machining process on fatigue damage will also be analyzed
in this study [31]. Ultrasonic vibration assisted grinding (UAG) has many advantages,
such as reducing grinding force and improving surface quality [32]. A hypothetic tensile
residual stress curve (the blue dashed line) with the same amplitude as compressive residual
compressive stress is depicted in Figure 10.
The residual stress generated by the plastic deformation near the surface is also
caused by surface strengthening processes such as shot peening and ultrasonic rolling.
Shot peening produces spoon-shaped residual compressive stress distribution, which is
conducive to improving the fatigue performance [33]. Zhao at al. [34] proposed a model for
calculating the residual stress distribution after shot peening. The maximum error between
the measured and simulated residual stress was 15.8%, which verified the accuracy of
the method. A residual stress curve distributed along the depth is designed based on the
residual stress data of shot peening in [34], as shown in Figure 11. The maximum plastic
deformation near the surface layer leads to the maximum residual compressive stress. The

172
Materials 2022, 15, 7850

maximum residual stress depth is 50 μm. The compressive stress decreases gradually with
the increase of depth and then tends to be stable.

,756VUVXUIDFH 03D


 56EDVHGRQ+HUWWHUPHWKRG
5HVLGXDOVWUHVVıU 03D

+\SRWKHWLFDOWHQVLOHVWUHVV
 0HDVXUHGGDWDLQUROOLQJGLUHFWLRQ



,&56VUVXUIDFH 03D




      


'HSWKIURPVXUIDFH PP ȱ
Figure 10. The measured residual stress data and the empirical stress curve.



 6356VUVXUIDFH 03D


5HVLGXDOVWUHVVıU 03D









      


'HSWKIURPVXUIDFH PP ȱ
Figure 11. The residual stress distributions induced by shot peening.

3.3. FEM-Based Elasto-Plastic Contact Analysis


The rolling contact between the eccentric cylindrical surface of the crankshaft and the
cylindrical roller bearing can be simplified as rigid cylindrical surface and a semi-infinite
space according to the contact mechanics theory [35]. The diagram of the contact model of
the crankshaft and the cylindrical roller is depicted in Figure 12.
In order to better reflect the three-dimensional stress–strain field response, a three-
dimensional elasto-plastic finite element contact model is established in the commercial
finite element software ABAQUS (Figure 13). The equivalent curvature radius of the rigid
cylindrical surface is R = 3.22 mm. The normal load, F, applied on the rigid cylindrical
surface, is determined by the output torque of the RV reducer during numerical simulation.
The bottom nodes of the mesh model are fully fixed, and symmetric boundary conditions
are set around the mesh model. That is, the nodes on the left and right sides of the mesh
model are constrained in x and y directions. A static implicit solver is selected for simu-
lation calculation. The computational domain is determined as −500 μm ≤ x ≤ 500 μm,

173
Materials 2022, 15, 7850

0 μm ≤ y ≤ 600 μm, and 0 μm ≤ z ≤ 200 μm. The rolling direction, x, is long enough to
minimize the influence of the boundary on the stress calculation in this direction.

Figure 12. The contact model of the crankshaft and the cylindrical roller bearing.

ȱ
Figure 13. The numerical elasto-plastic contact model.

Considering the efficiency and accuracy of finite element calculation, the grid size is
gradually expanded from the contact surface to the bottom. The grid size in the x and z
directions is 5 μm. The grid size of the near surface layer in the y direction is 5 μm, and the
grid size is set to 10 μm as the depth increases. The C3D8R element is selected because it
can well withstand distortion, and its stress–strain calculation is also more accurate. The
tangential contact property is defined as frictional contact with a friction coefficient of
0.05. The friction coefficient of 0.05, which was also used in [36,37], representing excellent
lubrication in rolling contact, is employed in this study. The rigid cylindrical surface
rolls from x = −400 μm to x = 400 μm, and multiple analysis steps are set to realize the
reciprocating circular motion of the cylindrical surface. The material points lying on the
yellow dashed line (x = 0 and y ∈ [0, 600] μm) will be analyzed in the present study because
the material at the same depth experiences the same cyclical stress.
Ultra-high stress exists in the kinematic pair (crankshaft and the cylindrical roller bear-
ing) during extreme operation conditions, which inevitably produces plastic deformation.
To accurately characterize the elastoplastic response, the varying yield stress with depth
is incorporated into the FEM model, according to the linear positive correlation between
the hardness and yield strength (Equation (5)). The efficiency for assigning material prop-
erties can be improved through the secondary development by Python. Therefore, the
isotropic hardening model is applied according to the uniaxial tensile experiment, and the
true stress-plastic strain fitting curve is shown in Figure 14. The uniaxial tensile test of
material 20CrNi2Mo specimens was completed to investigate the effect of quenching tem-
perature on the properties of strength and toughness [38]. Moreover, in [39], quasi-static

174
Materials 2022, 15, 7850

tensile/compressive experiments were carried out to obtain the mechanical properties


of high strength steel. These experimental results play a great role in this study. Be-
cause the yield strength of the bearing rolling element (1617 MPa) is much higher than
that of the crankshaft (1413 MPa), contact fatigue failure is more likely to occur on the
crankshaft. Therefore, the following mainly analyzes the contact fatigue characteristics of
the crankshaft.





7UXHVWUHVV 03D









     


7XUHVWUDLQ
Figure 14. The true stress-plastic strain curve.

The initial residual stress field is defined before starting the response history. The
hexahedral element has six stress components (σ11 , σ22 , σ33 , σ12, σ13, σ23 ). The shear stress
and the stress in the depth direction are negligible according to the above residual stress
measurement. Therefore, normal stress components σ11 and σ33 , confirmed at residual
stress σr (Figures 10 and 11), are applied to the element integration points of the model
(Figure 13). The repetitive work of initial residual stress application can be reduced through
the secondary development of Python. After the initial stress field is applied in the finite
element model, the equivalent node load formed by the stress field need to be in equilibrium
with the specified boundary conditions. Therefore, a static analysis step without external
load is set to obtain the actual residual stress field. The actual residual stress obtained by
adding different types of residual stress curves (Figures 10 and 11) is shown in Figure 15.
The variation of initial residual stress after equilibrium is negligible, which verifies that the
method of applying the initial residual stress field to this model is reasonable.

ȱ
Figure 15. The actual residual stress after equilibrium (without external load).

175
Materials 2022, 15, 7850

3.4. Contact Fatigue Life Assessment Model


The stress in the contact region varies non-proportionally in the loading cycle due to
the multiaxial characteristics of the stress state. Thus, it is necessary to apply a preferable
multiaxial fatigue criterion to capture stress–strain response and then estimate the fatigue
life of the crankshaft in such a complicated time-varying stress state. The FS multiaxial
fatigue criterion dominated by shear-type fatigue failure is employed in this work. The
normal stress and shear strain with maximum amplitudes are used as fatigue damage
parameters. The fatigue life and fatigue damage (FD) in this criterion are respectively
expressed as [40,41]:

Δγmax σn,max τ f
(1 + k )= (2N f )b + γ f (2N f )c (8)
2 σys G

Δγmax σn,max
FDFS = [1 + k ] (9)
2 σys
where Δγmax /2 = γa denotes the maximum shear strain amplitude, σn,max denotes the
maximum normal stress perpendicular to the critical plane, which is assumed to be the
plane experiencing maximum shear strain amplitude; b and c denote the shear fatigue
strength and the shear fatigue ductility indexes, set as −0.087 and −0.58 [42]; G and Nf
are the shear elastic modulus and the crack initiation life, respectively; k is the material
constant (set as 1 in this study) [43]; and τ  f and γ f denote the shear fatigue strength and
shear fatigue ductility coefficients, which are defined as follows [44]:
√ √
τ  f = σ f / 3, γ f = 3ε f (10)

where σ f and ε f denote the axial fatigue strength and axial fatigue ductility coefficients,
which are calculated by Baumel and Seeger’s method [42]:

σ f = 1.5σb , ε f = 0.59ψ (11)



1.0 (σb /E< 0 .003)
ψ= (12)
1.375 − 125(σb /E) (σb /E> 0 .003)
where E is the Young’s modulus and σb is the tensile strength.
When analyzing the fatigue data, including combined axial-torsion load paths, it is
found that if the maximum normal stress is normalized by the shear stress range, the
prediction accuracy of fatigue life can be improved. That is, replacing yield stress, σys ,
with the shear stress range, GΔγ, the modified fatigue damage (FDmod ) is calculated as
follows [45]:

Δγmax σn,max τ f
FDmod = (1 + k )= (2N f )b + γ f (2N f )c (13)
2 GΔγ G

The multi-axial stress–strain histories are calculated by using the model outlined in
Section 3.3. The stress tensor time history and the strain tensor time history at each element
is expressed as σ(t) and ε(t), respectively.
⎡ ⎤
⎡ ⎤ ε xx (t) 2 γxy ( t ) 2 γxz ( t )
1 1
σxx (t) τxy (t) τxz (t) ⎢1 ⎥
σ(t) = ⎣ τxy (t) σyy (t) τyz (t) ⎦, ε(t) = ⎢
⎣ 2 γxy (t) ε yy (t)

2 γyz ( t ) ⎦
1 (14)
τxz (t) τyz (t) σzz (t)
2 γxz ( t ) 2 γyz ( t ) ε zz (t)
1 1

Once the multi-axial stress–strain histories are obtained, Euler angle-based axis trans-
formation is employed to search for the critical planes that cover all the possible directions
in the 3D space [46], as shown in Figure 16. The transformation of stress vector and strain

176
Materials 2022, 15, 7850

vector is similar, so the former will be explained in detail below. Considering each element
integration point, O is in the center of the absolute reference frame, {O; x, y, z}. The orienta-
tion of a material plane, Ω, with unit normal vector, n (nx , ny , nz ), can be located by using
spherical coordinate parameters αp and βp . Parameter α is the angle between the projection
of unit normal vector n on plane x–y and axis x, while β is the angle between unit vector n
and axis z [47].

Figure 16. Definition of the Euler transformations and critical plane search method.

The transformation coordinate system {O; a , b , n} is defined by searching across two


angular ranges (αp ∈ [0, 2π], βp ∈ [0, π]), as shown in Figure 16. Axis n is parallel to unit
vector n, whereas axes a and b are parallel to material plane Ω, and the corresponding
unit vectors are a and b, respectively.
⎡⎤ ⎡ ⎤ ⎡ ⎤ ⎡ ⎤ ⎡ ⎤ ⎡ ⎤
nx sin β p · cos α p ax − sin α p bx − cos β p · cos α p
n = ny = sin β p · sin α p ⎦, a =
⎣ ⎦ ⎣ ⎣ay ⎦ = ⎣ cos α p ⎦, b = ⎣by ⎦ = ⎣ − cos β p · sin α p ⎦ (15)
nz cos β p az 0 bz sin β p

In order to easily calculate the normal stress, σn (t), and shear stress, τ n (t), associated
with the material plane, Ω, the total stress vector, t(t), depending on the stress tensor, σ(t),
needs to be calculated as:
⎡ ⎤ ⎡ ⎤⎡ ⎤
tx (t) σxx (t) τxy (t) τxz (t) nx
t(t) = ty (t) = σ(t)·n = τxy (t) σyy (t) τyz (t) ⎦⎣ny ⎦
⎣ ⎦ ⎣ (16)
tz ( t ) τxz (t) τyz (t) σzz (t) nz

The normal stress and component of shear stress on two axes can be obtained from
the above formula, as follows:

⎨ σn (t) = tx (t)·nx + ty (t)·ny + tz (t)·nz
{O; a , b , n} ⇒ τ (t) = tx (t)·ax + ty (t)·ay + tz (t)·az (17)
⎩ na
τnb (t) = tx (t)·bx + ty (t)·by + tz (t)·bz

The shear stress, τ n (t), is obtained as follows:



τn (t) = τna2 (t) + τ 2 (t)
nb (18)

During the load cycle, normal stress, σn (t), varies its magnitude, but the direction
remains parallel to vector n. Therefore, its maximum value can simply be calculated
as follows:
σn,max = max[σn (t)] (19)
Meanwhile, shear stress, τ n (t), varies its magnitude and direction with time, which
makes the solution of the shear stress amplitude more complicated. This coordinate system,
{O; a , b , n}, is then rotated about axis n until the amplitude of the shear stress component

177
Materials 2022, 15, 7850

along axis ap is maximized. The new coordinate system is recorded as {O; ap , bp , n}, and
Ωp is used to represent the critical plane, as shown in Figure 16, assuming that the rotation
angle is θ p (θ p ∈ [0, π]) and the transformation matrix is T(θ p ), defined by:
⎡ ⎤
1 0 0
T ( θ p ) = ⎣0 cos θ p sin θ p ⎦ (20)
0 − sin θ p cos θ p

The stress tensors around this coordinate system are given as:
⎧ P
⎨ σn (t) = T (θ p )·σn (t)
{O; a , b , n} ⇒
p p
τ P (t) = T (θ p )·τna (t) (21)
⎩ naP ( t ) = T ( θ )· τ ( t )
τnb p nb

That is, the θ p that maximizes the in-plane shear stress amplitude along axes ap is
p
searched for, and at this moment the amplitude of τnb (t) is the smallest [48]. Thus, the
stress vector satisfying the following relationship can be obtained:

1
τn,a = [maxτn (t) − minτn (t)] (22)
2
The overall computational methodology of estimating the fatigue life demonstrated
above is summarized in the flow chart of Figure 17. Firstly, the stable stress–strain field
is calculated by the elasto-plastic contact model, and the stress–strain components of the
target material pointing along the depth is extracted through the secondary development
of Python. Then, the stress, strain (σn,max and Δγmax ), and FDmod are calculated. Finally,
the fatigue initiation life is evaluated via the Newton–Raphson method.

Figure 17. Computational methodology of predicting fatigue life.

4. Results and Discussion


4.1. Effect of Friction Coefficient and Normal Stress on Fatigue Damage
The friction coefficients of 0, 0.05, and 0.2 correspond to ideal smooth contact without
friction, the reasonable friction coefficient for electrohydrodynamic lubricated bearing con-
tact, and the state where the excellent lubrication environment of the reducer is destroyed.
These three friction coefficients are selected firstly to investigate the effect of lubrication
on the shear strain, γa , and the modified fatigue damage, FDmod , on the material plane
(θ p ∈ [0, π]) with the initial compressive residual stress (ICRS) case under the framework of
modified FS criterion. The output torque is set as the rated output torque of 800 N·m and
the Maximum contact stress Ph = 1.56 GPa. As can be seen from Figure 18, the contour of
shear strain amplitude, γa , is symmetrical, and γa,max lies at the plane orientation angles
of 90◦ and 0◦ (180◦ ) when the friction coefficient, μ, is equal to 0. The distribution of γa
and FDmod over θ p is asymmetric when μ
= 0. Moreover, the worse lubrication, with a
friction coefficient of 0.2, leads to a much more obvious asymmetry of the FDmod contour.

178
Materials 2022, 15, 7850

According to the definition of the critical plane, the plane orientations are along θ p = 97◦
and θ p = 101◦ when the friction coefficients are 0.05 and 0.2. The location of the maximum
damage gradually moves to the contact surface, and the fatigue damage increases as the
friction coefficient increases; similar results are also found in Ref [49].

P  î
 P  î
 P  î

Ȗa    
i i i
 q q
 q  
'HSWK ȝP


  

  

  


   



 
 î î î
  
FDPRG i i i i

  

'HSWK ȝP

  



  

  


   


                    
3ODQH2ULHQWDWLRQ$QJOH ƒ 3ODQH2ULHQWDWLRQ$QJOH ƒ 3ODQH2ULHQWDWLRQ$QJOH ƒ ȱ
Figure 18. Variation of γa and FDmod under different lubrication conditions (Tout = 800 N·m,
Ph = 1.56 GPa).

Three residual stress conditions corresponding to the residual stress curves shown
in Figure 10, such as initial compressive residual stress (ICRS), without initial residual
stress (IRS), and initial tensile residual stress (ITRS), are chosen to explore the influence
of initial residual stress on the fatigue damage (Figure 19). The output torque and friction
coefficient are set as 800 N·m and 0.05, respectively. It can be observed from Figure 19 that
the distributions of the maximum normal stress under three initial residual stress states
are completely different at the subsurface, and the initial residual stress states appreciably
influences the maximum normal stress, σmax , near the θ p = 0◦ (180◦ ) plane. Compared with
the without the IRS case, ICRS reduces σmax by 48.4% while ITRS increases the value of
σmax by 99.8%.
The variations of FS damage (FDFS ) and modified FS damage (FDmod ) are also plotted
versus plane orientation in Figure 19, which can provide additional insight into the damage
mechanisms. Figure 19 shows that the variations of damage value against different plane
angles under the modified FS criterion is more obvious than that under the FS criterion.
The maximum fatigue damage, FDmod,max , occurs at the θ p = 97◦ plane with the ICRS
case, while the maximum damage, FDmod,max , shifts from θ p = 97◦ to θ p = 5◦ as the initial
residual stress changes from compressive stress to tensile stress. This might be attributed
to the fact that the increase of maximum normal stress leads to the increase of the damage
near the θ p = 0◦ (180◦ ) plane. Therefore, the initial residual stress affects the position of
FDmod,max by influencing the distribution of the maximum normal stress under the FS
criterion; similar results are also found in [50]. Moreover, the tensile normal stress with the
ITRS case results in higher damage predictions near the 0◦ (180◦ ) plane for modified FS
damage. Tensile stress help to accelerate shear crack growth, while compressive normal
stress with the ICRS case may serve to increase friction between the cracks and reduce
the crack driving force. The above also demonstrates that the ratio of maximum normal
stress to shear stress in modified FS damage (FDmod ) can better consider the interaction
effect between the two stress [45]. Therefore, the ICRS would lead to a decrease in fatigue
damage and increase the crack initiation life, Nf .

179
Materials 2022, 15, 7850

:LWK,&56VUVXUIDFH 03D 03D :LWKRXW,56 03D


:LWK,756VUVXUIDFH 03D
03D
ıQPD[    
   
'HSWK ȝP


  

  

   
   
  
FDFS  î î î
  

  
'HSWK ȝP


  

  

   
   
  
î î î
FDPRG    

  
'HSWK ȝP


  

  

   
   
                    
3ODQH2ULHQWDWLRQ$QJOH ƒ 3ODQH2ULHQWDWLRQ$QJOH ƒ 3ODQH2ULHQWDWLRQ$QJOH ƒ

Figure 19. Effect of initial residual stress on σn,max and the fatigue damage (Tout = 800 N·m).

4.2. Effect of Initial Residual Stress on Plastic Strain


The equivalent plastic strain is chosen as the evaluation index for the analysis of contact
fatigue damage. Figure 20 illustrates the variation of the maximum equivalent plastic strain
(Max PEEQ) corresponding to different output torques in the RCF cycles. The Max PEEQ
is close to 0 when the output torque is small, indicating that a pure elastic response exists
under this circumstance. It is evident that the Max PEEQ increases remarkably with the
increase of output torque. It can be concluded that heavy load conditions would aggravate
the rate of plastic strain accumulation, resulting in the redistribution of stress.




Tout 1āP
0D[LPXPHTXLYDOHQWSODVWLFVWUDLQ

 Tout 1āP


Tout 1āP
 Tout 1āP
Tout 1āP











      
1XPEHURI5&)F\FOHV ȱ
Figure 20. Effect of output torque on the maximum equivalent plastic strain (Max PEEQ).

Figure 21 demonstrates the variation of the maximum equivalent plastic strain (Max
PEEQ) corresponding to different residual stress cases in the RCF cycles when the output
torque is set as Tout = 1800 N·m and the maximum contact stress Ph = 2.34 GPa. Compared
with the case without IRS, ICRS reduces PEEQ by 6.3%, while ITRS increases PEEQ by
37.5%. The increase rate of PEEQ for ITRS is obviously greater than the decrease rate
of PEEQ for ICRS. RCF behaviors such as pitting and spalling more likely appear near
the surface owing to the position of Max PEEQ near to the contact surface with ITRS.

180
Materials 2022, 15, 7850

Meanwhile, the presence of ICRS makes the position where the maximum equivalent
plastic strain appear deeper. It is also found in [14] that the compressive residual stresses
make the crack initiation depth deeper. Because the crack initiation depth determines the
crack propagation length required to reach the surface, the ICRS enhanced the fatigue life
by prolonging the number of cycles of cracks reaching the surface.


Tout 1āP
D3((4PD[ ȝP

0D[LPXPHTXLYDOHQWSODVWLFVWUDLQ

D3((4PD[ ȝP

D3((4PD[ ȝP


:LWK,756VUVXUIDFH 03D

:LWKRXW,56
:LWK,&56VUVXUIDFH 03D



      
1XPEHURI5&)F\FOHV ȱ
Figure 21. The effect of initial residual stresses on the maximum equivalent plastic strain (Max PEEQ)
in RCF cycle (Tout = 1800 N·m, Ph = 2.34 GPa).

The initial residual stress also significantly affects the scope of the plastic region.
Figure 22 depicts the variation of the maximum normal stress corresponding to different
residual stress cases when the output torque is set as Tout = 2000 N·m (Ph = 2.47 GPa). It can
be clearly seen from the figure that the plastic region forms at a certain depth of material
rather than initiates on the surface. Under the same loading condition, the presence of
ICRS shrinks the plastic region, while ITRS expands the plastic region compared with
that without IRS. Almost the same maximum normal stress in the plastic region (around
−750 MPa) are found for all the three scenarios. However, the distribution and magnitude
of the maximum normal stress at a depth of 40 μm beneath the surface are totally different.
The maximum normal stress in this region for ITRS is the largest, while the increase of the
maximum normal stress is inhibited by the ICRS.

:LWK,&56VUVXUIDFH 03D 03D :LWKRXW,56 03D :LWK,756VUVXUIDFH 03D 03D


  
VQPD[ 
   
3ODVLF5HJLRQ 3ODVLF5HJLRQ 3ODVLF5HJLRQ

  
'HSWK ȝP


  


   

   


                    
3ODQH2ULHQWDWLRQ$QJOH ƒ 3ODQH2ULHQWDWLRQ$QJOH ƒ 3ODQH2ULHQWDWLRQ$QJOH ƒ

Figure 22. The effect of residual stress on the plastic region scope (Tout = 2000 N·m, Ph = 2.47 GPa).

4.3. Effect of Initial Residual Stress on Fatigue Life


The von Mises stress field and fatigue life corresponding to different initial residual
stress is further investigated in this section. Figure 23 shows the von Mises stress field at
the contact center of the x–z section when different initial residual stress states are applied.
It is worth noting that the SPRS represents the residual stress introduced by shot peening

181
Materials 2022, 15, 7850

(as shown in Figure 11). The maximum von Mises stress under the contact center in IRS
state is 615.9 MPa, which is 30.4% lower than that without IRS, while the von Mises stress
increase amplitude in ITRS state is 30.7% higher than that without IRS. What is more
noteworthy is that there is an obvious low stress region under the contact center with the
SPRS state. The maximum von Mises stress is an effective criterion to completely avoid the
occurrence of plastic deformation. The significant reduction of the maximum von Mises
stress means that the occurrence of plastic deformation can be suppressed or delayed [17].
The application of appropriate initial compressive residual stress can significantly reduce
the plastic deformation and thus improve the contact fatigue performance.

ȱ
Figure 23. The effect of initial residual stress on the von Mises stress field (Ph = 1.56 GPa).

Figure 24 illustrates the minimum fatigue life, namely the fatigue limit on the basis
of the modified FS criterion when the output torque is set as 800 N·m. Under the rated
load condition, the minimum contact fatigue life with the ICRS case is 9.08 × 107 , which is
32.7% larger than that without IRS, while the fatigue life with ITRS is 55.3% lower than that
without IRS. It is apparent that the fatigue life after applying residual stress introduced
by shot peening is increased by 125.7% compared with the state with ICRS. Therefore,
the fatigue life can be considerably improved by designing an appropriate shot peening
process and optimizing residual stress distribution.

:LWK6356VUVXUIDFH 03D
î
:LWK,&56VUVXUIDFH 03D
:LWKRXW,56
:LWK,756VUVXUIDFH 03D
 î
)DWLJXHOLIHNf

î

î

 ȱ
Figure 24. Variation of contact fatigue life with different initial residual stress states (Ph = 1.56 GPa).

182
Materials 2022, 15, 7850

Based on the modified Fatemi–Socie (FS) multiaxial fatigue criterion, the fatigue life
forecasting has considerably improved by taken hardness gradients and initial residual
stress into account especially under multiaxial loading conditions. Under the rated output
torque of 800 N·m, the experimental fatigue life of the RV reducer is 2150 h (7.55 × 107 ).
The error between the predicted fatigue life (9.08 × 107 )and the experimental value is 20.3%.
Considering the complex environmental conditions and the fatigue crack growth path
in the actual work process of the RV reducer, the calculation results under this error are
acceptable. However, there are still many shortcomings in the proposed method, such as
the wear evolution behavior of morphology during service, contact pressure distribution
under lubrication conditions, cumulative calculation of fatigue damage, etc. Fatigue life
assessment would be more challenging considering the coupling effect of these factors.

5. Conclusions
In this work, focusing on fatigue life assessment of an RV reducer crankshaft, an
innovative FEM-based three-dimensional elasto-plastic contact model is proposed. The
RCF characteristics of the crankshaft are investigated by applying the modified Fatemi–
Socie (FS) multiaxial fatigue criterion. Some distinct features can be highlighted, as follows:
(1) The location of the maximum shear strain depends on the friction coefficient. When
the friction coefficient is low, the position of the maximum shear strain is still on the
subsurface. Meanwhile, the fatigue initial location moves from the subsurface to the
surface, and the fatigue damage increases as the friction coefficient μ increases.
(2) The initial residual stress plays an influential role in fatigue damage and crack initia-
tion depth by altering the distribution of the maximum normal stress, σn,max , near the
contact surface. The compressive residual stress can reduce σn,max by 48.4% compared
with that without residual stress. Therefore, the ratio of maximum normal stress to
shear stress in the modified FS fatigue criterion can better consider the interaction
effect between the residual stress and the shear stress, which significantly improves
the prediction accuracy of the contact fatigue life model.
(3) The ICRS makes the plastic region shrink and improves the contact fatigue perfor-
mance by delaying the time of cracks propagating to the surface. Under the rated
load condition, the minimum contact fatigue life with ICRS is 9.03 × 107 , which is
29.6% larger than that without IRS, while the minimum contact fatigue life with SPRS
is 125.7% larger than that with ICRS. Residual stress distribution introduced by shot
peening significantly enhances the fatigue life of the crankshaft.
(4) Moreover, the fatigue life could be maximized by designing appropriate shot peening
parameters to obtain optimized residual stress distribution. The experimental verifica-
tion of the proposed fatigue life assessment method can also be conducted in future
study though it is costly and time-consuming.

Author Contributions: Conceptualization, X.L. and W.S.; methodology, X.L.; software, X.L.; formal
analysis, W.S. and J.T.; investigation, J.T. and H.D.; data curation, W.Z.; writing—original draft
preparation, X.L.; writing—review and editing, X.L., W.S., J.T., H.D. and W.Z.; visualization, X.L. and
W.S.; supervision, W.S.; project administration, J.T. and W.S.; funding acquisition, J.T. and W.S. All
authors have read and agreed to the published version of the manuscript.
Funding: This research was funded by the National Key R&D Program of China (No. 2019YFB2004700),
the State Key Laboratory of High Performance Complex Manufacturing (No. ZZYJKT2019-08); and
the China Postdoctoral International Exchange Program (No. 140050004).
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Not applicable.
Conflicts of Interest: The authors declare no conflict of interest.

183
Materials 2022, 15, 7850

References
1. Pham, A.-D.; Ahn, H.-J. High Precision Reducers for Industrial Robots Driving 4th Industrial Revolution: State of Arts, Analysis,
Design, Performance Evaluation and Perspective. Int. J. Precis. Eng. Manuf.-Green Technol. 2018, 5, 519–533. [CrossRef]
2. Han, L.; Guo, F. Global sensitivity analysis of transmission accuracy for RV-type cycloid-pin drive. J. Mech. Sci. Technol. 2016, 30,
1225–1231. [CrossRef]
3. Zhang, Z.; Wang, J.; Zhou, G.; Pei, X. Analysis of mixed lubrication of RV reducer turning arm roller bearing. Ind. Lubr. Tribol.
2018, 70, 161–171. [CrossRef]
4. Xu, L.; Yang, Y. Dynamic modeling and contact analysis of a cycloid-pin gear mechanism with a turning arm cylindrical roller
bearing. Mech. Mach. Theory 2016, 104, 327–349. [CrossRef]
5. Xu, L.X.; Chen, B.K.; Li, C.Y. Dynamic modelling and contact analysis of bearing-cycloid-pinwheel transmission mechanisms
used in joint rotate vector reducers. Mech. Mach. Theory 2019, 137, 432–458. [CrossRef]
6. Wang, H.; Shi, Z.-Y.; Yu, B.; Xu, H. Transmission Performance Analysis of RV Reducers Influenced by Profile Modification and
Load. Appl. Sci. 2019, 9, 4099. [CrossRef]
7. Deng, F.; Li, K.; Hu, X.; Jiang, H.; Huang, F. Life calculation of angular contact ball bearings for industrial robot RV reducer. Ind.
Lubr. Tribol. 2019, 71, 826–831. [CrossRef]
8. Li, S.; Kahraman, A.; Klein, M. A fatigue model for spur gear contacts operating under mixed elastohydrodynamic lubrication
conditions. J. Mech. Design 2012, 134, 041007. [CrossRef]
9. Liu, J.; Lv, X.; Wei, Y.; Pan, X.; Wang, Y. A novel model for low-cycle multiaxial fatigue life prediction based on the critical
plane-damage parameter. Sci. Prog. 2020, 103, 003685042093622. [CrossRef]
10. Vijay, A.; Sadeghi, F. A continuum damage mechanics framework for modeling the effect of crystalline anisotropy on rolling
contact fatigue. Tribol. Int. 2019, 140, 105845. [CrossRef]
11. Li, F.; Hu, W.; Meng, Q.; Zhan, Z.; Shen, F. A new damage-mechanics-based model for rolling contact fatigue analysis of cylindrical
roller bearing. Tribol. Int. 2018, 120, 105–114. [CrossRef]
12. Shen, F.; Zhou, K. An elasto-plastic-damage model for initiation and propagation of spalling in rolling bearings. Int. J. Mesh. Sci.
2019, 161–162, 105058. [CrossRef]
13. Guo, Y.B.; Warren, A.W. The impact of surface integrity by hard turning vs. grinding on fatigue damage mechanisms in rolling
contact. Surf. Coat. Technol. 2008, 203, 291–299. [CrossRef]
14. Choi, Y. A study on the effects of machining-induced residual stress on rolling contact fatigue. Int. J. Fatigue 2009, 31, 1517–1523.
[CrossRef]
15. Liu, H.; Liu, H.; Zhu, C.; He, H.; Wei, P. Evaluation of Contact Fatigue Life of a Wind Turbine Gear Pair Considering Residual
Stress. J. Tribol. 2018, 140, 041102. [CrossRef]
16. Wang, W.; Liu, H.; Zhu, C.; Bocher, P.; Liu, H.; Sun, Z. Evaluation of Rolling Contact Fatigue of a Carburized Wind Turbine Gear
Considering the Residual Stress and Hardness Gradient. J. Tribol. 2018, 140, 061401. [CrossRef]
17. Mahdavi, H.; Poulios, K.; Kadin, Y.; Niordson, C.F. Finite element study of cyclic plasticity near a subsurface inclusion under
rolling contact and macro-residual stresses. Int. J. Fatigue 2021, 143, 105981. [CrossRef]
18. Ooi, G.T.C.; Roy, S.; Sundararajan, S. Investigating the effect of retained austenite and residual stress on rolling contact fatigue of
carburized steel with XFEM and experimental approaches. Mater. Sci. Eng. A 2018, 732, 311–319. [CrossRef]
19. Guan, J.; Wang, L.; Mao, Y.; Shi, X.; Ma, X.; Hu, B. A continuum damage mechanics based approach to damage evolution of M50
bearing steel considering residual stress induced by shot peening. Tribol. Int. 2018, 126, 218–228. [CrossRef]
20. Walvekar, A.A.; Sadeghi, F. Rolling contact fatigue of case carburized steels. Int. J. Fatigue 2017, 95, 264–281. [CrossRef]
21. Zhang, S.; Wang, W.; Zhang, H.; Zhao, Z. The effect of hardness distribution by carburizing on the elastic–plastic contact
performance. Tribol. Int. 2016, 100, 24–34. [CrossRef]
22. Wang, W.; Liu, H.; Zhu, C.; Tang, J.; Jiang, C. Evaluation of contact fatigue risk of a carburized gear considering gradients of
mechanical properties. Friction 2019, 8, 1039–1050. [CrossRef]
23. Huang, J.; Li, C.; Chen, B. Optimization Design of RV Reducer Crankshaft Bearing. Appl. Sci. 2020, 10, 6520. [CrossRef]
24. Harris, T.A.; Kotzalas, M.N. Advanced Concepts of Bearing Technology: Rolling Bearing Analysis; CRC Press: Boca Raton, FL,
USA, 2006.
25. Lang, O.R.; Kernen i, R. Dimensionierung komplizierter Bauteile aus Stahl im Bereich der Zeit-und Dauerfestigkeit. Materialwiss.
Werkst. 1979, 10, 24–29. [CrossRef]
26. Thomas, J. Flankentragfähigkeit und Laufverhalten von Hartfeinbearbeiteten Kegelrädern. Ph.D. Thesis, Technische Universität
München, Munich, Germany, 1998.
27. Cahoon, J.; Broughton, W.; Kutzak, A. The determination of yield strength from hardness measurements. Metall. Trans. 1971, 2,
1979–1983. [CrossRef]
28. Pavlina, E.J.; Van Tyne, C.J. Correlation of Yield Strength and Tensile Strength with Hardness for Steels. J. Mater. Eng. Perform.
2008, 17, 888–893. [CrossRef]
29. MackAldener, M.; Olsson, M. Tooth interior fatigue fracture—Computational and material aspects. Int. J. Fatigue 2001, 23, 329–340.
[CrossRef]
30. Hertter, T. Rechnerischer Festigkeitsnachweis der Ermüdungstragfähigkeit Vergüteter und Einsatzgehärteter Stirnräder. Ph.D.
Thesis, Technische Universität München, Munich, Germany, 2003.

184
Materials 2022, 15, 7850

31. Réti, T. Residual stresses in carburized, carbonitrided, and case-hardened components. In Handbook of Residual Stress and
Deformation of Steel; ASM International: Almere, The Netherlands, 2002; pp. 189–208.
32. Zhou, W.; Tang, J.; Shao, W.; Wen, J. Towards understanding the ploughing friction mechanism in ultrasonic assisted grinding
with single grain. Int. J. Mesh. Sci. 2022, 222, 107248. [CrossRef]
33. You, S.; Tang, J.; Zhou, W.; Zhou, W.; Zhao, J.; Chen, H. Research on calculation of contact fatigue life of rough tooth surface
considering residual stress. Eng. Fail. Anal. 2022, 140, 106459. [CrossRef]
34. Zhao, J.; Tang, J.; Zhou, W.; Jiang, T.; Liu, H.; Xing, B. Numerical modeling and experimental verification of residual stress
distribution evolution of 12Cr2Ni4A steel generated by shot peening. Surf. Coat. Technol. 2022, 430, 127993. [CrossRef]
35. Johnson, K.L. Contact Mechanics; Cambridge University Press: Cambridge, UK, 1987.
36. Slack, T.; Sadeghi, F. Explicit finite element modeling of subsurface initiated spalling in rolling contacts. Tribol. Int. 2010, 43,
1693–1702. [CrossRef]
37. Bomidi, J.A.R.; Sadeghi, F. Three-Dimensional Finite Element Elastic–Plastic Model for Subsurface Initiated Spalling in Rolling
Contacts. J. Tribol. 2014, 136, 011402. [CrossRef]
38. Long, S.-L.; Liang, Y.-L.; Jiang, Y.; Liang, Y.; Yang, M.; Yi, Y.-L. Effect of quenching temperature on martensite multi-level
microstructures and properties of strength and toughness in 20CrNi2Mo steel. Mater. Sci. Eng. A 2016, 676, 38–47. [CrossRef]
39. Jiang, T.; Zhou, W.; Tang, J.; Zhao, X.; Zhao, J.; Liu, H. Constitutive modelling of AISI 9310 alloy steel and numerical calculation of
residual stress after shot peening. Int. J. Impact Eng. 2022, 166, 104235. [CrossRef]
40. Fatemi, A.; Socie, D.F. A critical plane approach to multiaxial fatigue damage including out-of-phase loading. Fatigue Fract. Eng.
Mater. Struct. 1988, 11, 149–165. [CrossRef]
41. Bannantine, A.; Socie, D. A Multiaxial Fatigue Life Estimation. Adv. Fatigue Lifetime Predict. Tech. 1992, 1122, 249.
42. Bäumel, A.J.; Seeger, T. Materials data for cyclic loading. Mater. Sci. Monogr. 1990, 61, 1076.
43. Shamsaei, N.; Fatemi, A. Effect of hardness on multiaxial fatigue behaviour and some simple approximations for steels. Fatigue
Fract. Eng. Mater. Struct. 2009, 32, 631–646. [CrossRef]
44. Dowling, N.E. Mechanical Behavior of Materials: Engineering Methods for Deformation, Fracture, and Fatigue; Pearson: London,
UK, 2012.
45. Gates, N.R.; Fatemi, A. On the consideration of normal and shear stress interaction in multiaxial fatigue damage analysis. Int. J.
Fatigue 2017, 100, 322–336. [CrossRef]
46. Hotait, M.A.; Kahraman, A. Estimation of Bending Fatigue Life of Hypoid Gears Using a Multiaxial Fatigue Criterion. J. Mech.
Design 2013, 135, 101005. [CrossRef]
47. Susmel, L. A simple and efficient numerical algorithm to determine the orientation of the critical plane in multiaxial fatigue
problems. Int. J. Fatigue 2010, 32, 1875–1883. [CrossRef]
48. Ding, H.; Zhang, Y.; Li, H.; Rong, K.; Tang, J.; Chen, S. Bending fatigue life oriented tooth flank dry-grinding tool modification for
cleaner manufacturing of spiral bevel gear product. J. Clean. Prod. 2021, 328, 129566. [CrossRef]
49. Chen, Z.; Jiang, Y.; Tong, Z.; Tong, S. Residual Stress Distribution Design for Gear Surfaces Based on Genetic Algorithm
Optimization. Materials 2021, 14, 366. [CrossRef] [PubMed]
50. Wang, W.; Liu, H.; Zhu, C.; Du, X.; Tang, J. Effect of the residual stress on contact fatigue of a wind turbine carburized gear with
multiaxial fatigue criteria. Int. J. Mesh. Sci. 2019, 151, 263–273. [CrossRef]

185
materials
Article
Analysis of the Failure Process of Elements Subjected to
Monotonic and Cyclic Loading Using the Wierzbicki–Bai Model
Urszula Janus-Galkiewicz and Jaroslaw Galkiewicz *

Faculty of Mechatronics and Mechanical Engineering, Kielce University of Technology, 25-314 Kielce, Poland;
[email protected]
* Correspondence: [email protected]; Tel.: +48-41-342-4711

Abstract: This article presents the results of a simulation in which smooth cylindrical and ring-
notched samples were subjected to monotonic and fatigue loads in an ultra-short-life range, made of
Inconel 718 super alloy. The samples displayed different behaviors as a result of different geometries
that introduced varying levels of stress triaxiality and loading methods. The simulations used the
Wierzbicki–Bai model, which took into account the influence of stress tensors and stress-deviator in-
variants on the behavior of the material. The difference in the behaviors of the smoothed and notched
specimens subjected to tensile and fatigue loads were identified and described. The numerical results
were qualitatively supported by the results of the experiments presented in the literature.

Keywords: Wierzbicki–Bai model; fatigue; tensile test; stress triaxiality; Lode parameter

1. Introduction
Citation: Janus-Galkiewicz, U.; Understanding the processes that occur in structural elements allows one to design
Galkiewicz, J. Analysis of the Failure more durable parts. The finite element method is currently the basic tool for analyzing the
Process of Elements Subjected to influence of various factors on material behavior. In the case of slowly changing monotonic
Monotonic and Cyclic Loading Using loads, it seems that we obtain reliable results, but there are still problems with both the
the Wierzbicki–Bai Model. Materials
calculations and the interpretation of the results for other types of loads. Methods to speed
2021, 14, 6265. https://doi.org/
up the calculation time [1] and to take into account various aspects of the loading process,
10.3390/ma14216265
such as dynamic loading [2] and the appearance of cracks [3], which are the final feature of
the critical cross-section, are constantly being developed.
Academic Editor: Andrea Spagnoli
A fundamental problem with these methods is the evaluation of the strength of the
material. The static tensile test, in many cases, does not give sufficient answers to the
Received: 24 September 2021
Accepted: 18 October 2021
question of when the failure will occur. When assessing the strength of elastic–plastic
Published: 21 October 2021
materials, the basic problem is with how to determine the moment of transition into a
plastic state. Many models can be used for this; the most popular models are Mises–Huber,
Publisher’s Note: MDPI stays neutral
Tresca, Drucker–Prager, and Mohr–Coulomb. They can be used with greater or lesser
with regard to jurisdictional claims in
success for various types of materials, although they have their limitations. For example,
published maps and institutional affil- the Mises–Huber model is not sensitive to stress triaxiality and the Lode parameter, whereas
iations. the Drucker–Prager model considers the fact that the material behaves differently under
tension and compression, which means that it is sensitive to stress triaxiality. The influence
of the Lode parameter can be found in the Tresca and Mohr–Coulomb models. The
experiments that were initiated by Wierzbicki and Bai [4], and later conducted by many
Copyright: © 2021 by the authors.
others, have shown [5,6] that the level of plastic deformation at the moment of failure of
Licensee MDPI, Basel, Switzerland.
the tested element depends on the stress triaxiality and the Lode parameter. On this basis,
This article is an open access article
a new model of plasticity was proposed. Numerous examples of the use of this model have
distributed under the terms and shown very good compliance between the numerical simulations and the experimental
conditions of the Creative Commons results. The model has previously been used for monotonic loads, but recently articles in the
Attribution (CC BY) license (https:// field of fatigue (so far, of ultra-low cycles) have started to appear, presenting very reliable
creativecommons.org/licenses/by/ results. In [7], the Wierzbicki–Bai model was used to reproduce measurable displacement
4.0/). and force parameters in the range of monotonic tensile or compression tests of samples

Materials 2021, 14, 6265. https://doi.org/10.3390/ma14216265 187 https://www.mdpi.com/journal/materials


Materials 2021, 14, 6265

with different geometries, which can be characterized by the triaxial level and the Lode
parameter. The above-mentioned model was used in [8,9] to analyze the fatigue loading
process in the range of up to 100 cycles.
The model itself has been gradually developed for several years [10–13]. The yield cri-
terion was extended to take into account the influence of triaxiality and the Lode parameter,
as well as the numerical application and the parameters supporting the calculation process.
In the present work, our attention was focused on analyzing the load process of smooth
cylindrical specimens and specimens with a ring notch, and tracing the development of the
damage in the material, as the accumulated effective plastic strain properly normalized. As
a result, the influence of the notch-root radius on the behavior of the specimens that were
subjected to a tensile load was identified and described, and the behavior of specimens
subjected to a fatigue load was compared with those under a tensile load. All results were
verified by experimental data.
The structure of the article is as follows: in the first section, the theoretical background
is provided; then, the materials and calculation methods are discussed; in the third section,
the achieved results are presented; and in the last section, our results and conclusions
are discussed.

2. Theoretical Background
Stress triaxiality is understood to mean the ratio of hydrostatic stress to effective stress,
that is, the ratio of the first stress tensor invariant by a function that depends only on the
second invariant of the deviatoric stress tensor [14].
σm
η= (1)
σe

The Lode parameter is slightly more complex to explain [15]. If the greatest tangential
stress expressed by principal stresses is τ = (σI − σI I I )/2 and the normal stress in the
plane of maximum tangential stress is equal to σN = (σI + σI I I )/2, the Lode parameter
can be written by the equation L = (σI I − σN )/τ [16]. Although it is not visible at first
glance, this parameter is related to the third invariant of the deviatoric stress tensor. Let us
introduce the parameter denoting the normalized third invariant of the deviatoric stress
tensor [17]:
27 J3
ξ= (2)
2 σe3
This quantity is related to the Lode angle, determined on the deviatoric plane accord-
ing to the equation:
ξ = cos(3θ ) (3)
and with the Lode parameter L:
 
3
ξ = L 9 − L2 / ( L2 + 3) (4)

Knowing the value of the effective stress, the Lode angle, and the stress triaxiality, it is
possible to uniquely describe the stress state, as shown in Figure 1.
On the deviatoric plane, the yield surface is reduced to a circle (Figure 1a). However,
the triaxiality of the stress affects its radius, and the Lode angle affects the position of the
point on the circle that defines the load state. Moreover, the Lode angle may influence
the shape of the yield surface [12]. It does not necessarily have to be a circle; this shape
should be determined and verified experimentally. Wierzbicki et al., therefore, proposed
universal functions based on both parameters. This function consists of two terms separated
by variables.

188
Materials 2021, 14, 6265

Figure 1. Components of the stress tensor in the principal stress space (a). Components of the stress tensor on the deviatoric
plane (b).

The impact of stress triaxiality is expressed by the function:

f ( η ) = 1 − c η ( η − η0 ) (5)

where cη is the coefficient of stress triaxiality depending on plasticity, η 0 is the reference


value of the stress triaxiality, and η is the current value of the stress triaxiality.
The influence of the Lode angle on the shape of the yield surface is more challenging.
The function that describes this impact has evolved as new, more complete experimental
results have appeared, and finally reached the form:



m+1 γ m +1
f (θ ) = csθ + (cθax − csθ ) γ− (6)
m m+1
 
cos(π/6)
where γ = 1−cos(π/6) cos(θ − 1
π/6)
− 1 . The values of “c” can be treated as material con-
stants, but in the fullest version they are described by functions:

ctθ θ ≥ 0
cθax = (7)
ccθ θ < 0
√    
csθ = 3/2 + B1 e− B2 ε pl f θ (8)
    B  B4
f θ = 1 − θ  3 (9)

where B parameters are the quantities selected, so that the simulation results are as close
as possible to the experimental results. Their task is to consider the influence of large
deformations and to correct the influence of the Lode angle. The constants ctθ and ccθ allow
the distinction between compression and tension.
It should be emphasized that Equation (6), proposed by Wierzbicki, was not the
only function used. In the papers inspired by the works of Wierzbicki et al. [16,18],
other functions were used based on the parameters (σe f f , θ, η), creating the so-called
Haigh–Westergaard space.

189
Materials 2021, 14, 6265

Taking into account Equations (5) and (6), the plasticity function proposed by Wierzbicki et al.
can be written as [7]:
 


  ! m+1 γ m +1
σ ε pl , η, θ = σ ε p 1 − cη (η − η0 ) csθ + (cθax − csθ ) γ− (10)
m m+1
 
where σ ε p is the effective-stress value read from the tensile diagram presented in the
logarithmic strain-true stress system. The terms (m + 1)/m and γm+1 /(m + 1) have been
added to facilitate numerical calculations.
An example of the results of plasticity surface modification through Equations (5) and (6)
is presented in Figure 2.

Figure 2. Yield surfaces for various yield criteria.

The fully developed Wierzbicki–Bai model contains many material constants, which
can be considered a weakness; however, these constants are easily determinable, and the
plasticity function itself can be used in a simplified form.
Research shows that in the case of elastic–plastic metals, both quantities, that is, the
Lode parameter and the stress triaxiality, play essential roles when the yield conditions
are analyzed. Triaxiality controls the void growth [19–21], whereas the Lode parameter
is associated with a change in the shape of the growing voids [22–24]. As a result, they
influence the critical strain, creating a fracture locus. An example diagram showing the
dependence of the strain at the critical moment of triaxiality and the Lode angle is shown
in Figure 3.

Figure 3. Exemplary fracture locus (source: own simulations).

190
Materials 2021, 14, 6265

3. Material
The material used in the work was Inconel 718 alloy, which is a nickel–chromium
alloy characterized by its high resistance to corrosion and creep. The chemical composition
of the alloy is presented in are given in Table 1 in [7].
This material was selected since the parameters required to fully characterize it in the
Wierzbicki–Bai model are readily available in the literature, and the process of obtaining
them is described in detail in [7]. The tensile curve was reconstructed with the help of
Equation (4), and is presented in Figure 4.

Figure 4. Comparison of engineering and true stress–strain curves (source: own computations).

After reaching the yield point, the material was described with the Ludwik curve in
the form:   −n
σ ε pl = σ0 + K ε pl (11)

The entire true stress–strain curve was represented by a set of material constants:
Young’s modulus E = 200,000 MPa, Poisson’s ratio ν = 0.284, yield stress σ0 = 945.1 MPa,
and parameters of the Ludwik curve K = 835.4 MPa and n = 0.425.
However, the numerical calculations required a much wider set of necessary data to
determine the yield surface. These constants are given in Table 5 in [7].
It is worth emphasizing at this point that the symmetry of the yield locus concerning
tension and compression was adopted.
Another problem was determining the moment of material failure. The model should
follow the experimental data; however, the damage can be modeled in various ways [25].
The parameter D was used for this, and calculated according to the equation:

ε
" pl dε pl
D=   (12)
ε f η, θ
0

where ε pl is the equivalent plastic strain. The critical strain ε f depended on the stress
triaxiality and the Lode parameter. Its full analytical form was described in [7]. The easiest
way to create it was to fit the experimental data as described in article [26]. The 3D fracture
locus in this case was constructed as follows:

ε f = ( N 1 , 1 η 2 + N 1 , 2 η + N 1 , 3 )θ 3 +
( N 2 , 1 η 2 + N 2 , 2 η + N 2 , 3 )θ 2 +
(13)
( N 3 , 1 η 2 + N 3 , 2 η + N 3 , 3 )θ +
( N4 , 1 η2 + N4 , 2 η + N4 , 3 )

191
Materials 2021, 14, 6265

where the table of coefficients had the form of Table 1. The fracture locus is shown in
Figure 5. To decrease the dynamic effects after crack initiation, an additional “softening”
function was introduced: 
σ0 0≤D<1
σo = (14)
βσ0 1 ≤ D ≤ Dc
where the softening factor β is:
w
Dc − D
β= (15)
Dc − 1

Table 1. Coefficients of the fracture locus (source: own computations).

−0.4773 0.319 −0.7304


N= 0.6683 −0.5705 1.5615
−0.2423 0.514 −1.8897
−0.0526 −0.3344 1.4992

Figure 5. Simplified fracture locus.

The parameters Dc and “w” after from [16] were assumed to be equal to 1.2 and 6,
respectively. The fracture process onset when D = 1, and when D = Dc , complete split
occurred. The shape of the dependence of β parameter on the damage indicator D described
by (15) is linear for w = 1, but in general it is nonlinear.

4. Numerical Model
The geometries shown in Figure 6 were adopted for the calculations.
The specimens were modeled in such a way that the minimum diameter (critical
section) was always a circle with a diameter of 4 mm.
The calculations were performed in Abaqus/Explicit version 6.12-2, in the Linux
environment. The VUMAT procedure was used for modeling the material, as it allows
the user to program the effect of the stress triaxiality on the development of plasticity and
failure of the element.
The geometry was modeled using CAX4R linear axisymmetric elements. The critical
cross-section for each geometry was filled with elements with a size of 2 mm/32 = 0.0625 mm.
The step-in time was determined automatically by the Abaqus routine (version 6.12-2),
based on the size of the smallest element.

192
Materials 2021, 14, 6265

(a) (b) (c) (d) (e)

Figure 6. Geometry of the specimens tested in the program. R = 1.0 (a), R = 2.5 (b), R = 5.0 (c),
R = 10.0 (d), smooth round (e).

As the plasticity theory used in the paper is valid under several assumptions, that is,
the homogeneity and material isotropy, and the material is taken to be elastic–plastic with
isotropic hardening, we applied isotropic hardening.
Using the existing symmetries, a quarter of the geometry shown in Figure 6 was
always modeled.
The uniform displacement was applied to the upper edge of each model (Figure 7).

Figure 7. Boundary conditions.

5. Tensile Tests
The reference state for simulating the tensile test was a cylindrical specimen with a
diameter of 4 mm (Figure 6e). The result of the simulation was the force-displacement plot
(Figure 8), and the distribution of the effective stress and the parameter D.
From the distribution of the effective stress and the D parameter, it can be seen
(Figure 9) that in a smooth round bar the damage process began in the specimen axis. The
result of this behavior on the specimen surface was the appearance of a cup-and-cone.
Figure 10 shows how the parameter D changed at the edge and in the center of the specimen.
It shows that at low loads the failure parameter had a constant value in the cross-section,
whereas increasing the load caused a slightly faster increase in D in the specimen center.

193
Materials 2021, 14, 6265

Figure 8. Force–displacement plots for different geometries.

(a) (b)

Figure 9. Distribution of effective stress (a) and parameter D (b) just before the fracture initiation of the smooth
round specimen.
Introducing a small-radius notch influenced the behavior of the sample during the
test. Loading the specimen with a notch of the radius of 1 mm allowed us to obtain a much
higher maximum force (Figure 8), and the specimen fractured with much less displacement
(strain). This was associated with the change in the level of triaxiality in the specimen.
The changes in effective stress and the D parameter are shown in Figure 11. The
presented results show that the failure process started from the notch root. The changes in
the D parameter in the specimen center, and the notch root during loading, as shown in
Figure 12, proved that the increase in the parameter describing the level of damage was
much higher compared with the smooth round specimen. Moreover, from the beginning of
loading, this process developed more intensively in the notch root.

194
Materials 2021, 14, 6265

Figure 10. Changes in the D parameter over time for a smooth round specimen.

(a) (b)

Figure 11. Distribution of effective stress (a) and parameter D (b) just before the moment of failure of the specimen with the
notch R = 1 mm.
It may seem that such a situation will be true for every notch; however, increasing the
radius to only 2.5 mm resulted in a change in the behavior of the specimen. As it is clear
to see in Figure 6, the maximum force obtained during the tensile test was much lower,
but the critical displacement was much larger. The diagrams of effective stress and the D
parameter were much more interesting (Figure 13).
It can be seen that the courses of the mentioned parameters were irregular, and the
maximum was not at the notch root, but in the specimen center.
The plot comparing the D parameter changes in the center of the specimen and at the
notch root (Figure 14) resembled that of a smooth sample (Figure 10).
It was interesting that these changes in the behavior of the specimens subjected to
monotonic loading translated into the behavior of cyclic loading.

195
Materials 2021, 14, 6265

Figure 12. Changes in parameter D over time for a specimen with a notch R = 1.0 mm.

(a) (b)

Figure 13. The distribution of effective stress (a) and the parameter D (b) just before the moment of failure of the sample,
with the notch R = 2.5 mm.

Figure 14. Changes in parameter D over time for a specimen with a notch R = 2.5 mm.

196
Materials 2021, 14, 6265

6. Fatigue Tests
All specimens were loaded with repeated stress cycles. The first sample for fatigue
testing was a specimen with a notch radius of 2.5 mm, loaded with a sinusoidal variable
load with an amplitude of 0.2 mm and a mean displacement in a cycle of 0.2 mm. The
material response for cyclic loading is shown in Figure 15.

Figure 15. Changes in the stress–strain hysteresis loop for a specimen with a notch R = 2.5 mm.

The changes in the D parameter during loading are shown in Figure 16. It can be seen
that from the initial stage of loading until the first maximum, the specimen behaved in
accordance with the results obtained for the monotonic loading, but the first unloading
caused a change in the behavior of the specimen, and the maximum shifted towards the
notch root.

Figure 16. Changes in parameter D under load.

The tests were repeated for specimens with much larger notch radii, i.e., 5 and 10 mm;
the load amplitude remained the same. The material response is shown in Figure 17.

197
Materials 2021, 14, 6265

(a) (b)

Figure 17. Changes in the stress–strain hysteresis loops for specimens with notches R = 5.0 mm (a) and R = 10 mm (b).

For these specimens, the loading process ended with a fracture initiated in the notch
root (Figure 18).

(a) (b)

Figure 18. Changes in parameter D under a load, for specimens with notch radius R = 5.0 mm (a) and R = 10 mm (b).

The presented results showed that increasing the radius of the notch root improved
the fatigue strength, with a greater number of load cycles. During the fatigue test, at
some point the damage of the material, initially developing faster in the specimen center,
began to dominate at the notch root [27,28]. The change in the place of damage dominance
depended on the notch root radius; the later it occurred, the greater the notch radius was.
Unfortunately, the behavior of the specimen similar to this under a under a monotonic load
was not obtained. In an attempt to obtain such behavior during the fatigue loading, tests
were carried out on a smooth round specimen.
Two levels of loading were used: the standard load used for other specimens, and
a load decreased by 25%. Reducing the load increased the number of cycles from 5 to 9
(Figure 19).

198
Materials 2021, 14, 6265

(a) (b)

Figure 19. Change in the stress–strain hysteresis loop for smooth round specimens with a regular load (a) and a load
decreased by 25% (b).

Changes in the damage level at two characteristic points are shown in Figure 20.

(a) (b)

Figure 20. Changes in parameter D for a smooth round specimen with a regular load (a) and a reduced load (b).

Parameter D, in both cases, had a higher value at the specimen center during almost
the entire loading time, but in the final phase the maximum damage position changed,
probably due to the influence of the neck curvature. Determining the location of the D
maximum required us to trace the distribution of the D parameter at the last moment
before the crack initiation (Figure 21). In the case of a smooth round specimen subjected
to ultra-low-cycle fatigue, the crack initiation occurred near the outer surface, but not at
the root of the notch produced by the neck. This was confirmed by numerous examples of
fracture surfaces obtained during the research [29–31].

199
Materials 2021, 14, 6265

(a) (b)

Figure 21. Distribution of the D parameter just before the crack initiation for the regular load (a) and the decreased load (b).

Figure 22 shows how the rate of material damage accelerated with the cycle. This
diagram also reveals the asymmetry of the damage accumulation. Despite the assumption
of the symmetry of behavior concerning tension and compression, a greater increase in
damage was recorded in the case of tension.

(a) (b)

Figure 22. Increment of parameter D for subsequent cycles.

Similar behavior was also observed with the notched specimens.

7. Discussion and Conclusions


This paper used the Wierzbicki–Bai model to analyze the damage process of elements
subjected to monotonic and fatigue loads. Smooth round and notched specimens were
analyzed. For the material of the specimens, Inconel 718 was selected with the assumption
of isotropic hardening. As this assumption could be too rough, further investigations are
necessary on the combined-material hardening rule. Nevertheless, the results showed that
the Wierzbicki–Bai model captured the differences in the behavior of the specimens with
different geometries, and the development of the damage in the analyzed elements. In a
smooth round specimen subjected to tension, the greatest increase in damage occurred

200
Materials 2021, 14, 6265

in the center of the sample, demonstrating that they are harder to fracture than notched
specimens. The introduction of the small radius notch to the specimen geometry in the
tensile specimens caused a crack initiation in the root of the notch; however, a sufficiently
large increase in the notch radius caused the fracture process to initiate in the specimen
center, as in the case of a smooth round specimen. As a result, the smooth specimens
fractured more dynamically than the notched ones that we could observe during the
simulation. The results proved that it was possible to locate the hotspot of the specimen:
the place where a combination of high-effective stress, stress triaxiality, and the Lode
parameter was favorable for the initiation of cracks. This information is of great importance
for engineers, with regard to how they can change the geometry of machine members to
improve their strength.
The situation was completely different for fatigue loads. In the case of the specimens
with a notch, a fracture began at the root of the notch, regardless of the size of the radius.
In the case of smooth round specimens, the crack initiation did not occur at the specimen
center, as in the case of monotonic loading. The initiation site was close to the outer
surface, that is, near the root of the forming neck. Such a state is confirmed by numerous
experimental studies. The position of the crack initiation site in the smooth round specimen
was to some extent affected by the load level; unfortunately, it was not possible to obtain a
state in which the fatigue crack initiation would occur in the center of the specimen. The
probable cause of this was the notch influence. Even in the smooth specimens at the final
stage, one can observe a notch, as is presented in Figure 23.

Figure 23. Smooth specimen shape at the final stage.

The increase in the damage of the specimens subjected to fatigue load was much more
intense during a tensile stage in the cycle, even though the symmetry of the yield locus
was adopted, but only when the Lode parameter was considered. This proved that the
main source of the phenomenon was a different level of stress triaxiality for tension and
compression. This result shows that ultra-low-cycle fatigue can be utilized to calibrate
the value of the coefficient of stress triaxiality dependency on plasticity cη . The most
popular calibration method so far has been to compare the results of the tensile test and the

201
Materials 2021, 14, 6265

upsetting test; however, the ultra-low-cycle fatigue test would be faster and less expensive,
as it uses only a single specimen.

Author Contributions: Conceptualization, J.G.; methodology, J.G.; software, J.G.; validation, J.G.,
U.J.-G.; formal analysis, J.G., U.J.-G.; investigation, J.G., U.J.-G.; data curation, U.J.-G.; writing—original
draft preparation, J.G.; writing—review and editing, U.J.-G.; visualization, U.J.-G. All authors have
read and agreed to the published version of the manuscript.
Funding: This research received no external funding.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: No new data were created or analyzed in this study. Data sharing is
not applicable to this article.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Funari, M.F.; Lonetti, P.; Spadea, S. A crack growth strategy based on moving mesh method and fracture mechanics. Theor. Appl.
Fract. Mech. 2019, 102, 103–115. [CrossRef]
2. Amini, M.R.; Shahani, A.R. Finite element simulation of dynamic crack propagation process using an arbitrary Lagrangian
Eulerian formulation. Fatigue Fract. Eng. Mater. Struct. 2013, 36, 533–547. [CrossRef]
3. Xu, D.; Liu, Z.; Liu, X.; Zeng, Q.; Zhuang, Z. Modeling of dynamic crack branching by enhanced extended finite element method.
Comput. Mech. 2014, 54, 489–502. [CrossRef]
4. Bai, Y.; Wierzbicki, T. Application of extended Mohr-Coulomb criterion to ductile fracture. Int. J. Fract. 2010, 161, 1–20. [CrossRef]
5. Khan, A.S.; Liu, H. A new approach for ductile fracture prediction on Al 2024-T351 alloy. Int. J. Plast. 2012, 35, 1–12. [CrossRef]
6. Danas, K.; Ponte Castañeda, P. Influence of the Lode parameter and the stress triaxiality on the failure of elasto-plastic porous
materials. Int. J. Solids Struct. 2012, 49, 1325–1342. [CrossRef]
7. Algarni, M.; Bai, Y.; Choi, Y. A study of Inconel 718 dependency on stress triaxiality and Lode angle in plastic deformation and
ductile fracture. Eng. Fract. Mech. 2015, 147, 140–157. [CrossRef]
8. Algarni, M.; Choi, Y.; Bai, Y. A unified material model for multiaxial ductile fracture and extremely low cycle fatigue of Inconel
718. Int. J. Fatigue 2017, 96, 162–177. [CrossRef]
9. Algarni, M.; Bai, Y.; Zwawi, M.; Ghazali, S. Damage evolution due to extremely low-cycle fatigue for inconel 718 alloy. Metals
2019, 9, 1109. [CrossRef]
10. Bao, Y.; Wierzbicki, T. On fracture locus in the equivalent strain and stress triaxiality space. Int. J. Mech. Sci. 2004, 46, 81–98.
[CrossRef]
11. Bao, Y.; Wierzbicki, T. On the cut-off value of negative triaxiality for fracture. Eng. Fract. Mech. 2005, 72, 1049–1069. [CrossRef]
12. Xue, L.; Wierzbicki, T. Ductile fracture initiation and propagation modeling using damage plasticity theory. Eng. Fract. Mech.
2008, 75, 3276–3293. [CrossRef]
13. Kofiani, K.; Nonn, A.; Wierzbicki, T. New calibration method for high and low triaxiality and validation onSENT specimens of
API X70. Int. J. Press. Vessel. Pip. 2013, 111–112, 187–201. [CrossRef]
14. Brocks, W.; Kunecke, G. On the Influence of Triaxiality of the Stress State on Ductile Tearing Resistance. In Defect Assessment in
Components—Fundamentals and Applications (ESIS/EGF9); Blauel, J.G., Schwalbe, K.-H., Eds.; Mechanical Engineering Publications
Limited: London, UK, 1991.
15. Gao, X.; Zhang, G.; Roe, C. A study on the effect of the stress state on ductile fracture. Int. J. Damage Mech. 2010, 19, 75–94.
[CrossRef]
16. Mohr, D.; Marcadet, S.J. Micromechanically-motivated phenomenological Hosford-Coulomb model for predicting ductile fracture
initiation at low stress triaxialities. Int. J. Solids Struct. 2015, 67–68, 40–55. [CrossRef]
17. Malvern, L.E. Introduction to the Mechanics of a Continuous Medium; Prentice-Hall, Inc.: Englewood Cliffs, NJ, USA, 1969.
18. Erice, B.; Gálvez, F. A coupled elastoplastic-damage constitutive model with Lode angle dependent failure criterion. Int. J. Solids
Struct. 2014, 51, 93–110. [CrossRef]
19. Rice, J.R.; Tracey, D.M. On the ductile enlargement of voids in triaxial stress fields*. J. Mech. Phys. Solids 1969, 17, 201–217.
[CrossRef]
20. McClintock, F.A. A criterion for ductile fracture by the growth of holes. J. Appl. Mech. 1968, 35, 363–371. [CrossRef]
21. Besson, J. Continuum models of ductile fracture: A review. Int. J. Damage Mech. 2010, 19, 3–52. [CrossRef]
22. Neimitz, A.; Galkiewicz, J.; Lipiec, S.; Dzioba, I. Estimation of the onset of crack growth in ductile materials. Materials 2018, 11,
2026. [CrossRef]
23. Barsoum, I.; Faleskog, J. Rupture mechanisms in combined tension and shear-Experiments. Int. J. Solids Struct. 2007, 44, 1768–1786.
[CrossRef]

202
Materials 2021, 14, 6265

24. Barsoum, I.; Faleskog, J. Rupture mechanisms in combined tension and shear-Micromechanics. Int. J. Solids Struct. 2007, 44,
5481–5498. [CrossRef]
25. Hamdia, K.M.; Msekh, M.A.; Silani, M.; Thai, T.Q.; Budarapu, P.R.; Rabczuk, T. Assessment of computational fracture models
using Bayesian method. Eng. Fract. Mech. 2019, 205, 387–398. [CrossRef]
26. Bai, Y.; Wierzbicki, T. A new model of metal plasticity and fracture with pressure and Lode dependence. Int. J. Plast. 2008, 24,
1071–1096. [CrossRef]
27. Pereira, F.G.L.; Lourenço, J.M.; Nascimento, R.M.D.; Castro, N.A. Fracture Behavior and Fatigue Performance of Inconel 625.
Mater. Res. 2018, 21. [CrossRef]
28. Sun, C.; Song, Q. A method for predicting the effects of specimen geometry and loading condition on fatigue strength. Metals
2018, 8, 811. [CrossRef]
29. Sharma, A.; Oh, M.C.; Ahn, B. Recent advances in very high cycle fatigue behavior of metals and alloys—A review. Metals 2020,
10, 1200. [CrossRef]
30. Mahtabi, M.J.; Shamsaei, N.; Rutherford, B. Mean Strain Effects on the Fatigue Behavior of Superelastic Nitinol Alloys: An
Experimental Investigation. Procedia Eng. 2015, 133, 646–654. [CrossRef]
31. Li, X.; Xiong, S.M.; Guo, Z. Failure behavior of high pressure die casting AZ91D magnesium alloy. Mater. Sci. Eng. A 2016, 672,
216–225. [CrossRef]

203
materials
Article
Robust Determination of Fatigue Crack Propagation Thresholds
from Crack Growth Data
Josef Arthur Schönherr 1, *, Larissa Duarte 2 , Mauro Madia 2 , Uwe Zerbst 2 , Max Benedikt Geilen 1 , Marcus Klein 1
and Matthias Oechsner 1

1 Center for Structural Materials (MPA-IfW), Technical University of Darmstadt, 64283 Darmstadt, Germany;
[email protected] (M.B.G.); [email protected] (M.K.);
[email protected] (M.O.)
2 Bundesanstalt für Materialforschung und -Prüfung (BAM), Division 9.4, 12205 Berlin, Germany;
[email protected] (L.D.); [email protected] (M.M.); [email protected] (U.Z.)
* Correspondence: [email protected]; Tel.: +49-6151-16-20348

Abstract: The robust determination of the threshold against fatigue crack propagation ΔKth is of
paramount importance in fracture mechanics based fatigue assessment procedures. The standards
ASTM E647 and ISO 12108 introduce operational definitions of ΔKth based on the crack propagation
rate da/dN and suggest linear fits of logarithmic ΔK– da/dN test data to calculate ΔKth . Since these
fits typically suffer from a poor representation of the actual curvature of the crack propagation curve,
a method for evaluating ΔKth using a nonlinear function is proposed. It is shown that the proposed
method reduces the artificial conservativeness induced by the evaluation method as well as the
susceptibility to scatter in test data and the influence of test data density.

Keywords: fatigue crack propagation threshold; ISO 12108; ASTM E647; data evaluation methods;
Citation: Schönherr, J.A.; Duarte, L.;
experimental determination
Madia, M.; Zerbst, U.; Geilen, M.B.;
Klein, M.; Oechsner, M. Robust
Determination of Fatigue Crack
Propagation Thresholds from Crack
Growth Data. Materials 2022, 15, 4737. 1. Procedures for the Determination of the Fatigue Crack Propagation Threshold from
https://doi.org/10.3390/
Crack Propagation Data
ma15144737 Typically, the outcome of fatigue crack growth (FCG) tests for the determination of
the fatigue crack propagation threshold ΔKth are the crack length a and load history data
Academic Editors: Lucjan Śnieżek,
(e.g., minimum and maximum force) on dependence of the number of cycles N. Consider-
Jaroslaw Galkiewicz and Sebastian
Lipiec
ing linear elastic fracture mechanics, the stress intensity factor range ΔK can be calculated
from the load and crack length history [1–3]. By using, e.g., the numerical differentiation
Received: 1 June 2022 technique like the secant method or the incremental polynomial method [4], fatigue crack
Accepted: 1 July 2022 propagation rate da/dN can be computed from the crack length readings and the cycles
Published: 6 July 2022 count. Due to measurement inaccuracy, influence of the testing environment, material inho-
Publisher’s Note: MDPI stays neutral mogeneities and other effects, test data are always affected by scatter. Furthermore, in most
with regard to jurisdictional claims in cases, there is no distinct reading at da/dN = da/dNth,ASTM and da/dN = da/dNth,ISO ,
published maps and institutional affil- respectively. Hence, the direct determination of the corresponding stress intensity ranges
iations. ΔKth,ASTM and ΔKth,ISO is not possible. Therefore, data fitting including inter- or extrapola-
tion techniques to determine ΔKth,ASTM and ΔKth,ISO are needed, which will be discussed
in the following.
Especially in the presence of extrinsic effects, i.e., e.g., crack closure effects [5], the ac-
Copyright: © 2022 by the authors. curate determination of the fatigue crack propagation threshold ΔKth is not trivial. The goal
Licensee MDPI, Basel, Switzerland. of this contribution is to investigate different methods for the evaluation of the fatigue
This article is an open access article
crack propagation threshold and discuss their robustness and conservativeness.
distributed under the terms and
Following the two most well-known standards regarding FCG tests, namely ASTM
conditions of the Creative Commons
E647 [4] and ISO 12108 [6], ΔKth is defined as the asymptotic value of stress intensity factor
Attribution (CC BY) license (https://
range ΔK at which the fatigue crack propagation rate da/dN approaches zero. The technical
creativecommons.org/licenses/by/
(or also operational) definition of ΔKth for most materials is given at finite crack growth rates
4.0/).

Materials 2022, 15, 4737. https://doi.org/10.3390/ma15144737 205 https://www.mdpi.com/journal/materials


Materials 2022, 15, 4737

da/dNth,ASTM = 10−7 mm/cycle according to ASTM and da/dNth,ISO = 10−8 mm/cycle


according to ISO. Since neither the fatigue crack growth rate nor the fatigue crack growth
threshold stress intensity factor range can be measured directly, the data evaluation is of
paramount importance.

1.1. Procedure Suggested by Both ASTM and ISO Standards


According to ASTM and ISO standards, the threshold stress intensity factor ranges are
evaluated by determining ‘the best fit straight line’ [4,6] to log ΔK–log da/dN data,

log10 ΔK = P1 · log10 da/dN + P0 , (1)

where P0 , P1 are fitting parameters, and then calculating the stress intensity ranges corre-
sponding to da/dNth,ASTM and da/dNth,ISO , respectively. Both standards define a mini-
mum number of five data points, approximately equally spaced in da/dN. The fitting
interval includes data pairs between 10−7 mm/cycle and 10−6 mm/cycle for ASTM and
between 10−8 mm/cycle and 10−7 mm/cycle for ISO, even though both standards allow for
using additional data with lower fatigue crack propagation rates, but require documenting
the modified range within the test protocol. Since nowadays FCG tests typically yield far
more than five data points within one decade of da/dN data, there is plenty of room for
interpretation of the suggested methods. Probably, the most straightforward interpretation
(named “interpretation one” in the following) is to just take all data points within the
specified ranges (as long as they are approximately equally distributed in da/dN direc-
tion) and then identify the best fit straight line for example by utilizing the least-squares
method; see Figure 1a. Thereby, Equation (1) is fitted to the (logarithmic) test data, using
ΔK as the dependent variable (i.e., the direction of the estimated error). The optimal set
of parameters obtained by the least-squares parameter optimization returns the “best fit”.
Another interpretation (named “interpretation two” in the following) might be that one
may freely select n ≥ 5 approximately equally spaced points that lie within the defined
boundaries, for example starting with the point next to the desired threshold fatigue crack
propagation rate. Then, the fit showing the maximum Pearson correlation coefficient is
selected; see Figure 1b. The results differ and, by adding data generated at lower decades
of da/dN, even a non-conservative (= higher) FCG threshold stress intensity range might
be calculated, see Section 3.1.3. Furthermore, neither of both interpretations (nor any other
straight line) is able to reflect the curvature of the depicted test data.
da/dN in mm/cycle

da/dN in mm/cycle

10−7 10−7

Experimental dataset Experimental dataset


Fitting dataset Fitting dataset
Fitting curve Fitting curve
(Eq. (1), all data points) (Eq. (1), first n data points)
10−8 ΔKth, ASTM = 2.72 MPa · m1/2 10−8 ΔKth, ASTM = 2.77 MPa · m1/2

3 4 3 4
ΔK in MPa · m1/2 ΔK in MPa · m1/2

(a) (b)
Figure 1. FCG test conducted with conventional K-decreasing at R = 0.8 in lab air, see corresponding
paragraph in Section 3.1.1: (a) linear fit over all data points with 10−7 mm/cycle ≤ da/dN ≤
10−6 mm/cycle; (b) best linear fit over the first n data points with da/dN ≥ 10−7 mm/cycle, where
n = 112 (out of 177 within the interval) is selected yielding the highest Pearson correlation coefficient.

The results obtained following interpretation are strongly affected by the test data
curvature because of its averaging character. The higher the curvature, the more conserva-
tive is the calculated ΔKth . In contrast, interpretation two is less affected by data curvature

206
Materials 2022, 15, 4737

because of its definition using the correlation coefficient of the straight line. In case of
a more pronounced scatter in the vicinity of da/dNth , interpretation two would have
accounted for a higher number of data points and therefore be more conservative. It is
trivial to state that a straight line is unable to characterize the curvature of the crack growth
curve and therefore leads to conservative ΔKth results for strictly monotonic increasing
FCG data fitted within the recommended ranges. In turn, the evaluation error depends on
the degree of the curvature, and it is therefore sensitive to the number (and location) of data
pairs taken into account as well as to the curve’s gradient, see also [7]. Therefore—as also
mentioned in the ASTM standard [4]—a nonlinear relationship between ΔK and da/dN
might be beneficial to obtain a good fit.

1.2. Procedures Suggested in the Literature


Bucci [7] considered a four-parameter Weibull function (see also [8,9]), fitted to the
entire data set, see Figure 2a. He stated that the Weibull approach addresses nonlinearity in
a better way, but if there are enough data points in the near-threshold region, straight line
fits are performing quite well and are easier to use. In Figure 2a, the test data have been
fitted using the four-parameter Weibull function
#
$
ΔK da/dN − e k
1− = exp − (2)
Kb v−e

and ΔKth,ASTM has been determined at da/dNth,ASTM . Smith and Hoeppner [9] observed
that the least-squares method is not suitable to fit the instability parameter Kb , the thresh-
old parameter e, the characteristic value v and the shape parameter k to da/dN–ΔK data
accurately within the threshold region due to the differences in orders of magnitude of
the non-logarithmic test data (da/dN ranges between 10−7 mm/cycle and 10−6 mm/cycle
vs. ΔK ranges between 100 MPa·m1/2 and 101 MPa·m1/2 , approximately). Therefore, they
proposed to use a least-squares optimization to calculate a preliminary optimized pa-
rameter set and afterwards improve the fitting results by optimizing only e and v in an
orthogonal distance regression (ODR) [10]. Since an ODR optimization is connected to
a quite high computational cost and compared to the 1990s the computation power in-
creased dramatically until nowadays, this might have been the reason why in [9] only
the two parameters showing the major contribution with regard to the mentioned errors
were optimized using an ODR. In fact, it has been observed in this work that the curve fit
may be improved by optimizing all four parameters instead of just two in an ODR after
least-squares minimization. As one can clearly see, the overall agreement to the test data
is quite good but locally diverges slightly. This can be observed in Figure 2a as well as in
Figure 2b, where the four-parameter Weibull fit predicts a too steep curvature towards the
ISO operational threshold definition (da/dNth = 10−8 mm/cycle) and therefore leads to
non-conservative results for ΔKth,ISO . Since non-conservative values are not acceptable,
the Weibull fit has not been considered any further.
Another well established method to evaluate the threshold stress intensity range
was proposed by Döker [11]. The method uses a straight line fit applied to the (non-
logarithmic) da/dN–ΔK data in the range 5 × 10−8 mm/cycle ≤ da/dN ≤ 10−6 mm/cycle,
see Figure 3a. The threshold definition is independent from the ASTM or ISO standards,
and ΔKth is evaluated at da/dNth = 0. The resulting threshold stress intensity factor
range is located somewhere between the values one would expect using the ISO or ASTM
operational definitions. Considering that only the ASTM E647 standard [12] is cited
within the original publication, this method shows quite conservative results using the
ASTM operational definition of ΔKth . The straightforward extension to the ISO threshold
by shifting the fit range by one decade of da/dN would require test results as low as
da/dN = 5 × 10−9 mm/cycle, which would be very time consuming to obtain and there-
fore not practicable. The curve behavior is poorly described, both for linear, Figure 3a,

207
Materials 2022, 15, 4737

and double-logarithmic, Figure 3b, scaled da/dN–ΔK data. Therefore, further analyses
with this approach have also been discarded.
Furthermore, there is a multitude of different crack propagation laws aiming at describ-
ing the whole FCG data starting from the threshold regime over the range where the FCG
curve grows linear in a double-logarithmic scaled plot (also known as Paris regime) and
some even include the region of instable crack growth; see [13–19]. Several of these models
contain ΔKth as a model parameter, which should not be confused with the operational
definitions of the threshold stress intensity factor range included in the standards.

10−6
da/dN in mm/cycle

da/dN in mm/cycle
10−7
10−7

Experimental dataset Experimental dataset


Fitting dataset Fitting dataset
Fitting curve Fitting curve
(Eq. (2), Weibull) (Eq. (2), Weibull)
10−8 ΔKth, ASTM = 2.79 MPa · m1/2 ΔKth, ISO = 2.28 MPa · m1/2
10−8
3 4 3 4
ΔK in MPa · m1/2 ΔK in MPa · m1/2

(a) (b)
Figure 2. Application of a four-parameter Weibull function on all test data, referring to: (a) a threshold
definition at da/dNth,ASTM = 10−7 mm/cycle. K-decreasing test at R = 0.8 in lab air, see correspond-
ing paragraph in Section 3.1.1; (b) a slightly non-conservative FCG threshold determination at
da/dNth,ISO = 10−8 mm/cycle. Kmax = const., Rmax ≈ 0.8, lab air, see corresponding paragraph in
Section 3.1.1.

8 × 10−7

7 × 10−7

6 × 10−7
da/dN in mm/cycle
da/dN in mm/cycle

5 × 10−7
10−7
4 × 10−7

3 × 10−7
Experimental dataset (linear) Experimental dataset (log–log)
2 × 10−7 Fitting dataset Fitting dataset
Fitting curve (linear, Döker) Fitting curve (linear, Döker)
ΔKth = 2.64 MPa · m1/2 10−8 ΔKth = 2.64 MPa · m1/2
0
3 4 5 3 4
ΔK in MPa · m1/2 ΔK in MPa · m1/2

(a) (b)
Figure 3. FCG threshold determination using a linear fit, applied to regular (non-logarithmic)
test data, following [11]. K-decreasing test at R = 0.8 in lab air, see corresponding paragraph in
Section 3.1.1. (a) linear-scaled data; (b) double-logarithmic scaled data. The method clearly provides
a poor data fit and determination of ΔKth at da/dNth = 0.

2. Experimental Procedure
Since none of the methods shown in Sections 1.1 and 1.2 allowed universal application
for both standards and simultaneously led to robust and not overly conservative results,
further fitting functions have been investigated. A comprehensive set of fatigue crack
growth data recorded at BAM Berlin and MPA-IfW Darmstadt has been used to calibrate
and validate the evaluation procedure. The fitting has been performed using least squares
minimization, where the threshold stress intensity factor range has been used as dependent
variable, if not otherwise mentioned.

208
Materials 2022, 15, 4737

2.1. Investigated Fitting Functions


The most straightforward approach was to extend Equation (1) to a more general
polynomial form as
I  i
log10 ΔK = ∑ Pi · log10 da/dN , (3)
i =0

where I denotes the polynomial degree and Pi are fit parameters. For I = 1, Equation (3)
equals Equation (1). Regarding I ≥ 2, in case there are only very few (or even no) data
points available in the vicinity of da/dNth , the fit may yield a curve having an inflection
point. Hence, the fit does not represent the data in a satisfactory manner, and (for I = 2)
it is even possible that the fit does not intersect the da/dNth line. Therefore, it is not
recommended to use polynomial fits of the type given in Equation (3) with I
= 1. The use
of non-logarithmic data, like in the approach of Döker [11], leads to the same problems.
Hence, further approaches have been investigated.
The log − log data depicted in Figure 1 considered swapping the axes, suggesting a
hyperbolic trend, leading to the fit function
  −1
log10 ΔK = P1 · − log10 da/dN + P2 , (4)

with the fitting parameters P1 and P2 . Fitting Equation (4) to the dataset leads to results
comparable to the linear fit using all data points in accordance with the standards, since
the exponent −1 does not represent the curvature of the crack growth curve, see Figure 4a.
By extending Equation (4) to a variable exponent as
 − P3
log10 ΔK = P1 · − log10 da/dN + P2 , (5)

with P3 ≥ 1, a much better fitting to the test data is possible, see Figure 4b. Rounding the
fit result of P3 ≈ 4.60 to the previous and next integer number, namely P3 = 4,
  −4
log10 ΔK = P1 · − log10 da/dN + P2 , (6)

see Figure 4c and P3 = 5,


  −5
log10 ΔK = P1 · − log10 da/dN + P2 , (7)

see Figure 4d provides two straightforward variants of Equation (5) with only two free
parameters. The curvature of the fitting curve increases with increasing P3 . It is worth
noting that, although P3 = 5 leads to non-conservative estimation of the FCG threshold
according to ISO, it is still conservative following the ASTM operational definition.
da/dN in mm/cycle

da/dN in mm/cycle

10−7 10−7

Experimental dataset Experimental dataset


Fitting dataset Fitting dataset
Fitting curve Fitting curve
(Eq. (5), P3 = 1) (Eq. (5), P3 = free)
10−8 ΔKth, ASTM = 2.75 MPa · m1/2 10−8 ΔKth, ASTM = 2.80 MPa · m1/2

3 4 3 4
ΔK in MPa · m1/2 ΔK in MPa · m1/2

(a) (b)
Figure 4. Cont.

209
Materials 2022, 15, 4737

da/dN in mm/cycle

da/dN in mm/cycle
10−7 10−7

Experimental dataset Experimental dataset


Fitting dataset Fitting dataset
Fitting curve Fitting curve
(Eq. (6), P3 = 4) (Eq. (7), P3 = 5)
10−8 ΔKth, ASTM = 2.79 MPa · m1/2 10−8 ΔKth, ASTM = 2.80 MPa · m1/2

3 4 3 4
ΔK in MPa · m1/2 ΔK in MPa · m1/2

(c) (d)
Figure 4. FCG threshold evaluation according to ASTM (da/dNth,ASTM = 10−7 mm/cycle) for the
data set presented in Figure 1 using Equation (5) with fixed or free parameter P3 : (a) fixed, P3 = 1;
(b) free, optimized value P3 ≈ 4.60; (c) fixed, P3 = 4; (d) fixed, P3 = 5.

2.2. Quantitative Data Analysis


The main datasets investigated stemmed from a total of 47 specimens manufactured
from 12 mm thick S690QL hot-rolled plates, tested at BAM Berlin and MPA-IfW Darmstadt.
The materials chemical composition and mechanical properties are given in Tables 1 and 2.
The microscopic analysis on etched samples showed a fine grained quenched and tempered
martensitic-bainitic microstructure, see Figure 5.

Table 1. Chemical composition (in weight percent) obtained by means of spark optical emission spectrometry.

C Si Mn P S Cr Mo Ni Al Cu Nb Fe
0.16 0.23 1.15 0.01 < 0.01 0.41 0.18 0.04 0.08 0.02 0.04 97.65

Table 2. Mechanical properties.

KV2 in J
σy in MPa σu in MPa E in GPa A in % (Orientation:
T–L [20])
810 825 207 16 126

Figure 5. S690QL microstructure (T–L plane).

All tests were performed using SENB specimens [6] with a cross section of 19 mm ×
6 mm. The specimens were oriented, such that the direction of crack propagation is parallel
to the rolling direction, i.e., the orientation T–L according to [20]. The test data have
been obtained on three resonance testing machines equipped with an eight-point bending
fixture, see [6]. These were a RUMUL MIKROTRON 654 with a maximum load capacity of
20 kN and an average testing frequency of about 108 Hz, a RUMUL TESTRONIC with a

210
Materials 2022, 15, 4737

maximum load capacity of 100 kN and an average testing frequency of about 60 Hz, both at
BAM Berlin, and a RUMUL TESTRONIC with a maximum load capacity of 250 kN and
an average testing frequency of about 90 Hz at MPA-IfW Darmstadt. The crack length
was monitored using direct current potential drop techniques with current reversal and
active temperature compensation (current source: HP 6033A, nanovolt meter: Keithley
2182A) at BAM and specimen compliance techniques (clip-gage: Sandner EXR10-0.5o) at
MPA-IfW. The crack length was corrected a posteriori by means of optical measurements on
the broken open fracture surfaces. Then, the crack propagation rates have been calculated
using the slope of piecewise straight line fits performed on filtered test data. Each segment
of the piecewise function referred to a crack extension of 0.02 mm. The corresponding stress
intensity factors have been calculated using the formulations reported in [6].
Since in these tests the same specimen types, manufactured from the same material
batch in the same specimen orientation, have been used, a high repeatability was expected.
Consequently, the standard deviation in ΔKth calculated for each method was the result of
the data scatter within the test (stemming from small variations in environment conditions,
material inhomogeneities, specimen misalignment, errors in the calculation of da/dN, etc.)
and an error induced by the fit used to evaluate ΔKth . Since the first part is independent
of the fitting procedure, the differences in standard deviations are a measure for the fit
robustness, whereas the corresponding mean value gives information on the fit quality and
therefore the inter- and extrapolation error, respectively.

3. Results and Discussion


3.1. Application to Data Obtained at R ≈ 0.8
First, data obtained at a load ratio of approximately R ≈ 0.8, which produce only a
negligible influence of crack closure effects have been investigated.

3.1.1. Evaluation for the Intervals Suggested by the Standards


In order to assess the performance of the polynomial functions with negative exponent
Equation (5) in comparison to the fit suggested by the standards, see Equation (1), test data
obtained at either a fixed load ratio R = 0.8 (K-decreasing procedures) or at Rmax ≈ 0.8
(Kmax tests) have been considered, see ([4], Section 8.6). To minimize influences of the
extrapolation method, the smallest recorded crack propagation rate has been required to
be smaller than 1.1 · da/dNth , i.e., min(da/dN ) ≤ 1.1 × 10−8 mm/cycle for the ISO and
min(da/dN ) ≤ 1.1 × 10−7 mm/cycle for the ASTM operational definition of ΔKth . It shall
be noted that these boundaries have been used only for comparability between tests within
this work and neither define the actual application boundaries of the method regarding
data extrapolation nor represent a general recommendation.
Since the fixed parameters P3 = 4 in Equation (6) and P3 = 5 in Equation (7) provided
a better description of the data curvature and a less conservative determination of ΔKth
in addition to the more general function with variable P3 , these have been compared with
the two interpretations of the linear fit method suggested by the standards, see Figure 1a,b.
The range used to fit the data was in all cases fixed to one decade of da/dN data, starting
from da/dNth and therefore equal to the ranges suggested by the standards in order to
ensure comparability between the methods.

Conventional K-Decreasing at R = 0.8 (ΔKLR )


The first datasets investigated stemmed from a total of nine SENB specimens made
of S690QL. These tests have been conducted using the standard K-decreasing (or load
shedding) procedure suggested by the standards at constant R = 0.8, tested in lab air.
Further information on the experimental procedure may be found in [21]. All nine speci-
mens contained data points below da/dN = 1.1 × 10−7 mm/cycle and are valid for ASTM
operational threshold evaluation, and four of them were also valid according to the ISO
definition. The threshold stress intensity factor range results with corresponding standard
deviations are presented in Table 3.

211
Materials 2022, 15, 4737

Table 3. Comparison of effective threshold values for S690QL obtained with various fit methods.
Tests conducted using the K-decreasing procedure at R = 0.8 in lab air.

Method ΔKth,ASTM in MPa·m1/2 ΔKth,ISO in MPa·m1/2


Linear, all data points (Equation (1)) 2.71 ± 0.06 2.27 ± 0.04
Linear, first n data points (Equation (1)) 2.76 ± 0.05 2.29 ± 0.04
Polynomial, neg. exp. (Equation (5)) 2.78 ± 0.04 2.34 ± 0.04
Polynomial, neg. exp. P3 = 4 (Equation (6)) 2.77 ± 0.05 2.32 ± 0.04
Polynomial, neg. exp. P3 = 5 (Equation (7)) 2.78 ± 0.05 2.32 ± 0.03

The comparison between the determined threshold stress intensity factor ranges (on
single specimen basis) and the value read out from test data proved conservative for every
single method and dataset. Considering the mean value of the threshold stress intensity
ranges, both linear fits showed a more pronounced underestimation of ΔKth , inducing an
artificial conservativeness, as already observed in Section 2.1. This is proven true especially
for the linear fit Equation (1) performed on all data points. For the linear fit over the first n
points, the artificial conservativeness of ΔKth,ASTM was found to be comparatively higher
than for ΔKth,ISO . The results obtained using Equation (5) agreed fairly well, whereas the
fixed exponents P3 = 4 and P3 = 5 showed a slightly higher conservativeness for ΔKth,ISO
compared to the three-parameter version of Equation (5). The standard deviation is very
low and comparable for all tests.

Load Shedding at Constant Kmax (Rmax ≈ 0.8)


A set of nine SENB prepared from the same material batch has been investigated using a
load shedding scheme at constant Kmax with a final load ratio Rmax ≈ 0.8 at about ΔKth,ISO and
R ≈ 0.72 . . . 0.76 at ΔKth,ASTM . All nine specimens provided data for evaluating ΔKth,ASTM
and among them four were also valid for ΔKth,ISO evaluation, see Table 4.

Table 4. Comparison of effective threshold values for S690QL obtained with various fit methods.
Tests conducted using the constant Kmax procedure (Rmax ≈ 0.8) in lab air.

Method ΔKth,ASTM in MPa·m1/2 ΔKth,ISO in MPa·m1/2


Linear, all data points (Equation (1)) 2.67 ± 0.03 2.22 ± 0.01
Linear, first n data points (Equation (1)) 2.77 ± 0.03 2.24 ± 0.01
Polynomial, neg. exp. (Equation (5)) 2.79 ± 0.03 2.29 ± 0.03
Polynomial, neg. exp. P3 = 4 (Equation (6)) 2.77 ± 0.03 2.27 ± 0.02
Polynomial, neg. exp. P3 = 5 (Equation (7)) 2.79 ± 0.03 2.27 ± 0.02

Since the crack propagation data obtained from different procedures at R ≈ 0.8 nearly
coincide and exhibit a very low scatter, the same conclusion as in the case of conventional
K-decreasing tests can be drawn.

Compression Precracking Load Reduction (CPLR) at R = 0.8


A further set of four SENB extracted from the same batch has been investigated using
compression precracking followed by a K-decreasing test with a constant load ratio R = 0.8.
Applying the same criteria for selecting valid data sets for comparison returned four
specimens for evaluating ΔKth,ASTM and two for ΔKth,ISO , see Table 5. Here, the standard
deviation for ISO is omitted due to the insufficient number of available data sets that
include points below da/dN = 1.1 × 10−8 mm/cycle. The results confirm those previously
shown in Table 4.

212
Materials 2022, 15, 4737

Table 5. Comparison of threshold values for S690QL obtained with various fit methods. Tests
conducted using Compression Precracking Load Reduction at R = 0.8 in lab air.

Method ΔKth,ASTM in MPa·m1/2 ΔKth,ISO in MPa·m1/2


Linear, all data points (Equation (1)) 2.67 ± 0.02 2.22
Linear, first n data points (Equation (1)) 2.74 ± 0.03 2.23
Polynomial, neg. exp. (Equation (5)) 2.77 ± 0.04 2.29
Polynomial, neg. exp. P3 = 4 (Equation (6)) 2.73 ± 0.02 2.26
Polynomial, neg. exp. P3 = 5 (Equation (7)) 2.74 ± 0.03 2.27

Constant Force Range (ΔF-Constant) at R = 0.8


The fourth and last test to determine the threshold at R ≈ 0.8 in ambient air has been
based on another set of four SENB specimens produced from the same material batch, using
conventional precracking followed by a test at constant force amplitude (ΔF-constant) at a
load ratio R = 0.8. All four specimens have been considered valid for evaluating ΔKth,ASTM
and three for ΔKth,ISO , see Table 6. Here, the same observations as above apply.

Table 6. Comparison of threshold values for S690QL obtained with various fit methods. Tests
conducted using ΔF-constant at R = 0.8 in lab air.

Method ΔKth,ASTM in MPa·m1/2 ΔKth,ISO in MPa·m1/2


Linear, all data points (Equation (1)) 2.58 ± 0.03 2.18 ± 0.06
Linear, first n data points (Equation (1)) 2.64 ± 0.03 2.21 ± 0.06
Polynomial, neg. exp. (Equation (5)) 2.69 ± 0.02 2.31 ± 0.02
Polynomial, neg. exp. P3 = 4 (Equation (6)) 2.67 ± 0.02 2.25 ± 0.03
Polynomial, neg. exp. P3 = 5 (Equation (7)) 2.69 ± 0.02 2.25 ± 0.02

Summary
The linear fits induce artificial conservativeness in the evaluation of fatigue crack
propagation thresholds obtained at R ≈ 0.8. Nevertheless, this behavior is observed less
pronounced for the linear fit incorporating only the first n points. In all cases, the three-
parameter polynomial Equation (5) provides less conservative results. Fixing its parameter
P3 to P3 = 4 or P3 = 5 sometimes induces conservativeness, but in most of the cases less
pronounced than the linear fits.
Figure 6 summarizes the evaluation of the four datasets presented earlier in
Section 3.1.1. The threshold stress intensity factor ranges determined according to the
ASTM operational definition (see Figure 6a) as well as those following the ISO operational
definition (see Figure 6b) for the four test methods (ΔKLR , Kmax , CPLR and ΔF-constant)
agree fairly well within each data evaluation method.
3.0 3.0
Linear, all data points (Eq. (1))
Linear, first n data points (Eq. (1))
2.8 2.8 Polynomial, neg. exponent (Eq. (5))
ΔKth, ASTM in MPa · m1/2

Polynomial, neg. exponent, P3 = 4 (Eq. (6))


ΔKth, ISO in MPa · m1/2

Polynomial, neg. exponent, P3 = 5 (Eq. (7))


2.6 2.6

2.4 2.4

2.2 2.2

2.0 2.0
ΔKLR Kmax CPLR ΔF−const. ΔKLR Kmax CPLR ΔF−const.

(a) (b)
Figure 6. Fatigue crack propagation thresholds obtained at R ≈ 0.8 applying the fit methods
to the four datasets presented in Section 3.1.1: (a) according to the ASTM operational definition;
(b) according to the ISO operational definition.

213
Materials 2022, 15, 4737

3.1.2. Robustness of the Fitting Methods in Handling Data Subjected to Augmented


Artificial Scatter
In order to assess the ability and robustness of the fitting methods to handle scattered
data, artificial scatter has been added to the test data presented in Figure 1. The additional
scatter has been generated by sampling random values from a normal distribution with
a mean μ = 1 and a standard deviation SD = 0.02 and multiplying them with ΔK data,
whereas da/dN remained unchanged.
The comparison of the linear fit using all scattered data points within the defined interval
(ΔKth,ASTM = 2.69 MPa·m1/2 , Figure 7a) with the original dataset (ΔKth,ASTM = 2.72 MPa·m1/2 ,
Figure 1a) did not reveal a notable difference. The polynomials with negative expo-
nents were also almost insensitive to scatter. The three-parameter polynomial provided
ΔKth,ASTM = 2.78 MPa·m1/2 for the scattered data (Figure 7d) compared to ΔKth,ASTM =
2.80 MPa·m1/2 for the original dataset (Figure 4b). The two-parameter polynomial with
P3 = 4 showed a similar trend with ΔKth,ASTM = 2.79 MPa·m1/2 for scattered data
(Figure 7c) in comparison to ΔKth,ASTM = 2.79 MPa·m1/2 for the original dataset (Figure 4c).
da/dN in mm/cycle

da/dN in mm/cycle
10−7 10−7

Artificially scattered Artificially scattered


experimental dataset experimental dataset
Fitting dataset Fitting dataset
Fitting curve Fitting curve
(Eq. (1), all data points) (Eq. (1), first n data points)
10−8 ΔKth, ASTM = 2.69 MPa · m1/2 10−8 ΔKth, ASTM = 2.72 MPa · m1/2

3 4 3 4
ΔK in MPa · m1/2 ΔK in MPa · m1/2

(a) (b)
da/dN in mm/cycle

da/dN in mm/cycle

10−7 10−7

Artificially scattered Artificially scattered


experimental dataset experimental dataset
Fitting dataset Fitting dataset
Fitting curve (Eq. (6)) Fitting curve (Eq. (5))
10−8 ΔKth, ASTM = 2.79 MPa · m1/2 10−8 ΔKth, ASTM = 2.78 MPa · m1/2

3 4 3 4
ΔK in MPa · m1/2 ΔK in MPa · m1/2

(c) (d)
Figure 7. FCG data reported in Figure 1 with additional artificial scatter. The da/dNth,ASTM has
been evaluated using four different methods: (a) linear, all data points Equation (1); (b) linear, first
n = 176 data points (out of 177 within the interval) Equation (1); (c) polynomial, negative exponent,
P3 = 4 Equation (6); (d) polynomial, negative exponent Equation (5).

In contrast, the best linear fit over the first n points showed a pronounced sensi-
tivity to scattered data. The best fit interval coincided almost with all data points (see
Figures 7b and 1b). Consequently, ΔKth,ASTM = 2.72 MPa·m1/2 calculated for scattered
data was more conservative than ΔKth,ASTM = 2.77 MPa·m1/2 calculated in case of the
original dataset.

214
Materials 2022, 15, 4737

3.1.3. Influence of an Augmented Fit Interval


Both ASTM E647 and ISO 12108 suggest a fit interval of one decade of da/dN data,
starting from da/dNth,ASTM and da/dNth,ISO , respectively. Nevertheless, both allow for
use data obtained at lower fatigue crack propagation rates for determining the threshold
stress intensity factor range. Therefore, the impact of an augmented fit interval on the deter-
mination of fatigue crack propagation thresholds has been investigated. Since no datasets
with crack propagation rates momentously below da/dNth,ISO = 1 × 10−8 mm/cycle were
available, only the threshold following the ASTM operational definition has been consid-
ered. Therefore, the data shown in Table 3, which have been obtained using a fit interval of
10−7 mm/cycle ≤ da/dN ≤ 10−6 mm/cycle, have been compared with threshold stress
intensity factor ranges obtained with augmented intervals. The upper bound has been
held constant, whereas the lower bound has been varied from 2.5 × 10−8 mm/cycle up to
10−7 mm/cycle. The respective threshold stress intensity factor ranges have been named
as ΔKth,ASTM,2.5 , ΔKth,ASTM,5 , ΔKth,ASTM,7.5 and ΔKth,ASTM,10 , according to the lower FCG
propagation rate bounds (2.5, 5, 7.5 and 10 × 10−8 mm/cycle, see Figure 8).

3.1 Linear, all data points (Eq. (1))


Linear, first n data points (Eq. (1))
Polynomial, neg. exponent (Eq. (5))
3.0
ΔKth, ASTM in MPa · m1/2

Polynomial, neg. exponent, P3 = 4 (Eq. (6))


Polynomial, neg. exponent, P3 = 5 (Eq. (7))
2.9

2.8

2.7

2.6

2.5
ΔKth,ASTM,10 ΔKth,ASTM,7.5 ΔKth,ASTM,5 ΔKth,ASTM,2.5

Figure 8. Comparison of threshold values obtained with various fit methods using vary-
ing fit intervals from 2.5 × 10−8 mm/cycle ≤ da/dN ≤ 10−6 mm/cycle (ΔKth,ASTM,2.5 ) up to
10 × 10−8 mm/cycle ≤ da/dN ≤ 10−6 mm/cycle (ΔKth,ASTM,10 ). The results refer to eight speci-
mens made of S690QL, tested using a K-decreasing procedure at R = 0.8 in lab air.

Regarding both linear fits, there is a clear tendency that, with augmenting the fit
interval towards lower minimal crack propagation rates (displayed in Figure 8 from left
to right), there is an increase in ΔKth,ASTM and therefore a decrease in conservativeness.
In contrast, the results for the polynomial with negative exponent, Equations (5)–(7), are
almost insensitive to the interval augmentation, but for P3 = 4 and P3 = 5, a reduction
in standard deviation can be observed with increasing the interval. For Equation (6), the
optimal lower bound was found at 2.5 × 10−8 mm/cycle and 5 × 10−8 mm/cycle with
equal magnitudes in mean and standard deviation.
Using the linear functions with augmented intervals increases the risk of non-conser-
vative extrapolation, as one can see comparing the values for ΔKth,ASTM,2.5 , where the
linear functions provided the highest threshold stress intensity factor ranges among all five
methods under comparison. This issue can be clearly understood looking at the evaluation
depicted in Figure 9. In particular, Figure 9a shows that the ΔKth,ASTM calculated using
the linear fit over all data points is on the right-hand side of the dataset, i.e., in the non-
conservative region. In contrast, the polynomial with P3 = 4 does not show this issue
(Figure 9b).

215
Materials 2022, 15, 4737

da/dN in mm/cycle

da/dN in mm/cycle
10−7 10−7

Experimental dataset Experimental dataset


Fitting dataset Fitting dataset
Fitting curve Fitting curve
(Eq. (1), augmented interval) (Eq. (6), augmented interval)
10−8 ΔKth, ASTM = 2.84 MPa · m1/2 10−8 ΔKth, ASTM = 2.80 MPa · m1/2

3 4 3 4
ΔK in MPa · m1/2 ΔK in MPa · m1/2

(a) (b)
Figure 9. Influence of augmented fit interval on various fits, using an interval 2.5 × 10−8 mm/cycle ≤
da/dN ≤ 10−6 mm/cycle: (a) the linear fit using all data within the specified range leads to a non-
conservative result; (b) the polynomial fit with negative exponent P3 = 4 gives a conservative result.
The same dataset as reported in Figure 1 has been used.

3.1.4. Data Extrapolation


The investigations presented in Section 3.1.1 included only data at crack propagation
rates as low as 1.1 · da/dNth . Nevertheless, it shall be noted that no data might be available
at low crack propagation rates, especially for the ISO operational definition with a threshold
crack propagation rate as low as da/dNth,ISO = 10−8 mm/cycle. Therefore, to ensure a
reliable and conservative evaluation of the fatigue crack propagation thresholds, a robust
extrapolation technique is needed. To assess the goodness of the extrapolation, the test
results given in Section 3.1.1 have been compared to artificially censored datasets, using
only data with da/dN ≥ 3 · da/dNth . Hence, the resulting fit intervals after censoring
were 3 × 10−8 mm/cycle ≤ da/dN ≤ 10−7 mm/cycle for ISO operational definition and
3 × 10−7 mm/cycle ≤ da/dN ≤ 10−6 mm/cycle for ASTM operational definition. Since
extrapolation is very sensitive to the data range available, in order to have a reliable
comparison, the investigations have been restricted to datasets that had data within the
whole censored interval (including the upper bound). Hence, the number of tests with
valid data for ASTM threshold determination is reduced in comparison to Section 3.1.1.
The comparison has been based on the change in ΔKth induced by censoring the FCG
data. ΔKth,cens denotes the fatigue crack propagation threshold obtained for the censored
version of the data set used to evaluate ΔKth . It follows that ΔKth − ΔKth,cens values greater
than or equal to zero are considered conservative, whereas values lower than zero are non-
conservative. The minimum difference throughout all specimens shows whether all tests
are extrapolated conservatively, whereas the mean value can be regarded as an index of the
goodness of the extrapolation. The results are given in Table 7. Conservative extrapolation
has been obtained for both linear fits and for the polynomial with negative exponent fixed
to P3 = 4, whilst the versions with P3 = 5 or free P3 returned a very limited number of
negative results, meaning non-conservative extrapolation results. Even though the latter are
just slightly non-conservative and rare, the occurrence of a non-conservative extrapolation
should be avoided whenever possible. Regarding the “mean( · )” columns, denoting the ex-
trapolation error, the linear functions performed far worse than the polynomial Equation (5)
with P3 = 4, especially with regard to the ΔKth,ASTM − ΔKth,cens,ASTM values, where the
mean and minimum extrapolation error were 0.20 MPa·m1/2 and 0.17 MPa·m1/2 , regarding
the fit over all data points and 0.23 MPa·m1/2 and 0.18 MPa·m1/2 for the fit over the first
n points, respectively, compared to 0.03 MPa·m1/2 and 0.01 MPa·m1/2 for the polynomial
with P3 = 4. Hence, in case extrapolation is inevitable, the polynomial with P3 = 4 brings
a notable improvement over the linear fits, suggested by the standards. With regard to
both standards, neither a statement on the minimum required crack propagation rate for a
valid evaluation of the FCG threshold nor any comments on the legitimacy of a potentially
necessary data extrapolation is given.

216
Materials 2022, 15, 4737

Table 7. Comparison of extrapolation errors obtained with various fit methods. The data refer
to K-decreasing tests at R = 0.8 and Kmax tests conducted on S690QL in lab air. The evaluation
comprised ten tests for ASTM threshold and thirteen specimens for ISO threshold. All values are
given in MPa·m1/2 .

ΔKth,ASTM − ΔKth,cens,ASTM ΔKth,ISO − ΔKth,cens,ISO


Method mean( · ) min( · ) mean( · ) min( · )
Linear, all data points (Equation (1)) 0.20 0.17 0.07 0.02
Linear, first n data points (Equation (1)) 0.23 0.18 0.07 0.02
Polynomial, neg. exp. (Equation (5)) 0.04 −0.02 0.03 −0.05
Polynomial, neg. exp. P3 = 4 (Equation (6)) 0.03 0.01 0.03 0.00
Polynomial, neg. exp. P3 = 5 (Equation (7)) 0.01 −0.02 0.02 −0.02

3.1.5. Application to the Full Dataset


Based on the conclusions drawn in Section 3.1.4, where a robust conservative extrap-
olation could be obtained for the linear fits as well as for the polynomial with negative
exponent fixed to P3 = 4, the full dataset, including all datasets that may be extrapolated,
has been reevaluated. Because an augmented fit interval may lead to non-conservative
results in case of linear fits, the intervals suggested in the standards have been used
(see Section 1.1). In contrast, a beneficial effect of an augmented interval has been observed
when using Equation (6), see Section 3.1.3. Hence, an augmented interval for calculat-
ing ΔKth,ASTM has been used in this case. Consequently, the fitting intervals have been
defined as da/dN = 5 × 10−8 mm/cycle ≤ da/dN ≤ 10−6 mm/cycle for ASTM and
da/dN = 10−8 mm/cycle ≤ da/dN ≤ 10−7 mm/cycle for ISO. The results are presented
in Table 8. The linear fit over the full interval resulted in the highest conservativeness in
combination with a small standard deviation. Using only the first n points of the interval
to generate the linear fit reduced the conservativeness with the drawback of increasing the
standard deviation, especially for ASTM. By using an appropriate nonlinear function like
Equation (6), the conservativeness as well as the standard deviation can be reduced.

Table 8. Comparison of threshold values obtained at R ≈ 0.8 with various fit methods using all
29 specimen data sets for ΔKth,ASTM and ΔKth,ISO ; results for S690QL in lab air.

Method ΔKth,ASTM in MPa·m1/2 ΔKth,ISO in MPa·m1/2


Linear, all data points (Equation (1)) 2.67 ± 0.06 2.22 ± 0.04
Linear, first n data points (Equation (1)) 2.73 ± 0.08 2.24 ± 0.05
Polynomial, neg. exp. P3 = 4 (Equation (6)) 2.76 ± 0.05 2.28 ± 0.04

In order to assess the methods capabilities when dealing with an augmented scatter
in ΔK, artificial scatter, as described in Section 3.1.2, is added to each specimen data set.
The results are given in Table 9. Like already observed in Figure 7, the two-parameter
polynomial Equation (6) is almost insensitive to scatter in test data, whereas both linear fits
partially suffer from pronounced susceptibility to scattered data.

Table 9. Comparison of threshold values determined from artificial scattered test data, obtained at
R ≈ 0.8 with various fit methods using all 29 specimen data sets for ΔKth,ASTM and ΔKth,ISO ; results
for S690QL in lab air.

Method ΔKth,ASTM in MPa·m1/2 ΔKth,ISO in MPa·m1/2


Linear, all data points (Equation (1)) 2.65 ± 0.06 2.16 ± 0.08
Linear, first n data points (Equation (1)) 2.68 ± 0.07 2.23 ± 0.05
Polynomial, neg. exp. P3 = 4 (Equation (6)) 2.76 ± 0.07 2.27 ± 0.04

217
Materials 2022, 15, 4737

3.2. Definition of the Fitting Function and Interval for the Determination of Thresholds Obtained at
R ≈ 0.8
Based on the results presented in Section 3.1, we suggested to use Equation (6) for fit-
ting the test data obtained at R ≈ 0.8. This polynomial exhibits only a minimal dependency
on scatter in test data and returns robust results that simultaneously show only a small,
but persistent conservativeness. Furthermore, data extrapolation has been shown to be
valid if the lowest available crack propagation rate fulfills da/dNmin ≤ 3 · da/dNth .
Since an augmented fit interval shows an additional reduction in standard devia-
tion, we suggested to use an augmented interval of 5 × 10−8 mm/cycle ≤ da/dN ≤
10−6 mm/cycle compared to the one proposed in the ASTM standard (10−7 mm/cycle ≤
da/dN ≤ 10−6 mm/cycle). Regarding the ISO operational definition, we recommended
to use the suggested interval of 10−8 mm/cycle ≤ da/dN ≤ 10−7 mm/cycle, since crack
propagation rates much lower than 10−8 mm/cycle are time consuming using conventional
FCG testing.
Using the herein proposed method for evaluating ΔKth , the artificial conservativeness
can be reduced and the fits’ robustness improved compared to the linear functions, as
shown in the previous paragraphs. Since the evaluation following the proposed method
involves a very low effort compared to the conducted experiments, the application thereof
is generally preferable over the linear fits.

3.3. Evaluation of the Fatigue Crack Propagation Threshold at R = −1


In contrast to tests carried out at R = 0.8, specimens tested at lower load ratios like
R = −1 may exhibit a distinct influence of extrinsic effects such as crack closure. This can
be observed in the example reported in Figure 10: due to the progressive development of
crack closure during the load shedding test, the crack propagation rate decreases rapidly
in the near-threshold regime, leading to a steep crack propagation curve towards the
threshold. This has major consequences on the evaluation of the fatigue crack propagation
threshold due to the fact that much fewer experimental points are available in the selected
fitting intervals.
The linear fit over all points within the interval (Figure 10a) is not capable of handling
the pronounced curvature. Hence, ΔKth,ASTM is calculated overly conservatively. This holds
true also for the polynomial with P3 = 4, Equation (6) (Figure 10c). In contrast, the linear fit
using the first n points (Figure 10b) shows a fairly good threshold approximation. The best
results are obtained using the three-parameter polynomial Equation (5) (Figure 10d), which
adapts to the curvature displayed by the data fairly well and results in the least (but still)
conservative ΔKth,ASTM value.

10−6 10−6
da/dN in mm/cycle

da/dN in mm/cycle

Experimental dataset Experimental dataset


Fitting dataset Fitting dataset
10−7 Fitting curve 10−7 Fitting curve
(Eq. (1), all data points) (Eq. (1), first n data points)
ΔKth, ASTM = 11.45 MPa · m1/2 ΔKth, ASTM = 11.87 MPa · m1/2

12 13 14 12 13 14
ΔK in MPa · m1/2 ΔK in MPa · m1/2

(a) (b)
Figure 10. Cont.

218
Materials 2022, 15, 4737

10−6 10−6

da/dN in mm/cycle

da/dN in mm/cycle
Experimental dataset Experimental dataset
Fitting dataset Fitting dataset
10−7 10−7
Fitting curve (Eq. (6)) Fitting curve (Eq. (5))
ΔKth, ASTM = 11.77 MPa · m1/2 ΔKth, ASTM = 11.89 MPa · m1/2

12 13 14 12 13 14
ΔK in MPa · m1/2 ΔK in MPa · m1/2

(c) (d)
Figure 10. FCG data showing a distinct effect of fatigue crack closure leading to crack arrest. The tests
have been conducted following a compression precracking load reduction procedure at R = −1
in lab air. Different fitting strategies have been used: (a) the linear fit using all data in the interval
10−7 mm/cycle ≤ da/dN ≤ 10−6 mm/cycle provided an overly conservative ΔKth,ASTM value;
(b) the linear fit using the first n = 5 data points (out of 29 within the interval) showed a fairly
good ΔKth,ASTM approximation; (c) Equation (6), using a fixed exponent P3 = 4, was not capable of
reproducing the curvature and therefore gave a very conservative result; (d) a very good result could
be achieved using the three-parameter variant of the polynomial (P3 ≈ 24.1).

3.3.1. Handling of Data Affected by Extrinsic Mechanisms


Extrinsic mechanisms affect the cyclic deformation in the crack wake and therefore
influence the crack growth rate [22]. The results about tests carried out at R = 0.1 and
R = −1 showed a kink of the crack propagation curve in the near-threshold regime. For an
in detail discussion on these findings, see [21]. The results depicted in Figure 11 about
two specimens tested in lab air, using a conventional K-decreasing procedure at R = −1,
pose the question of how to analyze the data to evaluate the fatigue crack propagation
threshold. In fact, both specimens show a distinct kink in the FCG data at about da/dN ≈
10−7 mm/cycle. In such cases, the calculation of a threshold stress intensity factor range
is—regardless of the standard used—questionable, since no asymptotic behavior of the
da/dN-ΔK curve towards ΔKth is observable (see Section 1).

10−6
da/dN in mm/cycle

da/dN in mm/cycle

10−6

10−7
10−7

10 11 12 13 14 15 16 17 18 11 12 13 14 15 16 17 18
ΔK in MPa · m1/2 ΔK in MPa · m1/2

(a) (b)
Figure 11. FCG data showing a kink at about the ASTM threshold (da/dNth,ASTM = 10−7 mm/cycle):
(a) kinking starts below da/dN = 10−7 mm/cycle; (b) kinking starts slightly above 10−7 mm/cycle.
The data refer to conventional K-decreasing at R = −1 in lab air. Both tests have been interrupted,
since neither crack arrest has been observed nor da/dNth,ISO has been reached.

No general rule is applicable to the evaluation of the fatigue crack propagation thresholds,
but each dataset shall be analysed separately. For instance, Figure 11a shows that the determi-
nation of ΔKth,ASTM would have been possible, since the kink started below da/dNth,ASTM =

219
Materials 2022, 15, 4737

10−7 mm/cycle. If the aim of this K-decreasing test would have been to determine ΔKth,ASTM,
one might have stopped the test after reaching da/dN < da/dNth,ASTM = 10−7 mm/cycle
for the first time. Hence, one would have never observed the effect of corrosion on FCG data
starting just below ΔK ≈ 12 MPa·m1/2. Nevertheless, as the crack propagation data did not
display a pronounced threshold behavior, a value for ΔKth,ASTM should not be provided. This
holds true also for the dataset depicted in Figure 11b, in which corrosion effects started above
da/dNth,ASTM = 10−7 mm/cycle. In these cases, we recommend providing the last da/dN–ΔK
reading recorded within the test as a pure indication of the lowest stress intensity factor range
obtained in the load reduction test, which, nevertheless, must not be taken as a fatigue crack
propagation threshold. Hence, an automated evaluation of FCG data should only be performed
after checking the crack propagation data in the near-threshold regime.

3.3.2. Influence of a Lower Data Density


The majority of test results presented herein show very low noise in conjunction
with a fairly high data density. Both properties depend on the quality of raw data and
the methodology used to calculate da/dN and ΔK. In Figure 12, the test data presented
in Figure 10 are reduced by a factor of two by skipping every second data point. Since
still 14 data points distributed over the whole interval of 10−6 mm/cycle ≤ da/dN ≤
10−7 mm/cycle are left, this data set clearly fulfills the requirement of providing at least
five points, defined in [4].

10−6 10−6
da/dN in mm/cycle

da/dN in mm/cycle

Artificially censored
experimental dataset Artificially censored
Fitting dataset experimental dataset
Fitting dataset
10−7 Fitting curve 10−7
(Eq. (1), first n data points) Fitting curve (Eq. (5))
ΔKth, ASTM = 11.77 MPa · m1/2 ΔKth, ASTM = 11.87 MPa · m1/2

12 13 14 12 13 14
ΔK in MPa · m1/2 ΔK in MPa · m1/2

(a) (b)
Figure 12. Test data presented in Figure 10 reduced by 50% to assess the influence of a lower data
density next to da/dNth,ASTM . (a) the linear fit using the best first n data points gives an overly
conservative ΔKth,ASTM value; (b) using the three-parameter polynomial Equation (5) shows a very
good agreement with test data.

Here, the advantages of an appropriate nonlinear fit functions apply. The linear


fit using the first n points shows a pronounced underestimation of the threshold stress
intensity range, whereas the three-parameter polynomial’s sensitivity to the number of
data points is very limited.

3.3.3. Validation of the Proposed Method


After examining all the possible issues which might influence the robust determination
of the fatigue crack propagation threshold, the methods have been validated against
various datasets.

Conventional K-Decreasing at R = −1 in Lab Air


The first datasets investigated stemmed from a total of eight SENB specimens made of
S690QL, whereof three showed a kink (see Figure 11) and therefore have not been consid-
ered in the analysis. These tests have been conducted using the K-decreasing procedure
included in the standards [4,6] at constant load ratio R = −1 in lab air. All five remaining
datasets contained data points below da/dN = 1.1 × 10−7 mm/cycle and therefore have

220
Materials 2022, 15, 4737

been considered valid for ASTM operational threshold evaluation, while two of them were
also valid regarding the ISO definition. The evaluated fatigue crack propagation thresholds,
including the corresponding standard deviations, are presented in Table 10. Note that,
in case of ISO, no standard deviation has been calculated due to insufficient data available.

Table 10. Comparison of threshold values obtained with various fit methods in case of K-decreasing
tests conducted on S690QL at R = −1 in lab air.

Method ΔKth,ASTM in MPa·m1/2 ΔKth,ISO in MPa·m1/2


Linear, all data points (Equation (1)) 13.43 ± 0.91 12.78
Linear, first n data points (Equation (1)) 13.67 ± 0.80 12.89
Polynomial, neg. exp. (Equation (5)) 13.69 ± 0.82 12.92
Polynomial, neg. exp. P3 = 4 (Equation (6)) 13.56 ± 0.85 12.90
Polynomial, neg. exp. P3 = 5 (Equation (7)) 13.57 ± 0.85 12.90

The conservativeness of each method has been analyzed: the two-parameter polyno-
mials Equation (6) and Equation (7), and especially the linear fit over all points, showed
a pronounced underestimation of ΔKth,ASTM in conjunction with a higher standard devia-
tion compared to the linear fit over the first n points or the three-parameter polynomial
Equation (5). The observations with regard to artificial conservativeness also apply to the
results for ΔKth,ISO . Comparing the latter two methods, both show comparable results,
both regarding the mean ΔKth value and the standard deviation.

Compression Precracking Load Reduction at R = −1 in Lab Air


The datasets investigated stemmed from a total of ten SENB specimens made of
S690QL. All specimens showed a distinct threshold behavior and therefore all have been
considered in the analysis. These tests have been conducted using compression precracking
followed by a load reduction procedure at constant R = −1 in lab air. All ten datasets
contained data points below da/dN = 1.1 × 10−7 mm/cycle and therefore have been con-
sidered valid for ASTM operational threshold evaluation, whereas none of them contained
points in order to calculate a threshold value according to ISO. The threshold stress intensity
factor ranges with corresponding standard deviations are presented in Table 11.

Table 11. Comparison of threshold values obtained with various fit methods in case of compression
precracking load reduction tests carried out at R = −1 in lab air.

Method ΔKth,ASTM in MPa·m1/2


Linear, all data points (Equation (1)) 9.88 ± 0.99
Linear, first n data points (Equation (1)) 10.52 ± 0.89
Polynomial, neg. exp. (Equation (5)) 10.54 ± 0.91
Polynomial, neg. exp. P3 = 4 (Equation (6)) 10.20 ± 0.94
Polynomial, neg. exp. P3 = 5 (Equation (7)) 10.22 ± 0.93

The same conclusions as in the previous paragraph can be drawn: the linear fit over
the first n points and the three-parameter polynomial Equation (5) provided the best results.

3.4. Definition of the Fitting Function and Interval for Tests Conducted at R 0.8
Data obtained at load ratios momentously lower than R = 0.8 might be affected by
phenomena like crack-closure or corrosion, which make the determination of the fatigue
crack propagation threshold difficult. When a steep gradient or a kink in the near-threshold
data are observed, no general or automatic extrapolation of these test datasets without
further investigation is advisable. Furthermore, the augmentation of fit intervals to crack
propagation rates smaller than da/dNth may lead to non-conservative results and therefore
it is not recommended.
If a valid threshold behavior is observed, the evaluation using the three-parameter polyno-
mial Equation (5) provided constantly conservative, but not overly conservative, results that

221
Materials 2022, 15, 4737

are neither sensitive to the data density nor to scatter. Therefore, we recommend this function
over the linear fit over the first n points, which indeed performed well on regular shaped
datasets. The fitting intervals shall follow the recommendations provided in the standards
(10−7 mm/cycle ≤ da/dN ≤ 10−6 mm/cycle for ASTM and 10−8 mm/cycle ≤ da/dN ≤
10−7 mm/cycle for ISO). In case no definite threshold is reached, i.e., no crack arrest is observed
(due for instance to anti-shielding effects, see Figure 11a), ΔKth cannot be determined. Therefore,
we recommended providing the last da/dN- ΔK data pair for pure orientation, which shall not
be intended as substitute for ΔKth .

3.5. Application of the Fitting Methods to the IBESS Dataset


The proposed method has been validated further against the data from the IBESS
project [23,24].Within the IBESS project, fatigue crack propagation tests have been per-
formed on two different structural steels, S355NL and S960QL, using both standard K-
decreasing procedures and CPLR tests at constant load ratios varying between R = −1 and
R = 0.7. The number of specimens tested at each stress ratio was limited; therefore, the
present validation considered just those tests for which a meaningful data analysis could
be performed. In particular, three tests at R = 0 for the S355NL and three tests at R = 0.5
for the S960QL have been considered. Furthermore, according to the recommendations on
the fitting intervals given in this work, the data have been further narrowed. In case of
the S355NL, all three datasets have been considered valid with respect to the ASTM fitting
interval, whereas just two among them could be used for the determination of the fatigue
crack propagation threshold according to ISO. For the S960QL, only two tests for ASTM and
one for ISO have been included in the analysis. The results displayed in Tables 12 and 13
confirm the conclusions drawn for the S690QL: the linear fit using the first n points and
the polynomial with three parameters reduce the conservativeness in the evaluation of
the fatigue crack propagation thresholds. The method proposed in the standards always
provides the most conservative results. Furthermore, for the datasets with enough valid
data in the fitting interval (ΔKth,ASTM ), the polynomial with three parameters provided the
smallest standard deviation.

Table 12. Comparison of the threshold values for the S355NL tested at R = 0 obtained with various
fitting methods—a total of three sets are eligible for ΔKth,ASTM evaluation and two for ΔKth,ISO .

Method ΔKth,ASTM in MPa·m1/2 ΔKth,ISO in MPa·m1/2


Linear, all data points (Equation (1)) 5.96 ± 0.40 5.75
Linear, first n data points (Equation (1)) 6.02 ± 0.38 5.86
Polynomial, neg. exp. (Equation (5)) 6.08 ± 0.33 5.88

Table 13. Comparison of the threshold values for the S960QL tested at R = 0.5 obtained with various
fitting methods—a total of two sets eligible for ΔKth,ASTM evaluation and one for ΔKth,ISO .

Method ΔKth,ASTM in MPa·m1/2 ΔKth,ISO in MPa·m1/2


Linear, all data points (Equation (1)) 3.30 3.04
Linear, first n data points (Equation (1)) 3.31 3.10
Polynomial, neg. exp. (Equation (5)) 3.31 3.08

4. Conclusions
The present paper compared several methods for the evaluation of the fatigue crack
propagation thresholds. New fitting strategies have been introduced and calibrated on a
large dataset of crack growth data for the S690QL. The goodness of the fitting methods has
been validated further against a dataset for S355NL and S960QL.
The following conclusions can be drawn:
• The ASTM E647 and ISO 12108 standards suggest to fit log ΔK over log da/dN data
using a linear fit, but leave plenty of room for interpretation with respect to the choice
of the points in the fitting interval;

222
Materials 2022, 15, 4737

• When using all data points within the suggested fitting intervals, the most conservative
values of ΔKth are obtained. However, the fit is not very subjected to scattered data;
• To use only the first n data points starting from the threshold crack propagation rate
in order to ensure the best linear fit reduces the conservativeness at the cost of a more
pronounced susceptibility to scatter and lower density of the data;
• The proposed fitting polynomials provided an improvement with respect to the
goodness of the fit and susceptibility to scatter;
• An extrapolation of data was possible within given bounds for the structural steel
S690QL, tested in lab air at room temperature at R ≈ 0.8. Further tests comprising
changes in materials, temperatures and the test environment should be conducted to
assess the validity ranges;
• Tests subjected to crack closure phenomenon cannot be assessed in a fully automatic
manner and require a manual dataset evaluation.

Author Contributions: Conceptualization, J.A.S., L.D., M.B.G., M.M., U.Z. and M.K.; methodology,
J.A.S., L.D., M.B.G., M.M., U.Z. and M.K.; software, J.A.S. and L.D.; validation, J.A.S. and L.D.; formal
analysis, J.A.S., L.D., M.B.G. and M.M.; investigation, J.A.S., L.D., M.B.G., M.M., U.Z. and M.K.;
resources, M.O.; data curation, J.A.S. and L.D.; writing—original draft preparation, J.A.S.; writing—
review and editing, L.D., M.B.G., M.M., U.Z., M.K. and M.O.; visualization, J.A.S.; supervision, M.M.,
U.Z., M.K. and M.O.; project administration, M.M., U.Z., M.K. and M.O.; funding acquisition, U.Z.
and M.O. All authors have read and agreed to the published version of the manuscript.
Funding: This work is part of the research project IGF 20530 N / 1263 “Ermittlung des intrinsischen
Schwellenwerts und dessen Validierung als Werkstoffparameter” from the Research Association for
Steel Application (FOSTA), Düsseldorf, which is supported by the Federal Ministry of Economic
Affairs and Climate Action through the German Federation of Industrial Research Associations (AiF)
as part of the program for promoting industrial cooperative research (IGF) on the basis of a decision
by the German Bundestag. The project is carried out at BAM Berlin and MPA-IfW Darmstadt.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Not applicable.
Acknowledgments: We acknowledge support by the Deutsche Forschungsgemeinschaft (DFG—German
Research Foundation) and the Open Access Publishing Fund of the Technical University of Darmstadt.
Conflicts of Interest: The authors declare no conflict of interest.

Symbols and Abbreviations


The following symbols and abbreviations are used in this manuscript:

ΔF applied force range


ΔK applied stress intensity factor range
ΔKLR K-decreasing FCG test procedure at constant load ratio
ΔKth fatigue crack propagation threshold
ΔKth,ASTM ΔKth referring to the ASTM operational definition
ΔKth,ISO ΔKth referring to the ISO operational definition
ΔKth,cens ΔKth obtained for the censoring of the data set
μ mean value of the distribution
σu ultimate tensile strength
σy upper yield strength
a crack size

223
Materials 2022, 15, 4737

A elongation at break
da/dN fatigue crack propagation rate
da/dNth da/dN referring to an operational threshold definition
da/dNth,ASTM ASTM operational threshold definition of da/dNth
da/dNth,ISO ISO operational threshold definition of da/dNth
e Weibull threshold parameter
E Young’s modulus
I degree of the polynomial
k Weibull shape parameter
Kb Weibull instability parameter
Kmax maximum stress intensity factor in a loading cycle
KV2 impact energy
n number of data points
N number of loading cycles
Pi fitting parameters, i ∈ N
R stress ratio
Rmax maximum stress ratio within Kmax -FCG test
v Weibull characteristic value
ASTM American Society for Testing and Materials
BAM Bundesanstalt für Materialforschung und -prüfung
CPLR compression precracking load reduction
FCG fatigue crack growth
ISO International Organization for Standardization
MPA-IfW Materialprüfungsanstalt Darmstadt, Institut für Werkstoffkunde
ODR orthogonal distance regression
SD standard deviation
SENB single edge notch bending

References
1. Paris, P.C.; Sih, G.C. Stress Analysis of Cracks; Technical Report; NASA: Washington, DC, USA, 1964.
2. Knott, J.F. Fundamentals of Fracture Mechanics; Butterworth: London, UK, 1973.
3. Tada, H.; Paris, P.C.; Irwin, G.R. The Stress Analysis of Cracks Handbook, 3rd ed.; ASME Press: New York, NY, USA, 2000.
4. ASTM E647-15e1; Standard Test Method for Measurement of Fatigue Crack Growth Rates. ASTM International: West Con-
shohocken, PA, USA, 2015. [CrossRef]
5. Ritchie, R. Mechanisms of Fatigue-Crack Propagation in Ductile and Brittle Solids. Int. J. Fract. 1999, 100, 55–83. [CrossRef]
6. ISO 12108:2018; Metallic Materials—Fatigue Testing—Fatigue Crack Growth Method. ISO: Geneva, Switzerland, 2018.
7. Bucci, R.J. Development of a proposed ASTM standard test method for near-threshold fatigue crack growth rate measurement. In
Fatigue Crack Growth Measurement and Data Analysis. ASTM STP 738; Hudak, S.J., Bucci, R.J., Eds.; ASTM International: West
Conshohocken, PA, USA, 1981; pp. 5–28. [CrossRef]
8. Salivar, G.; Hoeppner, D. A Weibull Analysis of Fatigue-Crack Propagation Data from a Nuclear Pressure Vessel Steel. Eng. Fract.
Mech. 1979, 12, 181–184. [CrossRef]
9. Smith, F.; Hoeppner, D.W. Use of the Four Parameter Weibull Function for Fitting Fatigue and Compliance Calibration Data. Eng.
Fract. Mech. 1990, 36, 173–178. [CrossRef]
10. Boggs, P.T.; Rogers, J.E. Orthogonal distance regression. In Statistical Analysis of Measurement Error Models and Applications;
American Mathematical Society: Providence, RI, USA, 1990; Volume 112, pp. 183–194. [CrossRef]
11. Döker, H. Fatigue Crack Growth Threshold: Implications, Determination and Data Evaluation. Int. J. Fatigue 1997, 19, 145–149.
[CrossRef]
12. ASTM E 647. Annual Book of ASTM Standards; Section 4; American Society of Testing and Materials: Philadelphia, PA, USA, 1995;
Volume 03.01, pp. 578–614.
13. Klesnil, M.; Lukáš, P. Influence of Strength and Stress History on Growth and Stabilisation of Fatigue Cracks. Eng. Fract. Mech.
1972, 4, 77–92. [CrossRef]
14. Romvari, P.; Tot, L.; Nad’, D. Analysis of Irregularities in the Distribution of Fatigue Cracks in Metals. Strength Mater. 1980,
12, 1481–1492. [CrossRef]
15. Zheng, X. A Simple Formula for Fatigue Crack Propagation and a New Method for the Determination of ΔKth. Eng. Fract. Mech.
1987, 27, 465–475. [CrossRef]
16. Ramsamooj, D.; Shugar, T. Model Prediction of Fatigue Crack Propagation in Metal Alloys in Laboratory Air. Int. J. Fatigue 2001,
23, 287–300. [CrossRef]
17. Day, B.; Goswami, T. Weibull Model Development for Fatigue Crack Growth. J. Mech. Behav. Mater. 2002, 13, 283–296. [CrossRef]

224
Materials 2022, 15, 4737

18. Paolino, D.S.; Cavatorta, M.P. Sigmoidal Crack Growth Rate Curve: Statistical Modelling and Applications: SIGMOIDAL
CRACK-GROWTH-RATE CURVE. Fatigue Fract. Eng. Mater. Struct. 2013, 36, 316–326. [CrossRef]
19. Maierhofer, J.; Pippan, R.; Gänser, H.P. Modified NASGRO Equation for Physically Short Cracks. Int. J. Fatigue 2014, 59, 200–207.
[CrossRef]
20. E1823-20; Standard Terminology Relating to Fatigue and Fracture Testing. ASTM International: West Conshohocken, PA,
USA, 2020. [CrossRef]
21. Duarte, L.; Schönherr, J.A.; Madia, M.; Zerbst, U.; Geilen, M.B.; Klein, M.; Oechsner, M. Recent Developments in Fatigue Crack
Propagation Threshold Determination. Int. J. Fatigue 2022. submitted.
22. Pippan, R.; Hohenwarter, A. Fatigue crack closure: A review of the physical phenomena. Fatigue Fract. Eng. Mater. Struct. 2017,
40, 471–495. [CrossRef] [PubMed]
23. Zerbst, U. Analytische bruchmechanische Ermittlung der Schwingfestigkeit von Schweißverbindungen (IBESS-A3); Technical
Report; Bundesanstalt für Materialforschung und -prüfung (BAM): Berlin, Germany, 2016.
24. Kucharczyk, P.; Madia, M.; Zerbst, U.; Schork, B.; Gerwin, P.; Münstermann, S. Fracture-mechanics based prediction of the fatigue
strength of weldments. Material aspects. Eng. Fract. Mech. 2018, 198, 79–102. [CrossRef]

225
materials
Article
Influence of Creep Damage on the Fatigue Life of P91 Steel
Stanisław Mroziński 1 , Zbigniew Lis 1 and Halina Egner 2, *

1 Faculty of Mechanical Engineering, Bydgosz University of Science and Technology, Al. Prof. S. Kaliskiego 7,
85-796 Bydgoszcz, Poland; [email protected] (S.M.); [email protected] (Z.L.)
2 Faculty of Mechanical Engineering, Cracow University of Technology, Al. Jana Pawła II 37,
31-864 Kraków, Poland
* Correspondence: [email protected]

Abstract: The following paper presents the results of tests on samples made of P91 steel under
the conditions of simultaneously occurring fatigue and creep at a temperature of 600 ◦ C. The load
program consisted of symmetrical fatigue cycles with tensile dwell times to introduce creep. Static
load (creep) was carried out by stopping the alternating load at the maximum value of the alternating
stress. The tests were carried out for two load dwell times, 5 s and 30 s. A comparative analysis of
the test results of fatigue load with a dwell time on each cycle confirmed that creep accompanying
the variable load causes a significant reduction in sample durability. It was shown in the paper that
regarding the creep influence in the linear fatigue damage summation approach, it is possible to
improve the compliance of the fatigue life predictions with the experimental results.

Keywords: low-cycle fatigue; creep; damage; strain energy

1. Introduction
Citation: Mroziński, S.; Lis, Z.; Egner,
Analysis of the operational loads of many structural elements indicates that they are
H. Influence of Creep Damage on the often subject to mechanical loads where the independent quantity is force (e.g., pressure)
Fatigue Life of P91 Steel. Materials and the dependent quantity is the deformation of the element. This applies, inter alia,
2022, 15, 4917. https://doi.org/ to facilities operating at elevated temperatures. This type of load may be additionally
10.3390/ma15144917 accompanied by material creep, which changes the nature of the load and the durability
of the designed elements. Creep is much more pronounced at high temperatures, for
Academic Editors: Jaroslaw
example, in pipelines that contain a pressurized hot medium or gas turbine components
Galkiewicz, Lucjan Śnieżek and
that are statically loaded but operate at high and variable temperatures. In such cases, when
Sebastian Lipiec
predicting the fatigue life of a technical object, only the fatigue characteristics are taken into
Received: 9 June 2022 account, obtained for example, under the conditions of controlled deformation (ε ac = const)
Accepted: 12 July 2022 or stresses (σa = const), while in real applications, the conditions are different, which may
Published: 14 July 2022 lead to divergent results obtained from calculations and tests. In papers [1–4], the results of
Publisher’s Note: MDPI stays neutral low-cycle fatigue tests are presented for samples made from the same material, obtained in
with regard to jurisdictional claims in the conditions of σa = const and ε ac = const. Based on the comparative analysis of the fatigue
published maps and institutional affil- diagrams in the 2N f − ε ac coordinate system, it was found that at the same deformation
iations. levels, the fatigue life under the conditions σa = const is lower than that obtained in the
conditions of controlled deformation (ε ac = const). The deformation asymmetry during
the half-cycles of tension and compression, which caused the cyclic creep of the material,
was given as the reason for the reduction in durability under the conditions σa = const.
Copyright: © 2022 by the authors. Creep-fatigue experiments were conducted in [5] under stress-controlled conditions to
Licensee MDPI, Basel, Switzerland. understand dwell time’s effects on the durability of a DS superalloy. It was shown that
This article is an open access article the most important damage source is creep-fatigue interaction damage which is controlled
distributed under the terms and
by load amplitude, stress ratio, and dwell time per cycle. In paper [6], multiaxial creep-
conditions of the Creative Commons
fatigue life and its variations with the increasing holding time under different loading
Attribution (CC BY) license (https://
programs were investigated for 304 stainless steel. The introductions of holding periods
creativecommons.org/licenses/by/
under creep-fatigue loadings resulted in lifetime reductions as compared with pure fatigue
4.0/).

Materials 2022, 15, 4917. https://doi.org/10.3390/ma15144917 227 https://www.mdpi.com/journal/materials


Materials 2022, 15, 4917

loadings. The problem of creep-fatigue interaction for various engineering materials was
also recently investigated in [7–9]. In papers [10,11], tests were carried out to determine
the effect of the sequence of fatigue cycles and static loads in the load program. It was
found that the sequence of both loads in the load program affects the durability. Variable
loads preceding static loads result in obtaining higher durability than the variant in which
the static load precedes the variable loads. In light of the obtained results, it was found
that creep damage and fatigue damage are not independent. These results confirm the
observations included, inter alia, in [12], where the authors found that short cracks appear
during the variable load, and their density is larger if the material is initially subjected to
a static load. In paper [10], experimental verification of the linear hypothesis of damage
summation [13] was also carried out. Since the linear summation hypothesis is insensitive
to the sequence of events in the load program, it may provide erroneous results of the
durability assessments in comparison to the experimental data.
The problem of fatigue life predictions under the conditions of simultaneous occur-
rence of the variable and the static load was analyzed with the use of various fatigue
descriptions, i.e., deformation, stress, or energy approach. In the works [11,14–20], the
problem of material fatigue was attempted to be explained with the use of energy descrip-
tion. The results indicate significant advantages of the energy approach, even if a relatively
simple combination of the static and variable load effects was applied. The main advantage
of the energy description is the ability to accumulate damage using plastic strain energy as
a criterion parameter.
In papers [21–23], the authors investigated the influence of creep on fatigue life and
verified currently used calculation models. Several modifications of the classic Palmgren-
Miner’s linear fatigue damage summation hypothesis [13] were proposed. The authors
concluded that disregarding creep damage in the calculations may lead to a significant
differentiation in comparison with the test results. The creep duration influences the results
diversion from the experiment.
The present work is a continuation of research on the improvement of constitutive
modeling of low cycle fatigue (cf. [1,4,11]), as well as on the explanation of the phenomena
accompanying the load program containing constant and variable loads.

2. Materials and Methods


Experimental Testing
Samples for fatigue tests were prepared from P91 steel (Rm = 716 MPa, Re = 564 MPa,
A5 = 35% at room temperature). The chemical composition of P91 steel is shown in Table 1.

Table 1. Chemical composition of P91 steel.

C Si Mn P S Cr Mo Ni
0.197 0.442 0.489 0.017 0.005 8.82 0.971 0.307
Al Co Cu Nb Ti V W
0.012 0.017 0.036 0.074 0.004 0.201 0.02

The samples were cut out from a thick-walled pipe for power industry applications,
with an external diameter of 200 mm and a wall thickness of 20 mm. The samples were
shaped under the guidelines specified in the standard [24]. The dimensions of the sample
are shown in Figure 1.

228
Materials 2022, 15, 4917

φ K


φ K
0 
φ 
5 5


 
 


(a) (b)
Figure 1. (a) Test sample dimensions (in [mm]); (b) test sample view.

The load program of experimental tests included: static tensile tests, creep tests, low-
cycle fatigue tests, and tests in which the samples were subjected to fatigue with a dwell
time in each cycle. The load programs and their symbols are shown in Figure 2.

(a) (b)

(c)
ߪଵ ൌ ߪ௔ଵ = 219 MPa, ߪଶ ൌ ߪ௔ଶ = 225 MPa, ߪଷ ൌ ߪ௔ଷ = 249 MPa, ߪସ ൌ ߪ௔ସ = 258 MPa
߬ଵ = 0 s, ߬ଶ = 5 s, ߬ଷ = 30 s
(d)

Figure 2. Load programs: (a) static load, (b) fatigue load, (c) static load + fatigue load, (d) load parameters.

Experimental tests (static and fatigue tests) were carried out at a temperature of
T = 600 ◦ C on an Instron 8502 testing machine, equipped with a heating chamber with a
maximum temperature range of 1000 ◦ C (Figure 3). The temperature of the sample was
monitored with a thermocouple attached to the sample. The sample deformation was
measured with an extensometer with a measurement base of 12.5 mm. Fatigue tests in the
conditions σa = const were carried out on four stress levels (see Figure 2d), determined
based on the analysis of static tension diagrams (Figure 4b). The load frequency was
0.2 Hz. During the tests, the instantaneous values of the force loading the sample and its
deformation were recorded. Two fatigue tests were performed at each load level.

229
Materials 2022, 15, 4917


 
 
 


(a) (b)
Figure 3. Test stand: (a) heating chamber; (b) sample in grips. 1—sample, 2—thermocouple, 3—upper
grip, 4—extensometer, 5—lower grip, 6—heating chamber, 7—machine frame.

(a) (b)
Figure 4. Monotonic stress–strain curves: (a) for different test temperatures, (b) stress amplitude
levels chosen for creep-fatigue tests (T = 600 ◦ C).

3. Results
3.1. Static Tension Test
Figure 4a present a stress–strain curve obtained at temperature T = 600 ◦ C. To illustrate
the influence of temperature on strength properties, the figure additionally includes the
tensile curves of P91 steel obtained at two other temperatures, 20 ◦ C and 400 ◦ C. The stress
in the tested sample was calculated as the ratio of the instantaneous force loading the
sample and the initial cross-sectional area of the sample (nominal stress). Figure 4b include
a fragment of the tensile diagram obtained at the temperature T = 600 ◦ C, with the stress
amplitude levels for the creep-fatigue tests indicated in it, see Figure 2d.
The tensile diagrams (Figure 4) were subjected to a detailed analysis aimed at deter-
mining the basic strength parameters (see Table 2).

Table 2. Strength parameters of P91 steel at different temperatures.

T, ◦ C Rp0.2 , MPa Rm , MPa Z, % A12.5 , % E, MPa


20 564 716 62 35 209,850
400 463 571 69 44 183,220
600 317 353 72 56 132,890

The comparative analysis of the data in Table 2 confirm the common literature reports
on the influence of temperature on the basic strength parameters of P91 steel [25–28]. Based
on the analysis of the tensile diagrams (Figure 4) and the values of the strength parameters
listed in Table 2, it can be concluded that the increase in temperature leads to a reduction
in the yield point (R p0.2 ). A similar relationship is observed in the case of the tensile
strength (Rm ) and Young’s modulus (E), for which a decrease in this parameter is also

230
Materials 2022, 15, 4917

visible. The decrease in strength properties with temperature is accompanied by an increase


in elongation (A12.5 ) and constriction (Z).

3.2. Fatigue Tests


The analysis of the fatigue test results was carried out using the hysteresis loop param-
eters (σa , ε ap , ε ac ), which are necessary for the analytical description of the cyclic properties
of steel following the standard [24]. Changes in the hysteresis loop parameters in function
of the number of a load cycle were observed during all the fatigue tests. As expected,
cyclic creep of the sample material was observed under the conditions of controlled stress
(σa = const). The phenomenon consisted of the shift of the hysteresis loop along the strain
axis, increasing the maximum strain on a cycle ε max . Figure 5 show exemplary hysteresis
loops at two stress levels (σa = 225 MPa, σa = 249 MPa). The presented loops refer to a load
program in which creep time τ = 0 (see Figure 2).

εmax εmax
 
σ N N N N σ
N N N N N
03D 03D


 

 
í      í   
ε ε
í í

í í
N  Δεap
Δεap
í í
(a) (b)

Figure 5. Hysteresis loops recorded in fatigue tests (τ = 0, T = 600 ◦ C): (a) σa = 225 MPa, (b) σa = 249 MPa.

Based on the hysteresis loops analysis, it can be concluded that the amount of max-
imum strain on a cycle (ε max ) during the fatigue test is influenced by the stress level σa .
As expected, this influence is the smallest at the lowest stress levels and increases with
increasing stress. It can also be seen that the strain increments are characterized by different
rates, low at the beginning of the test and increasing with the number of a load cycle
(Figure 6a).


ε max  4
1 − σa1 03D Δεap
 4 í1
2 − σa2 03D mm·mm 3
3 − σa3 03D  2

δΔε4

4 − σa4 03D
3 1


δΔε1

 2 1

 
        
1XPEHURIF\FOHn n/N
(a) (b)
Figure 6. Evolution of: (a) maximum strain on cycle; (b) plastic strain on cycle Δε ap .

231
Materials 2022, 15, 4917

Based on the analysis of the loops shown in Figure 5, it can be concluded that with the
increase in the number of the load cycle, the range of plastic deformation Δε ap also increases
at a growing rate. The amount of plastic deformation in a cycle, Δε ap , is also influenced by
the stress level σa . To illustrate this influence, Figure 6b summarize the changes of Δε ap in
function of a load cycle number n related to the fatigue life N.
The observed increase in the range of plastic deformation proves a clear cyclic softening
of the P91 steel at the temperature of 600 ◦ C. To quantify the softening, a coefficient δΔε was
introduced (see Figure 6b), defined by the following relationship:

δΔε = Δε ap( N ) − Δε ap(1) (1)

where Δε ap(1) is the range of plastic deformation in the first cycle, while Δε ap( N ) is the range
in the last cycle. It can be seen in Figure 6b that with the increase in stress amplitude, the
value of the coefficient δΔε also increases (δΔε4 > δΔε3 > δΔε2 > δΔε1 ).

3.3. Creep Tests


During the creep tests, similarly to the fatigue tests, the instantaneous values of the
sample elongation as a function of the creep test time were recorded. As expected, the
results of the creep tests are influenced by the constant stress level σ. Figure 7 present
the sample elongation vs. time recorded during the creep tests at four stress levels (cf.
also [11]).

 
1 − σ1 03D
(ORQJDWLRQ

 2 − σ2 03D
4 3 2 1
 3 − σ3 03D
4 − σ4 03D


 Ι
ΙΙ ΙΙΙ

   
WLPHPLQ

Figure 7. Sample elongation in creep test.

All creep-to-fracture curves exhibit three stages with different elongation rates of the
specimen over time. These stages were indicated on the creep curve obtained for the lowest
stress (σ1 = 219 MPa), stage I with a variable elongation rate, stage II with a constant rate,
and stage III also with a variable creep rate. The stress level σ affects both the length of
these stages as well as the creep rate in the individual stages.

3.4. Fatigue-Creep Tests


As expected, during the programs containing both variable and static loads (τ > 0,
see Figure 2c), cyclic creep of the material was observed. It manifested in the hori-
zontal shift of the hysteresis loops and the increase in the maximum strain on a cycle
ε max . Figures 8 and 9 present chosen hysteresis loop positions at two stress levels,
σa = 219 MPa and σa = 258 MPa.

232
Materials 2022, 15, 4917

εmax εmax 
σ  N N N N N  N N N N N
σ
03D  
03D
 
 
 
 
    ε 
í ε  í
í í
í í
í í
Δεap Δεap
 N  N
í í

(a) (b)

Figure 8. Hysteresis loops in creep-fatigue conditions for σa = 219 MPa, and (a) τ = 5 s, (b) τ = 30 s.

εmax εmax
 
σ N N N N N
σ N N N N N
03D 03D
 

 

 
í    í   
ε ε
í í

í í
Δεap Δεap
 N
í  N
í

(a) (b)
Figure 9. Hysteresis loops in creep-fatigue conditions for σa = 258 MPa, and (a) τ = 5 s, (b) τ = 30 s.

It can be seen from Figures 8 and 9 that the creep time affects both the maximum strain
ε max and the range of plastic deformation in a cycle. Figure 10 shows the maximum strains
at two stress levels and three values of the dwell time τ.

  

 
εPD[
εPD[

 

τ V τ V
 
τ V τ V
τ  τ 
 

 
         
1XPEHURIF\FOHn 1XPEHURIF\FOHn
(a) (b)
Figure 10. Maximum strain on cycle ε max vs. number of cycles for different dwell times τ and
different stress amplitudes: (a) σa = 219 MPa, (b) σa = 258 MPa.

On the basis of the above results (Figures 8 and 9), it can be concluded that creep occurring
during the variable load (τ > 0) causes a significant increase in the maximum strain ε max .
With the increase in a dwell time, the elongation of the sample in the fatigue test increases
significantly. The analysis of the position of successive hysteresis loops (Figures 8 and 9) in the
function of the number of a cycle allows concluding that under the conditions of combined
variable and constant load, P91 steel does not exhibit a stabilization of its cyclic properties.

233
Materials 2022, 15, 4917

Regardless of the level of alternating stress amplitude σa, and the duration τ of the permanent
load, the steel always exhibits significant cyclic softening. This is illustrated in Figure 11
by the increase in plastic strain Δε ap as a function of the number of a load cycle for various
test conditions. In order to compare changes in plastic strains for different dwell times τ,
Figure 11b,d present results in the function of relative durability n/N.

 
Δεap, PPāPPí

1 − σa 03Dτ  1 − σa 03Dτ 

Δεap, PPāPPí
2 − σa 03Dτ V 2 − σa 03Dτ V
 
3 − σa 03Dτ V 3 − σa 03Dτ V
3
3
2
 2 
1
1

 
        
1XPEHURIF\FOHn n/N

(a) (b)
 

  1 − σa 03Dτ 


1 − σa 03Dτ 
2 − σa 03Dτ V
2 − σa 03Dτ V


 3 − σa 03Dτ V


Δεap, PPāPPí


Δεap, PPāPPí

3 − σa 03Dτ V
3
 
3 2
 
2 1
 1 

 
         
1XPEHURIF\FOH n n/N
(c) (d)
Figure 11. Changes in plastic strain on cycle Δε ap for different stress amplitude levels: (a,b) σa = 219 MPa;
(c,d) σa = 258 MPa.

Based on the analysis of changes in the plastic strain range Δε ap , it can be concluded
that the magnitude of changes in this parameter is influenced, among others, by the dwell
time τ. At the same amplitude stress levels, under the conditions of combined fatigue
and creep loads, plastic strains Δε ap are larger than under the conditions of pure fatigue
σa = const (δε2 > δε1 , see also Figure 6b).
During the experimental tests, a very clear effect on the fatigue life of the dwell time
τ was observed. The fatigue life experimental results obtained for various load program
sequences are summarized in Figure 12 in the form of fatigue diagrams. The durability
results at four stress levels were approximated in the semi-logarithmic coordinate system
by a regression equation of the form:

σa = a log N + b (2)

It can be observed that the creep time τ significantly affects the fatigue life. With the
increase of time τ, the fatigue life decreases. The results confirm the research described
in [21–23].

234
Materials 2022, 15, 4917

 τ 
  τ, s a b 
   í τ V
τ V
   í


'XUDELOLW\ N
   í


σa, 03D
GDWD 
 SRLQWV

 
  
 


    
    
'XUDELOLW\N σa03D

(a) (b)
Figure 12. Durability results (experiment): (a) fatigue curves, (b) summary of durability results.

3.5. Fractographic Observations


After testing, the samples were subjected to fractographic analysis. The fractographic
samples were prepared from the measurement parts of the samples, parallel to the load
direction. Figure 13 show the sample surfaces after selected variants of the load pro-
gram. For comparison, the surface of the P91 steel sample before the test is also included
in Figure 13a.

(a) (b)

(c) (d)

Figure 13. P91 steel microstructure: (a) the initial material, (b) after uniaxial tension test (T = 600 ◦ C),
(c) after creep test (T = 600◦ C, σ = 258 MPa), (d) after fatigue test (T = 600 ◦ C, σa = 258 MPa).

P91 steel (Figure 13a) is characterized by a typical microstructure for the class of steels
containing 8 ÷ 12% Cr; the microstructure of highly tempered martensite with numerous
precipitates of carbides and nitrides is observed. A lamellar microstructure of tempered
martensite with numerous precipitates visible at the boundaries of the former austenite
grains and martensite lamellae are visible. After the tensile test at T = 600 ◦ C (Figure 13b),
as expected, a highly deformed microstructure with visible banding is detected—the
direction of a material flow corresponds to the direction of the principal stress. Numerous
precipitations are visible on the deformation bands.
A similar structure as after the tensile test was observed after the creep test (Figure 13c).
The visible structure is also lamellar, with numerous precipitates.

235
Materials 2022, 15, 4917

The structure of the P91 steel after the fatigue test is also characterized by lamellar
microstructure (Figure 13d); however, the grain deformation is less pronounced than in the
creep test or monotonic test. The deformed microstructure shows the process of coagulation
of the precipitates. The main crack propagates transcrystalline, while the secondary cracks
are formed along the grain boundaries of the former austenite.
The results of microscopic analyses confirm the necessity regard creep damage during
the fatigue life calculations. However, the determination of the quantitative impact of
constant and variable load on changes in the microstructure requires further extensive
research program.

3.6. Durability Assessments


Currently, there are several hypotheses for summing the fatigue damage that evolves
under the conditions of simultaneously occurring fatigue and creep loads. These were
discussed, inter alia, in [14]. In the present research, a linear model was used for durability
calculations, which refers directly to the approach proposed in [13]. For the load program
shown in Figure 2, the total damage Dt will be equal to the sum of fatigue damage D f and
creep damage Dc :
D f + Dc = Dt . (3)
The experimental verification of the linear damage summation model under the
conditions of simultaneous creep and fatigue was then carried out. After realizing n cycles
of variable load with additional creep (τi > 0) in each load cycle, the total damage Dt will
be equal to:
n τ
∑ N + ∑ Tc = Dt (4)

where N is the number of cycles to failure in pure fatigue (σa = const and τ = 0), while Tc
denotes the total lifetime in creep (σ = const). It should be emphasized that the approach
described by Equation (4) is insensitive to the order of occurrence of particular types of load
and ignores changes in the microstructure of the material under various load conditions.
This problem was discussed, inter alia, in papers [10,11].
The results of calculations of damages D f , Dc , and Dt are summarized in Figure 14.

σ =σa, MPa Df Dc Dt  
 τ V τ V τ V
Dt
219 (τ=5 s) 0.11 1.49 1.60 
Ncalc>Nexp

τ V

219 (τ=30 s) 0.02 1.53 1.55 τ τ V
 A τ V V
τ V
225 (τ=5 s) 0.13 1.61 1.74 
Ncalc<Nexp


225 (τ=30 s) 0.02 1.38 1.40

250 (τ=5 s) 0.20 0.81 1.01 

250 (τ=30 s) 0.03 0.69 0.72

   
258 (τ=5 s) 0.24 0.85 1.10
6WUHVVDPSOLWXGHσa03D
258 (τ=30 s) 0.05 1.10 1.15
(a) (b)

Figure 14. (a) Damage levels resulting from calculations; (b) interpretation of results.

Based on the comparative analysis of the values of the total damage components Dt
(composed of D f and Dc ), it can be concluded that the dominant component is the creep
damage, Dc . The application of the linear model of damage summation under the condi-
tions of simultaneous occurrence of static load (creep) and variable load (fatigue) results in
obtaining total damage larger than unity (Dt > 1) for most of the analyzed load program
sequences. Consequently, the calculated durability is lower than that obtained from the tests.

236
Materials 2022, 15, 4917

This conclusion is confirmed in Figure 15, presenting durability diagrams obtained from
calculations and experimental tests. The figure includes the following graphs:
(1) a calculation-based diagram taking into account only the fatigue properties (fatigue);
(2) a calculation-based diagram regarding fatigue and creep (fatigue + creep);
(3) a diagram obtained based on experimental results (experiment).

 GLDJUDP GHVFULSWLRQ a b  GLDJUDP GHVFULSWLRQ a b


 IDWLJXH  í
  IDWLJXH  í   IDWLJXHFUHHS  í
 IDWLJXHFUHHS  í
6WUHVVDPSOLWXGHσa03D

6WUHVVDPSOLWXGHσa03D
 H[SHULPHQW  í  H[SHULPHQW  í
 
σa=a logN+b σa=a logN+b
 
  
    

 GDWD  GDWD


SRLQWV SRLQWV
 
 
         
'XUDELOLW\N 'XUDELOLW\N
(a) (b)
Figure 15. Durability diagrams (experimental and calculated): (a) τ = 5 s, (b) τ = 30 s.

Analysis of the results presented in Figure 15a,b indicates that regarding creep during
the fatigue life calculations significantly improves the efficiency of the durability assess-
ments. If both the creep and fatigue damage is considered within the classical damage sum-
mation approach (4), the predicted durability calculations are lower but close to the results
of durability obtained from the experimental tests. If only the fatigue damage is considered,
the simulated durability is much larger than the real one observed in the experiment.

4. Conclusions
During fatigue tests on P91 steel samples at the temperature of 600 ◦ C under alternating
variable and constant loads, cyclic softening of the material was observed, both for the
programs containing only the pure fatigue load cycles and for the cases when alternating
fatigue and creep took place. The amount of P91 steel softening is influenced by both the
load level and the dwell time (creep). With the increase of the dwell time in the cycles, the
cyclic softening becomes more pronounced.
The increase in the static load time (τi > 0) causes a marked increase in the cyclic creep
observed during classic fatigue tests conducted under the conditions of σa = const (τi = 0),
and in the range of inelastic stain on a cycle Δε ap .
The dwell periods occurring during the variable load reduce the fatigue life. The dwell
time value influences the durability.
Fatigue life predictions that do not regard the material damage due to creep may lead
to a significant overestimation of durability. Taking into account the creep damage in the
calculations improves the compliance of the numerical and experimental results.

Author Contributions: S.M.: Supervision, Conceptualization, Methodology, Investigation, Valida-


tion, Visualization, Writing—Original draft preparation. Z.L.: Software, Visualization, Investigation.
H.E.: Methodology, Writing—Reviewing and Editing, Project administration, Funding acquisition.
All authors have read and agreed to the published version of the manuscript.
Funding: This research was funded by the NATIONAL SCIENCE CENTRE of Poland through Grant
No. 2017/25/B/ST8/02256.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.

237
Materials 2022, 15, 4917

Data Availability Statement: Not applicable.


Conflicts of Interest: The authors declare no conflict of interest.

References
1. Mroziński, S.; Piotrowski, M.; Egner, H. Effects of fatigue testing on low-cycle properties of P91 steel. Int. J. Fatigue 2019, 120,
65–72. [CrossRef]
2. Mroziński, S.; Lis, Z. Comparative Analysis of Methods for Determining Cyclic Properties of Metals, Experimental Mechanics of
Solids. Materials Research Forum RLC. Mater. Res. Proc. 2019, 12, 139–145.
3. Wu, D.-L.; Zhao, P.; Wang, Q.-Q.; Xuan, F.-Z. Cyclic behavior of 9–12% Cr steel under different control modes in low cycle regime:
A comparative study. Int. J. Fatigue 2015, 70, 114–122. [CrossRef]
4. Egner, W.; Sulich, P.; Mroziński, S.; Egner, H. Modelling thermo-mechanical cyclic behavior of P91 steel. Int. J. Plast. 2020, 135,
102820. [CrossRef]
5. Hu, X.; Zhang, Q.; Jiang, Y.; Rao, G.; Miao, G.; He, W.; Nie, X. The effect of cyclic loading on the creep fatigue life and creep
strength of a DS superalloy: Damage mechanism and life modeling. Int. J. Fatigue 2020, 134, 105452. [CrossRef]
6. Xu, L.; Wang, R.-Z.; Wang, J.; He, L.; Itoh, T.; Miura, H.; Zhang, X.-C.; Tu, S.-T. On multiaxial creep–fatigue considering the
non-proportional loading effect: Constitutive modeling, deformation mechanism, and life prediction. Int. J. Plast. 2022, 155,
103337. [CrossRef]
7. Liu, H.; Yang, X.; Jiang, L.; Lv, S.; Huang, T.; Yang, Y. Fatigue-creep damage interaction model of asphalt mixture under the
semi-sine cycle loading. Constr. Build. Mater. 2020, 251, 119070. [CrossRef]
8. Barat, K.; Sivaprasad, S.; Kar, S.K.; Tarafder, S. A novel rate based methodology for creep fatigue life estimation of superalloys.
Int. J. Press. Vessel. Pip. 2020, 182, 104064. [CrossRef]
9. Igumnov, L.A.; Volkov, I.A. Defining relations of mechanics of damaged media effected by fatigue and creep. Procedia Struct.
Integr. 2020, 28, 2086–2098. [CrossRef]
10. Mroziński, S.; Lis, Z. Research on the influence of creep on low-cycle fatigue life. Energetyka 2019, 11, 760–762(In Polish).
11. Mroziński, S.; Lis, Z.; Egner, H. Energy Dissipated in Fatigue and Creep Conditions. Materials 2021, 14, 4724. [CrossRef] [PubMed]
12. Tien, J.K.; Nair, S.V.; Nardone, V.C. Creep fatigue Interactions in Structural Alloys. In Flow and Fracture at Elevated Temperatures;
Raj, R., Ed.; ASM Materials Science Seminar: Philadelphia, PA, USA, 1983; Chapter 6; p. 179.
13. Miner, M. Cumulative Damage in Fatigue. J. Appl. Mech. 1945, 67, A159–A164. [CrossRef]
14. Zhuang, W.Z.; Swansson, N.S. Thermo-Mechanical Fatigue Life Prediction: A Critical Review; DSTO-TR-0609; DSTO Aeronautical and
Maritime Research Laboratory: Victoria, Australia, 1998.
15. Ahmadzadeh, G.R.; Varvani-Farahani, A. Concurrent ratcheting–fatigue damage analysis of uniaxially loaded A-516 Gr.70 and
42CrMo Steels. Fatigue Fract. Eng. Mater. Struct. 2012, 35, 962–970. [CrossRef]
16. Chang, L.; Wen, J.B.; Zhou, C.Y.; Zhou, B.B.; Li, J. Uniaxial ratcheting behavior and fatigue life models of commercial pure
titanium. Fatigue Fract. Eng. Mater. Struct. 2018, 41, 2024–2039. [CrossRef]
17. Samuel, K.G.; Rodriguez, P. An empirical relation between strain energy and time in creep deformation. Int. J. Press. Vessel. Pip.
1998, 75, 939–943. [CrossRef]
18. Payten, W.M.; Dean, D.W.; Snowden, K.U. A strain energy density method for the prediction of creep–fatigue damage in high
temperature components. Mater. Sci. Eng. A 2010, 527, 1920–1925. [CrossRef]
19. Cheng, H.; Chen, G.; Zhang, Z.; Chen, X. Uniaxial ratcheting behaviors of Zircaloy-4 tubes at 400 C. J. Nucl. Mater. 2015, 458,
129–137. [CrossRef]
20. Song, G.; Hyun, J.; Ha, J. Creep-fatigue prediction of aged 13CrMo44 steel using the tensile plastic strain energy. Eur. Struct.
Integr. Soc. 2002, 29, 65–73.
21. Zheng, X.T.; Xuan, F.Z.; Zhao, P. Ratcheting-creep interaction of advanced 9–12% chromium ferrite steel with anelastic effect. Int.
J. Fatigue 2011, 33, 1286–1291. [CrossRef]
22. Zhao, Z.; Yu, D.; Chen, X. Creep-ratcheting-fatigue life prediction of bainitic 2.25Cr1MoV steel. Procedia Struct. Integr. 2019, 17,
555–561. [CrossRef]
23. Zhao, Z.; Yu, D.; Chen, G.; Chen, X. Ratcheting fatigue behaviour of bainite 2.25Cr1MoV steel with tensile and compressed hold
loading at 455 ◦ C. Fatigue Fract. Eng. Mater. Struct. 2019, 42, 1937–1949. [CrossRef]
24. ASTM E606-92; Standard Practice for Strain—Controlled Fatigue Testing. ASTM: West Conshohocken, PA, USA, 1998.
25. Skocki, R. Research on the Influence of Elevated Temperature on the Cyclic Properties of P91 Steel. Ph.D. Thesis, Faculty of
Mechanics UST, Bydgoszcz, Poland, 2017. (In Polish).
26. Mroziński, S.; Lipski, A. The effects of temperature on the strength properties of aluminium alloy 2024-T3. Acta Mech. Autom.
2012, 6, 62–66.
27. Oh, Y.-J.; Yang, W.-J.; Jung, J.-G.; Choi, W.-D. Thermomechanical fatigue behavior and lifetime prediction of niobium bearing
ferritic stainless steels. Int. J. Fatigue 2012, 40, 36–42. [CrossRef]
28. Brnic, J.; Turkalk, G.; Canadija, M.; Lanc, D. AISI 316Ti (1.4571) steel—Mechanical, creep and fracture properties vs. temperature.
J. Constr. Steel Res. 2011, 67, 1948–1952. [CrossRef]

238
materials
Article
Experiment and Theoretical Investigation on Fatigue Life
Prediction of Fracturing Pumpheads Based on a Novel
Stress-Field Intensity Approach
Yun Zeng 1 , Meiqiu Li 1 , Han Wu 2 , Ning Li 1 and Yang Zhou 3, *

1 School of Mechanical Engineering, Yangtze University, Jingzhou 434023, China; [email protected] (Y.Z.);
[email protected] (M.L.); [email protected] (N.L.)
2 CNOOC Ener Tech-Drilling & Production Co., Shanghai 200444, China; [email protected]
3 School of Mechatronic Engineering and Automation, Shanghai University, Shanghai 200444, China
* Correspondence: [email protected]

Abstract: Fracturing pumpheads are typical pressure vessels that experience frequent fatigue failure
under the effect of notches in their cross-bore. To enhance the fatigue life of fracturing pumpheads,
the study of the notch effect is indispensable and important to establish a reliable mathematical
model to predict their fatigue life. In the present paper, two novel fatigue life prediction models are
proposed for notched specimens. In these models, two new geometric fatigue failure regions are
defined to improve the weight function. Finally, the elaborated novel stress-field intensity approach
was applied to three different types of notched specimens. Experiment results indicate that the new
SFI approach achieves 47.82%, 39.48%, and 31.85% higher prediction accuracy than the traditional
SFI approach, respectively. It was found that the modified SFI approach provided better predictions
than the traditional SFI approach and the TCD method. The II-th novel SFI approach had the highest
Citation: Zeng, Y.; Li, M.; Wu, H.; Li,
accuracy, and the I-th novel SFI approach was more suitable for sharply notched specimens.
N.; Zhou, Y. Experiment and
Theoretical Investigation on Fatigue
Keywords: fatigue life prediction; notched specimen; S-N curves; fatigue fracture region;
Life Prediction of Fracturing
stress-field intensity
Pumpheads Based on a Novel
Stress-Field Intensity Approach.
Materials 2022, 15, 4413. https://
doi.org/10.3390/ma15134413
1. Introduction
Academic Editors: Jaroslaw
In the oil and gas exploration field, with the rapid development of unconventional
Galkiewicz, Lucjan Śnieżek and
Sebastian Lipiec
oil and gas drilling technology, requirements for fracturing and acidizing technology have
also gradually increased. Therefore, as important components of the equipment, fracturing
Received: 13 May 2022 pumps require long life and high reliability [1]. However, the average life of pumpheads is
Accepted: 20 June 2022 much lower than that of other conventional components, and the manufacturing cost of
Published: 22 June 2022 pumphead components is surprisingly high, seriously affecting the economic benefits of
Publisher’s Note: MDPI stays neutral shale gas [2,3]. Hence, it is important to accurately estimate the fatigue life of pumpheads
with regard to jurisdictional claims in to improve the performance of fracturing pumps.
published maps and institutional affil- The nominal stress method [4,5], the critical distance method [6–9], and the effective
iations. stress volume method [10,11] are most commonly used to estimate the high-cycle fatigue
life of notched components. The accuracy of NSM mainly depends on the correction factor
(stress concentration factor loading, stress size coefficient, roughness coefficient) and S-N
curves of materials. However, owing to the complexity in petroleum engineering, empirical
Copyright: © 2022 by the authors. equations are often used to calculate the coefficients of NSM. In addition, it is proven that
Licensee MDPI, Basel, Switzerland.
NSM is not suitable for predicting the fatigue life of complex structural parts [12].
This article is an open access article
NSM considers the “hot point stress” as the core parameter of fatigue failure; however,
distributed under the terms and
some advanced volume methods [13–15] indicate that the “volume stress” in a damaged
conditions of the Creative Commons
region should also be considered for fatigue failure. The critical distance method and the
Attribution (CC BY) license (https://
effective stress volume method are two main advanced volume methods. The theory of
creativecommons.org/licenses/by/
critical distance method was first proposed by Taylor and Tanaka [16–21] based on LEFM.
4.0/).

Materials 2022, 15, 4413. https://doi.org/10.3390/ma15134413 239 https://www.mdpi.com/journal/materials


Materials 2022, 15, 4413

The average stress of the maximum principal stress field in the critical region near a notch
root (critical region can be a point, line, area, or volume) is regarded as a fatigue damage
parameter for fatigue life estimation. The formulas for different forms of TCD can be found
in [17–20]. The methods are schematically shown in Figure 1, and the formulas of the
methods are as follows:
 
PointMethod (PM) : σmean = Δσ1 r = L20 , θ = 0
% 2L
LineMethod (LM) : σmean = 2L1 0 0 0 σ1 (r, θ = 0)dr
% π/2 % L0 (1)
AreaMethod (AM) : σmean = 1.1πL 2
σ1 (r, θ )rdrdθ
0 − π/2 % 0 %
2
%
2π π/2 1.54L0
VolumeMethod (VM) : σmean = 3
3 0 0 0 σ1 (r, θ, ϕ)r2 sin θrdrdθdϕ
2π (1.54L0 )


2
1 ΔKth
El Haddad equation: L0 = (2)
π Δσ−1

Figure 1. Visualization of the point method, the line method, the area method, and the volume method.

Qylafku [22], Adib [23], and Pluvinage [24] have defined the effective distance as the
length between the minimum point of the stress gradient and a notch root. Therefore, there
is a certain relationship between fatigue damage and the stress gradient. The effective
stress volume method introduces the concept of the effective fatigue failure region, and
the calculation is used to obtain the effective stress by integrating the elastic–plastic fa-
tigue crack-opening stress and the weight function of the effective critical damage region
(schematically shown in Figure 2). The mathematical definition of the effective stress
function (σeff ) is expressed by Equation (3):
%X
σeff = X1 0 eff σyy ( x ) × ϕ( x, χ)dx
 eff  (3)
 1 dσyy ( x) 
χ =  σmax dr 

where Xeff is the effective distance diameter of the fatigue process zone, σyy (r ) is the
fatigue crack-opening stress, ϕ(r, χ) is the relative stress gradient, and Kρ is the notch stress
intensity factor.

240
Materials 2022, 15, 4413

Figure 2. Schematic of elastic–plastic stress distribution along the notch ligament and the relative
stress gradient concept.

In addition, these effective stress volume approaches have been verified in simple
geometric structures; thus, it is unknown whether they are suitable for fatigue life estima-
tion of thick-walled pressure vessels. The mean value of the fatigue failure parameter, l0
(critical distance), for aluminum alloys, cast iron, and steel is equal to 0.113 mm [25,26].
The initiation of fatigue cracks depends on grain size, grain dislocations, grain slips, and
other features of microstructures. Especially, the initiation of fatigue cracks is a grain
movement nucleation process at the mesoscopic scale. To simplify the calculation of the
fatigue process zone, the fatigue failure parameter, l0 , can be defined as the field diameter
of the fatigue failure region. In addition, the plastic zone is also assumed as the fatigue
damage failure region. Numerous empirical formulas are developed to calculate the cyclic
plastic zone size [27–33] (Table 1). However, the fatigue plastic yielding zones of specimens
and structures are different. In fracture mechanics, the fatigue damage failure region and
the plastic yield region are defined as the fracture tip plastic zone and the crack initiation
region, respectively [34–38]. To investigate the effects of the cyclic plastic zone on notch tip
stress and fatigue life, Zhu et al. [34,35] proposed a novel approach.
Different studies have proposed different fatigue failure parameters to estimate the
fatigue life of engineering components; however, these approaches are only verified on
notched specimens of regular components. Fracturing pumpheads have a complex circular
cross-bore structure; therefore, it is doubtful whether the fatigue failure region size can be
defined as a fatigue failure parameter to predict the fatigue life of pumpheads.
In the current study, a new fatigue life prediction model was proposed for notched
specimens and a novel cardioid geometric fatigue failure region was defined to improve
the weight function. In addition, existing problems of the original SFI approach were
corrected. The accuracy of the modified SFI approach was verified in fatigue tests on
notched specimens and based on theoretical approaches for improving the fatigue life
of pumpheads.

241
Materials 2022, 15, 4413

Table 1. Empirical formulas for cyclic plastic zone size calculation.

Reference Formula

2
σ
Nicholls and Martin [27] r  y = a σ2 −app
y σapp
2
 2
Bathias and Pelloux [31] 
r y = 0.1 σy ΔK
 2
Pineau and Pelloux [28] r  y = 0.053 ΔK σ
 2y +s
Saxena and Antolovich [29] r  y = α ΔK σ
y 2
Park et al. [30] r  y = 144
π ΔK
σ
 y 2
Chapetti et al. [21] r  y = 12π
1 ΔK
σ
 y 2
Edmunds and Willis [33] r  y = 24π
1 ΔK
σy
σapp represents the applied stress, and α and s denote correlation and variation coefficients, respectively.

2. Theoretical Analysis
2.1. Brief Review of Stress-Field Intensity Approach and Characteristic Parameters
2.1.1. Traditional Stress-Field Intensity Approach
The fatigue failure of materials occurs due to the propagation of fatigue cracks from
grains. The formation of cracks can be defined as the cumulative damage of multiple grains
in local areas of materials. The SFI model proposed by Yao [39] is displayed in Figure 3,
and the mathematical definition of the stress-field intensity function (σFI ) is expressed as:
" "
1 → 1
σFI = f (σij ) ϕ( r )dv = σ(r−θ ) ϕ(r, θ )dv (4)
V V
Ω Ω

where σFI is the stress-field intensity of a notched specimen, Ω is the fatigue failure region,
V is the volume of the fatigue failure region, R is the field diameter, f (σij ) is the equivalent
stress function, and ϕ(r, θ ) is the weight
 function, which signifies the contribution of
 
stresses at point P to the peak stress at  r .

³
y FI
³

max r-ș
P
ș
ȍ
r

x
³

FI

z ȱ
Figure 3. Basic model of the stress-field intensity approach.

2.1.2. Parameters for Critical Fatigue State


All advanced volume methods generally include a “characteristic region parame-
ter”. Table 2 summarizes some characteristic region parameters proposed in previous
papers [6,9–11,17]. Although these methods were previously successfully verified, they
possess some common problems.

242
Materials 2022, 15, 4413

Table 2. Fatigue damage regions of different approaches.

Approach Authors Fatigue Damage Region Influence Factors


Neuber’s and Peterson’s
Neuber and Peterson [40,41] Material constants, a1 , a2 a1 , a2 depend on fracture strength σb
empirical formula
R0 depends on the threshold value of fatigue crack
SED approach Lazzarin and Berto [42] Radius of critical volume (area), R0 propagation, fatigue limit, Poisson’s ratio, and
notch shape
 
SFI approach Yao [41] Radius of the fatigue damage field, r r depends on grain size
Modified SFI approach Qylafku [22] Radius of the fatigue damage field, r r depends on metallic-plastic deformation
l0 depends on the threshold value of fatigue crack
TCD and M-TCD Taylor and Susmel [19] Intrinsic crack length, l0
propagation, fatigue limit, and stress ratio
Advanced volumetric method Adib-Ramezani [10] Effective distance, Xeff Xeff depends on stress distributions at notch roots

(1) The acquisition of critical distance parameters requires the testing of numerous
material constants; however, it takes a long time to experimentally obtain these mate-
rial constants.
(2) The geometry of a fatigue-damaged failure area is generally defined as a common
shape; hence, it is difficult to confirm whether previous advanced volume methods are
suitable for calculating the fatigue damage failure region of complex pressure vessels.
(3) Owing to the size effect, the geometry and size of fatigue damage failure regions
need to be reconsidered.

2.2. Determination of Solutions for New Fatigue Failure Regions


2.2.1. Comparison of Stress Gradient Distributions near Different Notch Roots
Generally, engineering structures have different and complex geometries; thus, it is
difficult to select a fatigue failure region. Therefore, the determination of fatigue failure
regions for notched components with common geometries is a better choice. According to
the research of Susmel and Taylor [19], four different notched specimens were designed
(Figure 4). Generally, FEM was applied to calculate the bisector stress distributions of the
notched specimens, and LEFM and EPFM were used for calculations.
In the paper, finite element analysis was performed on the specimen in Figure 4, with
a fixed constraint on one end and a load on the other end, and the stress distribution in
the elastic state of material at the notch root bisector of each specimen was elaborated
and represented by a bi-logarithmic coordinate system. Furthermore, the corresponding
relative stress gradient was calculated by the volumetric method and the SFI method
(Figure 5). Owing to the notch effect, relative stress gradients near the notch root of
the four specimens had different forms. However, the variation trends of relative stress
gradients for the four specimens were similar. According to the theory of critical distance
method (TCD) [14,15,17], the minimum point of a relative stress gradient was defined
as the characteristic region parameter. Figure 5 displays the elastic stress distributions
and relative stress gradients near the notch root of the four specimens. Some numerical
truncation errors appear based on the discrete feature of numerical calculations. Since
plastic action is not considered, the relative stress gradient trend of linear elastic analysis
can only be used as a reference to judge the trend. In Figure 5a,b, the trend of log(σ0 /σyy )
is to be linear in zone II. On the contrary, in Figure 5c,d, the trend of log(σ0 /σyy ) is to be
linear in zone III. Therefore, elastoplastic analysis is also required to perform a comparative
analysis of the differences between zone II and zone III.

243
Materials 2022, 15, 4413

25 25 25
ı025

4 5

ĭ3.5
ĭ8

150
150
150

150
60°

R=1.5
R=0.1

(a) (b) (c) (d)


Big circle hole Little circle hole Sharp V-notch U-notch plate
plate plate plate
ȱ
Figure 4. Geometries of different notched specimens (thickness = 6 mm, units: mm).

Figure 5. Bi-logarithmic diagrams for stress distributions at the notch bisector of different specimens
at the elastic state of material. (a) big circle hole plate, (b) little circle hole plate, (c) sharp V-notch
plate, (d) U-notch plate.

244
Materials 2022, 15, 4413

Moreover, fatigue process zones resulted from the accumulation of the fatigue damage
region, where micro- or macro-plastic cyclic strains occur. The fatigue failure region radii
of the four specimens were calculated based on stress distributions for the elastic–plastic
state at the notch root bisector and relative stress gradients in a bi-logarithmic coordinate
system. Figure 6 displays the elastic–plastic stress distributions and relative stress gradients
near the notch root of the four specimens. The peak in stress for the elastic state always
appeared at the notch root surface (the origin of the X-axis in the Cartesian coordinate
system). However, the stress peak in the elastic stress appeared at a certain distance from
the notch root. The point of the stress peak and its counterpart distance, Xm , were defined
as zone I, which was a completely plastic region in the bi-logarithmic coordinate system. In
zone II, the peak stress manifested a decreasing trend. The distribution trends of relative
stress gradients in zones II and III were opposite, especially the relative stress gradient,
which dropped gradually in zone II. The point at which the relative stress gradient started
to manifest an increasing trend was considered as the effective distance. In zone III, the
elastic–plastic stress distribution was linear; thus, a gradual plastic to elastic material
transition occurred in this region. In zone IV, the elastic–plastic stress distribution was no
longer linear, and the relative stress gradient expressed a new decreasing trend. The relative
stress gradient was far from the notch root and had a minimum effect on fatigue failure.

ȱ ȱ

ȱ ȱ
Figure 6. Bi-logarithmic diagrams for elastic–plastic stress distributions at the notch bisector of
different specimens. (a) big circle hole plate, (b) little circle hole plate, (c) sharp V-notch plate,
(d) U-notch plate.

2.2.2. Summary of Fatigue Damage Failure Area Shapes for Different Notched Specimens
It is propounded that in addition to the “stress peak point”, the stress field in a specific
region should also be considered during fatigue life analysis [6,7,36,43–46]. TCD, the
volumetric method, and the SFI approach are defined as macro-mechanical methods; thus,
failure criteria of these methods should be combined with the critical damage region. In
this work, the SFI approach was used to predict the fatigue life of fracturing pumpheads.

245
Materials 2022, 15, 4413

The shape and size of the fatigue damage region are key factors for accurate fatigue
life prediction.
Figure 7 displays the stress nephograms for the elastic–plastic stress at the notch
root of the four specimens under a pressure of 251 MPa. The high-stress contours of the
circular plate specimen had a crescent shape, whereas those of the V-notched and U-notched
specimens had a heart shape. Obviously, regions enclosed by these high-stress contour
lines were plastic. Elastic–plastic transition zones near high-stress plastic zones also had
an important influence on fatigue failure. It was found that elastic–plastic stress contour
lines of the four notched specimens manifested roughly similar regions with a heart shape.
Therefore, it is feasible to assume that the shape of the fatigue failure region was cardioid.

Figure 7. Elastic–plastic stress nephograms at the notch root of the four specimens under pressure of
251 MPa. (a) big circle hole plate, (b) little circle hole plate, (c) sharp V-notch plate, (d) U-notch plate.

2.3. Improvement of the Stress-Field Intensity Approach


2.3.1. Traditional Stress-Field Intensity Approach
The traditional SFI approach was modified to effectively predict the fatigue life of
notched components. However, the traditional SFI approach and other critical damage
region approaches have several common problems.
1. Owing to the randomness of grain damage in fatigue damage failure regions and
the influence of stress gradients near notch roots, uniform shapes cannot be used to define
damage regions in notched components.
2. Micro-plastic cyclic strains are one of the main reasons for fatigue failure. However,
macro-mechanical methods do not consider the influence of microcosmic factors. Hence,
material parameters at the microstructural scale are essential factors for fatigue failure.
3. The mechanical resistance of microstructure barriers inhibits the propagation of
short cracks. However, whether grains near the peak stress point at a notch root play a
“contribution” or “hindrance” role in the initiation and propagation of cracks has not been
clearly defined.

246
Materials 2022, 15, 4413

2.3.2. Optimization of the Traditional Stress-Field Intensity Approach


The accuracy of the traditional SFI approach depends on the failure equivalent stress
function, the fatigue failure region, and the weight function. In Section 2.3.1, it is asserted
that the effect of inner grains existing far away from a high-stress region also needs to be
considered for fatigue life prediction. Therefore, the revised formula can be expressed as:
"
1 →
σFI = f (σij ) ϕ( r )dv = σmax − ξ (Ω) (5)
V
Ω

where ξ (Ω) is the “auxiliary part”, which can be expressed as:


%
(σmax − σr−θ ) ϕ(r, θ )dv
ξ (Ω) = Ω % (6)
ϕ(r, θ )dv
Ω

The novel method proposed in this work is different from the traditional SFI approach
and does not require an artificial field diameter. It was considered that the continuous
accumulation of damages resulted from the combined effect of the peak stress point and
other points in the damage region. The hypothetical concept of an “invisible boundary
fatigue failure region” is illustrated in Figure 8. In Figure 8a, rpc is the size of the cyclic
plastic zone, and rpm is the size of the monotonic plastic zone.

Figure 8. Schematic diagram of the novel elastic–plastic zone at the crack tip (notch): (a) monotonic
plastic zone and cyclic plastic zone (b) invisible boundary region.

The weight function of the traditional SFI approach is expressed as:


 
 1 dσr−θ 
ϕ(r, θ ) = 1 − χr (1 + sin θ ) = 1 −  r (1 + sin θ ) (7)
σmax dr 

The weight function of the  traditional SFI approach is a generalized monotonically

decreasing function about  r . However, it is also reported that as the distance from a
notch root increases, the variation trend of the weight function does not comply with the
original definition [36]. Therefore, it is important to correct the weight function for accurate
fatigue life prediction. It can be inferred from Equation (9) that the weight function is
mainly composed of three parts: relative stress gradient, distance function of the notch root,
and function of angle θ.
In this work, common problems of the weight function of the traditional SFI approach,
similar to those mentioned in Section 2.3.1, are summarized and revised.
(1) The direction angle of the relative stress gradient is not clearly expressed by the

traditional weight function. Hence, the variable of distance r for a notch root was revised
to a distance function r (θ ).

247
Materials 2022, 15, 4413

(2) Grain size is an essential factor for fatigue life prediction; therefore, it was intro-
duced as a coefficient in the calculation formula of the weight function of the SFI approach.
(3) The weight function defines the contribution of stresses at point P to the peak stress

at |r |; however, considering the inhibition effect of the “auxiliary part”, (1 + sin θ ) is no
longer applicable. Hence, the function f (θ ) of angle θ was proposed to revise the weight
function.
(4) Many researchers [43,47,48] have also proven that the grain size is related to value
of the stress-intensity field; moreover, the grain size means the scale of the grain dimension.
Therefore, the authors consider the grain size of the material to modify the weight function.
Considering the above factors and the function of ξ (Ω) as a separate part, the revised
formula can be expressed as:

  r (θ )
 1 dσr−θ  G
ϕ( x, θ ) =  r ( θ ) f ( θ ) (8)
 σmax dr 

where G is the grain size and r(θ) is the geometric relationship function of stress contour
distribution in a damaged region.

2.4. Hypothesis of the “Cardioid” Fatigue Damage Failure Region


Finite element analysis results revealed that fatigue failure regions of the notched
specimens had two geometrical shapes. In Figure 7, the highly stressed damage volume
geometry of the plate specimen with holes has a crescent shape, whereas those of the
U-notched and V-notched specimens have a heart shape.

2.4.1. Crescent-Shaped Fatigue Failure Region


The relationship between stress contours and the peak point is geometrically expressed
in Figure 9. To simplify the calculation, the approximate geometric relationships of some
parameters are displayed in Figure 9b, and the corresponding differential equations are
expressed as:

⎪ OA = OA1 = OD = ri


⎪ OD = OP


⎨  + PD = ri = Δr +r √
A1 P = OA1 2 − OP2 = ri 2 − Δr2 = r2 + 2rΔr
(9)


⎪ AC =
⎪ r 2 − (Δr − r0 )2

⎪ i 

AP = ri 2 − (Δr − r0 )2 + r0 2 = r2 + 2Δr (r + r0 )

To solve these differential equations, the results of r(θ) were considered infinite. There-
fore, in Figure 9b, lines A1 P and PD are two special boundary solutions, which can be
expressed as: 
& π
r( 2 ) = ri 2 − Δr2 , r (0) = r
 (10)
r (θ )|θ = 0 = 0
Hence,
θ 
r (θ ) = (2 sin2 )( r2 + 2rΔr − r ) + r (11)
2

248
Materials 2022, 15, 4413

Figure 9. Schematic representation of the notched specimen with holes. (a) Stress contour lines at the
notch root. (b) Relationship between stress contours and the peak point.

2.4.2. Cardioid-Shaped Fatigue Failure Region


According to FEA results, the geometry volume of the fatigue failure region of the
U-notched and V-notched specimens in the elastic–plastic stress had a cardioid shape.
The relationship between stress contours and the peak point was expressed by periodic
distribution (Figure 10).
⎧ 2 
⎨ x + y + ax = a x2 + y2
2

(r (θ ))2 + ar (θ ) cos θ = ar (θ ) (12)



r (0) = r

Figure 10. Schematic representation of the U-notched specimen. (a) Cartesian cardioid. (b) Relation-
ship between stress contours and the peak point.

249
Materials 2022, 15, 4413

In addition, according to the concept of the “invisible boundary damage region”, the
field diameter of an invisible boundary does not require an artificial definition. As the
distance increased, the “inhibition” gradient (dΩ/dr) reached zero; thus, the distance was
determined as the field diameter. Therefore, the field diameter of the cardioid fatigue
failure region can be redefined as:
&
2a = r
(13)
r (θ ) = a(1 − cos θ )

2.4.3. Simplification of the Advanced SFI Approach


In the traditional SFI approach, the function of angle θ is expressed as (1 + sinθ). In
addition, the contributions of different stress contours to fatigue failure require different
field diameters; however, none of these contributions exceed the corresponding stress
values. Therefore, the function of angle θ can be redefined as:

f (θ ) = cos θ (14)

In addition, based on Equations (8), (11)–(14), the field-intensity formula can be


expressed as:

%    r (θ )
 1 dσr−θ  G
(σmax − σr−θ )  σmax dr r ( θ ) cos( θ ) dv
σFI = σmax − Ω (15)
%  1 dσr−θ   r (θ )
G
 σmax dr r (θ ) cos(θ ) dv
Ω
& √
I − crescent : r (θ ) = (2 sin2 2θ )( r2 + 2rΔr − r ) + r
(16)
I I − cardioid : r (θ ) = 2r (1 − cos θ )
Considering the symmetry of a notched specimen, the simplified field-intensity calcu-
lation formula can be expressed as:

%∞%    r (θ )
π
 1 dσr  G
0 0
2
(σmax − σr )  σmax dr r ( θ ) cos( θ ) dθdr
σFI = σmax −    r (θ )
(17)
% ∞ % π2  1 dσr  G
0 0  σmax dr r ( θ ) cos( θ ) dθdr

3. Materials and Experiment


3.1. Specimen Geometry
The geometry and dimensions of the smooth solid-bar specimen used for the S-N
curve test are presented in Figure 11a. In addition, three notched specimens of different
elastic SCFs (kt) were designed to verify the reliability of the new approach, and their
geometries and dimensions are presented in Figure 11b–d. All fatigue experiments were
conducted on test equipment (Figure 12).

250
Materials 2022, 15, 4413

Figure 11. Geometries and dimensions of test specimens (in mm) used for S-N curve and notch
specimens tests: (a) smooth specimen, and (b) circumferentially notched solid-bar specimen with
kt = 2, (c) kt = 3, and (d) kt = 5.

Figure 12. Illustration of the specimen clamping mechanical testing system.

3.2. S-N Curve Test Procedure


3.2.1. Static Mechanical Properties of Test Material
According to the GB/T 3075-2008 standard and ISO 6892:1998, the mechanical properties
of the test material (40CrNi2 MoVA) were tested prior to the S-N curve experiment (Table 3).

251
Materials 2022, 15, 4413

Table 3. Static mechanical properties of nickel-chromium alloy.

Material Serial Number Young’s Modulus Poisson’s Ratio Yield Stress (sY ) Ultimate Tensile Strength (σ b )
40Cr Ni2 MoV 1 204 GPa 0.3 980 MPa 1060 MPa
40Cr Ni2 MoV 2 204 GPa 0.3 980 + 3 MPa 1060 + 3 MPa
40Cr Ni2 MoV 3 204 GPa 0.3 980 − 2 MPa 1060 − 3 MPa

3.2.2. Uniaxial Test


Uniaxial fatigue tests were conducted on a PLD-300 electro-hydraulic servo fatigue
testing machine (Figure 12), and an axial tensile sinusoidal waveform load was selected to
control the process. The stress ratio (R = −1) was constant under the varying frequencies
of 10–30 Hz. Specimens’ buckling was reduced to a minimum because the shapes of
measurement distances were represented by hourglass sections and notches that become
the weakest regions. It means fatigue damages have occurred in the mentioned zones, while
avoiding the other ones (Figure 12). According to the value of ultimate tensile strength
(σb ), 35 specimens were divided into 7 groups with different stress levels. According to the
GB/T 24176-2009 standard and ISO 12107:2003, the confidence level of the S-N curve of
40CrNi2 MoV was obtained as 50%. Nf is the fatigue life recorded for each experiment. S-N
curve fitting was carried out by using the two-parameter method, and the fitting formula
can be expressed as:
σm × N = C (18)
where m and C are fitted values, σ is the test stress, and N is the number of cycles to fracture.
Results of the S-N curve experiment are presented in Table 4, and the fitted S-N curve
is plotted in Figure 13.

Figure 13. Fatigue data and fitted S-N curve for the pumphead material.

252
Materials 2022, 15, 4413

Table 4. Empirical data of different fatigue test groups for the pumphead material.

Material Stress (Smax , MPa) Nf (Life) Median (x)


Level/count 1 2 3 4 5 ——
756.3 1037 1055 980 957 1123 1030.4
716.5 5037 4958 5968 7853 6842 6131.6

40CrNi2 MoV 636.9 35,600 25,500 20,360 31,186 32,100 28,949.2


517.5 110,200 175,100 125,900 140,500 152,900 140,920
437.8 403,100 285,600 605,800 367,500 702,600 472,920
398.1 321,580 715,600 456,800 555,200 985,600 606,956
318.4 5,569,800 6,771,100 7,865,200 6.735 × 106

3.3. S-N Curve Test of Notched Specimens


S-N curve tests of three different notched specimens were conducted on a Zwick
HB250 fatigue testing machine Figure 12. Based on the actual working condition of the
pumphead, axial tensile sinusoidal waveform loading with R = −1 was selected to control
the testing machine. The loading conditions and fatigue experimental results of the notched
specimens are presented in Table 5.

Table 5. Cyclic loading parameters and results of fatigue test for 40CrNi2 MoV notched specimens.

Specimen No. kt F (kN) Frequency (Hz) Nf (Cycles) Rσ Median Life


1-1 2 20 kN 10.0 62,724 −1
1-2 2 20 kN 10.0 68,563 −1 65,215
1-3 2 20 kN 10.0 64,359 −1
2-1 2 19 kN 10.0 122,133 −1
2-2 2 19 kN 10.0 134,572 −1 134,011
2-3 2 19 kN 10.0 145,328 −1
3-1 2 17 kN 10.0 363,235 −1
3-2 2 17 kN 10.0 384,527 −1 390,452
3-3 2 17 kN 10.0 423,596 −1
4-1 3 10 kN 10.0 6680 −1
4-2 3 10 kN 10.0 6899 −1 6949
4-3 3 10 kN 10.0 7268 −1
5-1 3 7 kN 10.0 85,362 −1
5-2 3 7 kN 10.0 81,237 −1 82,295
5-3 3 7 kN 10.0 80,198 −1
6-1 3 5.5 kN 10.0 638,340 −1
6-2 3 5.5 kN 10.0 534,568 −1 576,630
6-3 3 5.5 kN 10.0 556,982 −1
7-1 5 8 kN 10.0 36,024 −1
7-2 5 8 kN 10.0 45,396 −1 37,226
7-3 5 8 kN 10.0 30,258 −1

253
Materials 2022, 15, 4413

Table 5. Cont.

Specimen No. kt F (kN) Frequency (Hz) Nf (Cycles) Rσ Median Life


8-1 5 7.5 kN 10.0 104,417 −1
8-2 5 7.5 kN 10.0 109,853 −1 109,131
8-3 5 7.5 kN 10.0 110,123 −1
9-1 5 6 kN 10.0 190,823 −1
9-2 5 6 kN 10.0 198,095 −1 192,826
9-3 5 6 kN 10.0 189,562 −1

4. Results and Discussion


4.1. Fatigue Life Verification of Notched Specimens
In Section 2, the theoretical SFI approach and the numerical simulation method were
introduced. According to experimental loading conditions, three different stress levels were
set for each notched specimen prior to the elastic–plastic finite element analysis. The static
structural module of ANSYS workbench software was used for numerical calculations. To
ensure the accuracy of FEA results, meshes at the notch root of the specimens were refined.
In Section 2.4, two different geometrical shapes of the fatigue failure region (crescent and
cardioid) were proposed; thus, the improved SFI approach was also divided into two
models (Equations (12) and (13)). Extracting stress distribution data along the symmetry
line (focus path) of the notched specimens, stress distributions and stress-field intensities
were calculated by the novel SFI approach (Figure 14). The stress gradient decreased
rapidly in highly stressed volumes, and the “inhibition” gradient presented a similar trend
(Figure 14a). However, with the increase of the field diameter, the “inhibition” gradient
gradually approached zero and the stress-field intensity gradually changed to a constant
value. The same trend was noticed for the other loading cases, and the corresponding
results are presented in Figure 14b and Table 6.

Figure 14. Analysis of stress fields near the notch by the simplified and improved SFI approach:
(a) loading result for kt = 2 and σn = 20kN, and (b) theoretical calculation results for 9 different cases.

254
Materials 2022, 15, 4413

Table 6. Theoretical results of fatigue life for three notched specimens, calculated by four approaches.

Novel SFI Novel SFI Traditional Traditional


Stress Approach I Approach II SFI Approach TCD (LM)
Median
Concentration F/kN
Life σFI / Nf / σFI / Nf / σFI / Nf / σmean / Nf /
Factor
MPa Cycles MPa Cycles MPa Cycles MPa Cycles
20 kN 65,215 502.21 71,072 510.68 63,525 558.3 30,468 572.3 24,911
kt = 2 19 kN 134,011 455.36 157,695 466.16 132,202 509.5 64,033 529.5 46,846
17 kN 390,452 407.71 397,452 406.18 404,984 453.2 166,555 482.6 100,499
10 kN 6949 725.55 3649 706.44 4425 762.3 2428 809.2 1497
kt = 3 7 kN 82,295 508.94 66,667 495.75 80,965 561.3 29,164 592.4 18,966
5.5 kN 576,630 398.06 476,335 389.29 573,213 452.1 157,535 473.8 117,128
8 kN 37,226 540.68 39,948 521.21 53,424 596.7 17,747 611.5 14,545
kt = 5 7.5 kN 109,131 478.02 107,568 469.18 125,477 530.4 46,194 562.9 28,506
6 kN 192,826 425.79 279,319 416.39 332,268 472.9 117,319 498.6 77,084

Based on the aforementioned principle, two other approaches (traditional SFI approach
and TCD method) were chosen to calculate the fatigue life of the notched specimens. In
comparison to the traditional SFI approach and the TCD method, the novel SFI approach
had the narrowest error band (Figure 14), indicating its higher fatigue life prediction
accuracy. The error mean absolute percent index parameter (μ) was expressed as:

 
1 n  NP − NE 
μ = ∑
ni = 1  NE  × 100% (19)

where NP is the predicted life and NE is the experimental life. The index parameters of the
notched specimens are presented in Figure 15.
The fatigue life prediction accuracies of different approaches were quantitatively
analyzed. The novel SFI approach II had the highest fatigue life prediction accuracy for
the kt = 2 notched specimen. For the kt = 2 notched specimen, the deviations between the
novel SFI approach II-based calculation results and the experimental fatigue life values
under 20, 19, and 17 kN were 2.59%, 1.35%, and 3.72%, respectively. For the kt = 3 notched
specimen, the deviations between the novel SFI approach II-based calculation results
and the experimental fatigue life values under 10, 7, and 5.5 kN were 36.2%, 1.61%, and
0.59%, respectively. For the kt = 5 notched specimen, the deviations between the novel
SFI approach II-based calculation results and the experimental fatigue life values under 8,
7.5, and 6 kN were 43.5%, 14.9%, and 72.3%, respectively. Similarly, for the kt = 2 notched
specimen, the deviations between the novel SFI approach II-based calculation results
and the experimental fatigue life values under 20, 19, and 17 kN were 8.9%, 17.6%, and
1.79%, respectively. For the kt = 3 notched specimen, the deviations between the novel SFI
approach II-based calculation results and the experimental fatigue life values under 10, 7,
and 5.5 kN were 47.4%, 18.9%, and 17.3%, respectively. For the kt = 5 notched specimen, the
deviations between the novel SFI approach II-based calculation results and the experimental
fatigue life values under 8, 7.5, and 6 kN were 7.3%, 1.43%, and 44.8%, respectively.
Therefore, when the stress concentration factors were 2 and 3, the SFI approach II was
more accurate in predicting fatigue life; however, when the stress concentration factor was
5, the SFI approach I had better fatigue life prediction efficiency than the SFI approach II. In
terms of the absolute percent index parameter (μ), the SFI approach I (25.76%) was more
accurate than the SFI approach II; thus, the SFI approach I was more suitable for the fatigue
life prediction of sharp notched specimens.

255
Materials 2022, 15, 4413

Figure 15. Index parameters of notched specimens calculated by four different approaches.

The deviations between fatigue life values calculated by the traditional SFI approach
and the TCD method under different loading conditions were approximately 70%. There-
fore, both the novel SFI approaches I and II are highly accurate in predicting high-cycle
fatigue life.

4.2. Fatigue Life Verification of Pumphead


To further verify the fatigue life prediction accuracy of the novel SFI approach for the
pumphead, numerical simulations for stress distribution and fatigue life were performed
in ANSYS workbench software (Figure 16).
For extracting stress distribution data along the symmetry line (focus path) of the
pumphead cross-bore, field intensities under four different working conditions were calcu-
lated by the novel SFI approach II. The gradient value of “inhibition action” was calculated
as shown in Figure 17. As shown in Figure 17, the slope of the “inhibition” effect (dΩ/dr)
is close to 0 in almost all the four working conditions when R is close to 1.5 mm. With
the increase of the field diameter, the value of field strength does not alter and steadily
becomes constant. It can be judged that the damage in this area can be defined as the
effective damage area of the ultra-high-pressure pumphead body. The field intensity under
65, 70, 80, and 105 MPa was 318, 340, 389, and 518 MPa, respectively. Further, substituting
these four field strengths into the S-N curve of the initial pumphead material, the fatigue
life was calculated by the new method.

256
Materials 2022, 15, 4413

Figure 16. (a) Analysis model of the hydraulic part of the fracturing pump, (b) 1/2 inner cavity of
the pumphead, (c) typical mesh structure, (d) boundary conditions for numerical simulations, and
(e) maximum stress contour and focus path.

Figure 17. Simplified novel SFI method to analyze the stress field near the concentration point of the
intersecting line under the four load conditions: (a) 65 MPa, (b) 70 MPa, (c) 80 MPa, and (d) 105 MPa.

257
Materials 2022, 15, 4413

According to the actual working conditions of the fracturing pump, the loading
frequency of the circular cross-bore was 110 cycles/min. Owing to the component size,
material differences, and loading distribution, even under the same operating conditions,
the fatigue life of pressure vessels becomes dispersed; thus, fatigue reliability methods
are found to be better for data processing [45,46]. To further verify the new method
proposed in this paper, the nominal stress method commonly used in engineering was
used for comparison. The mathematical definition of the nominal stress method is given in
Equation (20):
σa
σn = εβCL (20)
KT
where σn is the stress of the mechanical structural part in the structural S-N curve, σa is the
stress of material in the S-N curve, KT is the stress concentration factor, ε is the size factor, β
is the surface processing coefficient, and CL is the loading coefficient.
In this paper, the value of the size factor, ε, is 0.9, because the specimens are taken from
the same nickel-chromium alloy material, so the value of the surface processing coefficient,
β, is 1. The loading method of the fracturing pump is the same as in the case of the fatigue
test, so the value of the loading coefficient is 1. In addition, based on Equations (18) and
(20) and the S-N curve, the fatigue life prediction equation can be expressed as:

(4.97σn )8.1225 · N = 6.24 × 1026 (21)

Therefore, in 65, 70, 80, and 105 MPa working conditions, combined with the S-N
curves, fatigue life is calculated by the SFI approach and NSM, respectively. The fatigue life
values of the pumphead under different conditions are displayed in Figure 18. It is evident
that the novel SFI approach had the narrowest error band, indicating its higher fatigue life
prediction accuracy for the pumphead.

Figure 18. Pumphead fatigue life prediction accuracies of NSM and the novel SFI approach.

5. Conclusions
Based on the theories of the critical distance method and the SFI approach, criteria
for defining the fatigue damage failure region of the SFI approach were revisited. The

258
Materials 2022, 15, 4413

proposed novel SFI approach manifested high fatigue life prediction accuracy for notch
specimens and pumpheads. The main findings of this work are summarized below:
(1) Based on numerical simulation results and the concept of the critical volume
method, an improved weight function accounting for the fatigue damage failure geometries
of cardioid and crescent areas was proposed to improve fatigue life prediction accuracy.
(2) Based on the concept of the “invisible boundary”, the effective critical distance for
crescent and cardioid areas of the fatigue failure region was revised and simplified. The
novel SFI approach was modified considering the effect of grain size.
(3) To verify the fatigue life prediction effectiveness of the novel SFI approaches for
notch specimens, fatigue tests were conducted on three different notched bar specimens.
The novel SFI approach II had higher fatigue life prediction accuracy.
(4) In comparison with the SFI approach II, the SFI approach I was more suitable for
fatigue life prediction of sharply notched specimens.

Nomenclatures
σmax Maximum stress kt Stress concentration factor
σFI Stress-field intensity kf Fatigue notch factor
Distance to the peak
r χ Stress gradient
stresspoint
f (σij ) Equivalent stress function χ Relative stress gradient

ϕ( r ) Weight function σ0 Nominal stress
Weight function with
ϕ(r, θ ) distance r and deviation εa Strain amplitude
angle θ
σa Stress amplitude R Field diameter
The load of component
σn σU Ultimate stress
weakest region
σb Ultimate tensile strength σmean Effective stress
Range of threshold value
Range of the fully reverse
ΔKth for fatigue crack Δσ−1
axial fatigue limit
propagation
The maximum principal
σmean Equivalent stress of TCD σ1
stress
SY Yield stress G Grain size function
Angular coordinates in
ξ (Ω) Obstructive effect θ
Cartesian coordinates
Fatigue crack-opening
σyy (r) L0 Critical distance
stress
ν Poisson’s ratio

Abbreviations
Theory of critical distance
SFI Stress-field intensity TCD
method
Linear elastic fracture Elastic–plastic fracture
LFEM EPFM
mechanics mechanics
NSM Nominal stress method LSSM Local stress–strain method

Author Contributions: The author H.W. provided help in the preliminary investigation of this article.
N.L. provided resources such as experimental equipment. M.L. Completed the numerical simulation.
Y.Z. (Yang Zhou) conducted in-depth research on the evaluation the fatigue life of pump heads, then
proposed the concept of fatigue failure region. Y.Z. (Yun Zeng) was in charge of the entire research
experiment, statistics all the data and sorted it out. Finally wrote and revised the manuscript. All
authors have read and agreed to the published version of the manuscript.
Funding: This research received no external funding.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: All data are presented in the article.
Acknowledgments: This project was supported financially by the National Natural Science Founda-
tion of China (No. 52174018 and No. 51904181), CNPC Innovation Foundation 2020D-5007-0503, the
Project of case Teaching for graduate degree of Yangtze University (No. YAL202106, Application of
engineering software analysis), and State Key Science and Technology Projects (No. 2016ZX05038-
001-LH002). The authors would like to thank all the members of the project team for their support.

259
Materials 2022, 15, 4413

Conflicts of Interest: We declare that we have no financial and personal relationships with other
people or organizations that can inappropriately influence our work, there is no professional or
other personal interest of any nature or kind in any product, service and/or company that could
be construed as influencing the position presented in, or the review of, the manuscript entitled,
“Experiment and Theoretical Investigation on Fatigue Life Prediction of Fracturing Pumpheads Based
on a Novel Stress-Field Intensity Approach”.

References
1. Zhu, W.B. Research on Reciprocating Sealing Performance and Mechanism of Fracturing Pump. Nat. Gas Industry 2006, 60–62,
161–162.
2. Bokane, A.; Stewart, M.; Sanda, R.; Babar, D.; Katke, N. Personal Safety and Human Factors Risk Assessment during Maintenance
of Fracturing Pumps at Well Service Operations. In Proceedings of the International Petroleum Technology Conference, Kuala
Lumpur, Malaysia, 10–12 December 2014.
3. Badr, E.A.; Sorem, J.R.; Tipton, S.M.; Yang, S. An analytical procedure for estimating residual stresses in blocks containing
crossbores. Int. J. Press. Vessel. Pip. 2000, 77, 737–749. [CrossRef]
4. Zanetti, M.; Babini, V.; Meneghetti, G. Definition of nominal stress-based FAT classes of complex welded steel structures using the
Peak Stress Method. Procedia Struct. Integr. 2019, 19, 627–636. [CrossRef]
5. Fomichev, P.A.; Zarutskii, A.V. Fatigue Life Prediction for Specimens with an Open Hole with a Pre-Compressed Boundary Via
the Nominal Stresses Under Regular Cyclic Loading. Strength Mater. 2019, 51, 746–752. [CrossRef]
6. Zhu, S.-P.; He, J.-C.; Liao, D.; Wang, Q.; Liu, Y. The effect of notch size on critical distance and fatigue life predictions. Mater. Des.
2020, 196, 109095. [CrossRef]
7. He, J.-C.; Zhu, S.-P.; Liao, D. Probabilistic fatigue assessment of notched components under size effect using critical distance
theory. Eng. Fract. Mech. 2020, 235, 107150. [CrossRef]
8. Liao, D.; Zhu, S.-P.; Correia, J.A.; De Jesus, A.M.; Berto, F. Recent advances on notch effects in metal fatigue: A review. Fatigue
Fract. Eng. Mater. Struct. 2020, 43, 637–659. [CrossRef]
9. Susmel, L.; Taylor, D. The Theory of Critical Distances to estimate lifetime of notched components subjected to variable amplitude
uniaxial fatigue loading. Int. J. Fatigue 2011, 33, 900–911. [CrossRef]
10. Adib, H.; Gilgert, J.; Pluvinage, G. Fatigue life duration prediction for welded spots by volumetric method. Int. J. Fatigue 2004, 26,
81–94. [CrossRef]
11. Esmaeili, F.; Zehsaz, M.; Chakherlou, T.N. Investigation the effect of tightening torque on the fatigue strength of double lap
simple bolted and hybrid (bolted–bonded) joints using volumetric method. Mater. Des. 2014, 63, 349–359. [CrossRef]
12. Sobek, J. Proper Nominal Stress Distribution Subjected to Combination of Wedge-Splitting and Bended Geometry Tests. Procedia
Eng. 2017, 190, 406–413. [CrossRef]
13. Ostash, O.P.; Panasyuk, V.V. Fatigue process zone at notches. Int. J. Fatigue 2001, 23, 627–636. [CrossRef]
14. Susmel, L.; Taylor, D. An elasto-plastic reformulation of the theory of critical distances to estimate lifetime of notched components
failing in the low/medium-cycle fatigue regime. J. Eng. Mater. Technol. 2010, 132, 1–8. [CrossRef]
15. Susmel, L.; Meneghetti, G.; Atzori, B. A simple and efficient reformulation of the classical manson–coffin curve to predict lifetime
under multiaxial fatigue loading-Part II: Notches. J. Eng. Mater. Technol. 2009, 131, 1–9. [CrossRef]
16. Tanaka, K. Engineering formulae for fatigue strength reduction due to crack-like notches. Int. J. Fract. 1983, 22, 39–46. [CrossRef]
17. Taylor, D. Analysis of fatigue failures in components using the theory of critical distances. Eng. Fail. Anal. 2005, 12, 906–914.
[CrossRef]
18. El Haddad, M.H.; Dowling, N.E.; Topper, T.H.; Smith, K.N. J integral applications for short fatigue cracks at notches. Int. J. Fract.
1980, 16, 15–30. [CrossRef]
19. Susmel, L.; Taylor, D. A novel formulation of the theory of critical distances to estimate lifetime of notched components in the
medium-cycle fatigue regime. Fatigue Fract. Eng. Mater. Struct. 2007, 30, 567–581. [CrossRef]
20. Susmel, L.; Taylor, D. The Modified Wöhler Curve Method applied along with the Theory of Critical Distances to estimate finite
life of notched components subjected to complex multiaxial loading paths. Fatigue Fract. Eng. Mater. Struct. 2008, 31, 1047–1064.
[CrossRef]
21. Susmel, L.; Taylor, D. Estimating lifetime of notched components subjected to variable amplitude fatigue loading according to the
elastoplastic theory of criticaldistances. J. Eng. Mater. Technol. 2015, 137, 1–15. [CrossRef]
22. Qylafku, G.; Azari, Z.; Kadi, N.; Gjonaj, M.; Pluvinage, G. Application of a new model proposal for fatigue life prediction on
notches and key-seats. Int. J. Fatigue 1999, 21, 753–760. [CrossRef]
23. Qilafku, G.; Kadi, N.; Dobranski, J.; Azari, Z.; Gjonaj, M.; Pluvinage, G. Fatigue of specimens subjected to combined loading: Role
of hydrostatic pressure. Int. J. Fatigue 2001, 23, 689–701. [CrossRef]
24. Adib-Ramezani, H.; Jeong, J.; Pluvinage, G. Structural integrity evaluation of X52 gas pipes subjected to external corrosion defects
using the SINTAP procedure. Int. J. Press. Vessel. Pip. 2006, 83, 420–432. [CrossRef]
25. Meneghetti, G.; Susmel, L.; Tovo, R. High-cycle fatigue crack paths in specimens having different stress concentration features.
Eng. Fail. Anal. 2007, 14, 656–672. [CrossRef]

260
Materials 2022, 15, 4413

26. Ostash, O.P.; Panasyuk, V.V.; Kostyka, E.M. Phenomenological model of fatigue macrocrack initiation near stress concentrators.
Fatigue Fract. Engng. Mater. Struct. 1999, 22, 161–172. [CrossRef]
27. Nicholls, D.J.; Martin, J.W. An examination of the reasons for the discrepancy be-tween long and small fatigue cracks in Al-Li
alloys. Mat. Sci. Eng. A 1990, 128, 141–145. [CrossRef]
28. Pineau, A.G.; Pelloux, R.M. Influence of strain-induced martensitic transformations on fatigue crack growth rates in stainless
steels. Met. Trans. 1974, 5, 1103–1112. [CrossRef]
29. Saxena, A.; Antolovich, S.D. Low cycle fatigue, fatigue crack propagation and substructures in a series of polycrystalline cu-Al
alloys. Met. Trans. 1975, 6, 1809–1828. [CrossRef]
30. Park, H.B.; Kim, K.M.; Lee, B.W. Plastic zone size in fatigue cracking. Int. J. Press. Vessel. Pip. 1996, 68, 279–285. [CrossRef]
31. Bathias, C.; Pelloux, R.M. Fatigue crack propagation in martensitic and austenitic steels. Met. Trans. 1973, 4, 1265–1273. [CrossRef]
32. Chapetti, M.D.; Miyatab, H.; Tagawab, T.; Miyatab, T.; Fujiokac, M. Fatigue crack propagation behavior in ultra-fine grained low
carbon steel. Int. J. Fatigue 2005, 27, 235–243. [CrossRef]
33. Edmunds, T.M.; Willis, J.R. Matched asymptotic expansions in nonlinear fracture mechanics—III. In-plane loading of an elastic
perfectly-plastic symmetric specimen. J. Mech. Phys. Solids 1977, 25, 423–455. [CrossRef]
34. Taddesse, A.T.; Zhu, S.-P.; Liao, D.; Huang, H.-Z. Cyclic plastic zone modified critical distance theory for notch fatigue analysis of
metals. Eng. Fail. Anal. 2021, 121, 105163. [CrossRef]
35. Taddesse, A.T.; Zhu, S.-P.; Liao, D.; Keshtegar, B. Cyclic plastic zone-based notch analysis and damage evolution model for fatigue
life prediction of metals. Mater. Des. 2020, 191, 108639. [CrossRef]
36. Wu, Y.-L.; Zhu, S.-P. Assessment of notch fatigue and size effect using stress field intensity approach. Int. J. Fatigue 2021, 149,
106279. [CrossRef]
37. Ai, Y.; Zhu, S.-P.; Liao, D. Probabilistic modelling of notch fatigue and size effect of components using highly stressed volume
approach. Int. J. Fatigue 2019, 127, 110–119. [CrossRef]
38. Liao, D.; Zhu, S.-P.; Keshtegar, B.; Qian, G.; Wang, Q. Probabilistic framework for fatigue life assessment of notched components
under size effects. Int. J. Mech. Sci. 2020, 181, 105685. [CrossRef]
39. Yao, W. Stress field intensity approach for predicting fatigue life. Int. J. Fatigue 1993, 15, 243–245.
40. Neuber, H. Theory of Notch Stress: Principles for Exact Calculation of Strength with Reference to Structural Form and Material, 2nd ed.;
Springer: Berlin/Heidelberg, Germany, 1958.
41. Peterson, R.E.; Sines, G.; Waisman, J.L. (Eds.) Notch Sensitivity; Metal Fatigue; McGraw Hill: New York, NY, USA, 1959;
pp. 293–306.
42. Berto, F.; Lazzarin, P. The volume-based Strain Energy Density approach applied to static and fatigue strength assessments of
notched and welded structures. Procedia Eng. 2009, 1, 155–158. [CrossRef]
43. Zhao, B.; Song, J.; Xie, L.; Hu, Z.; Chen, J. Surface roughness effect on fatigue strength of aluminum alloy using revised stress field
intensity approach. Sci. Rep. 2021, 11, 19279. [CrossRef]
44. Zeng, Y.; Li, M.; Zhou, Y.; Li, N. Development of a new method for estimating the fatigue life of notched specimens based on
stress field intensity. Theor. Appl. Fract. Mech. 2019, 104, 102339. [CrossRef]
45. Niu, X.-P.; Zhu, S.-P.; He, J.-C.; Shi, K.; Zhang, L. Fatigue reliability design and assessment of reactor pressure vessel structures:
Concepts and validation. Int. J. Fatigue 2021, 153, 106524. [CrossRef]
46. Li, X.-K.; Zhu, S.-P.; Liao, D.; Correia, J.A.F.O.; Berto, F.; Wang, Q. Probabilistic fatigue modelling of metallic materials under
notch and size effect using the weakest link theory. Int. J. Fatigue 2022, 159, 106788. [CrossRef]
47. Zhao, B.; Xie, L.; Song, J.; Ren, J.; Wang, B.; Zhang, S. Fatigue life prediction of aero-engine compressor disk based on a new stress
field intensity approach. Int. J. Mech. Sci. 2020, 165, 105190. [CrossRef]
48. Zhao, H.; Liu, J.; Hua, F.; Ran, Y.; Zi, R.; Li, B. Multiaxial fatigue life prediction model considering stress gradient and size effect.
Int. J. Press. Vessel. Pip. 2022, 199, 104703. [CrossRef]

261
materials
Article
Further Investigations and Parametric Analysis of
Microstructural Alterations under Rolling Contact Fatigue
Muhammad Usman Abdullah and Zulfiqar Ahmad Khan *

NanoCorr, Energy & Modelling (NCEM) Research Group, Department of Design & Engineering,
Bournemouth University, Poole BH12 5BB, UK
* Correspondence: [email protected]

Abstract: Bearing elements under rolling contact fatigue (RCF) exhibit microstructural features,
known as white etching bands (WEBs) and dark etching regions (DERs). The formation mecha-
nism of these microstructural features has been questionable and therefore warranted this study
to gain further understanding. Current research describes mechanistic investigations of standard
AISI 52100 bearing steel balls subjected to RCF testing under tempering conditions. Subsurface
analyses of RCF-tested samples at tempering conditions have indicated that the microstructural
alterations are progressed with subsurface yielding and primarily dominated by thermal tempering.
Furthermore, bearing balls are subjected to static load tests in order to evaluate the effect of lattice
deformation. It is suggested from the comparative analyses that a complete rolling sequence with
non-proportional stress history is essential for the initiation and progression of WEBs, supported
by the combination of carbon flux, assisted by dislocation and thermally activated carbon diffusion.
These novel findings will lead to developing a contemporary and new-fangled prognostic model
applied to microstructural alterations.

Citation: Abdullah, M.U.; Khan, Z.A. Keywords: bearing steel; diffusion; microstructural alterations; rolling contact fatigue; tempering
Further Investigations and
Parametric Analysis of
Microstructural Alterations under
Rolling Contact Fatigue. Materials
1. Introduction
2022, 15, 8072. https://doi.org/
10.3390/ma15228072
Automotive components subjected to severe rolling cyclic loadings not only result
in deterioration of interacting surfaces [1,2] but also cause subsurface microstructural
Academic Editors: Jaroslaw damage [3,4]. This subsurface damage is pronounced in bearing steel under rolling con-
Galkiewicz, Lucjan Śnieżek
tact fatigue (RCF) where the parent microstructure is evolved resulting in formation of
and Sebastian Lipiec
microstructural alterations. The subsurface microstructural alterations can be observed
Received: 2 October 2022 under the optical microscope after Nital etching as a ring profile or semi-elliptical shape,
Accepted: 11 November 2022 depending on the plane of cut. The primary features involved in the altered microstructure
Published: 15 November 2022 are white etching bands (WEBs) preceded by dark etching regions (DERs) [4,5]. It has been
reported that the microstructural features are formed due to the decay of parent marten-
Publisher’s Note: MDPI stays neutral
site where carbon migrates within the matrix to form carbon-rich and carbon-depleted
with regard to jurisdictional claims in
published maps and institutional affil-
areas [6,7]. Extensive numerical and experimental research is available in the literature
iations.
on bearing steel evolution and microstructural features; however, a comprehensive un-
derstanding is still needed to comprehend the general (physical and chemical) formation
mechanism of DERs/WEBs.
DERs are considered to be an over-tempered form of martensite constituting a mixture
Copyright: © 2022 by the authors. of decayed microstructure, ferrite, and remaining martensite [5]. DERs are governed
Licensee MDPI, Basel, Switzerland. by the subsurface stress fields and their structure consists of deformed and misoriented
This article is an open access article martensite plates including reprecipitation of nano carbides, resulting in over-tempering
distributed under the terms and of the microstructure [8]. WEBs are reported as disc-shaped ferrite plates surrounded by
conditions of the Creative Commons carbon-rich areas (also known as lenticular carbides, LCs) [9]. The mechanism by which
Attribution (CC BY) license (https:// DERs/WEBs are formed has been debated for the last few decades and redistribution of
creativecommons.org/licenses/by/ carbon within the white bands has been reported in the literature [7]. Earlier research [10,11]
4.0/).

Materials 2022, 15, 8072. https://doi.org/10.3390/ma15228072 263 https://www.mdpi.com/journal/materials


Materials 2022, 15, 8072

has indicated that WEBs are caused by stress and are developed as a result of subsurface
plastic deformation. The structural changes observed during RCF have been attributed
to recrystallisation, in which carbon atoms diffuse from supersaturated ferrite regions to
create LCs. Carbon diffuses from ferrite interstitially to LCs and requires instigation energy
to break their binding energies. This instigation energy has been reported as dissipated
plastic energy during cyclic loadings [12–14]. Another theory [15] proposed that the carbon
depletion within ferrite bands is caused by dislocation annihilation. Recently, Fu et al. [16]
proposed a carbon migration theory based on dislocation gliding which indicated that
carbon diffuses from ferrite bands to thicken neighbouring Lenticular Carbides. However,
a bulk of the research has been centred on the migration of carbon within the ferrite matrix
and could be restrained by either thermally driven diffusion or gliding dislocations [17]. A
clear understanding of band formation is still required to resolve the ambiguity regarding
white band formation.
Current research presents a systematic approach to understanding the development
and formation mechanism of DERs/WEBs during RCF. Standard martensitic AISI 52100
bearings balls are tempered in a muffle furnace. The tempered and untempered ball samples
are then subjected to specific RCF conditions in a rotary tribometer. Additionally, the RCF
tests are performed and then subjected to similar tempering conditions. It is observed
that the bearing pre-tempered RCF-tested ball samples showed significant microstructural
alterations under the contact track as compared with untempered RCF-tested samples. The
initially conducted RCF test exhibited the formation of DERs accompanying WEBs. After
tempering, little to no difference in the WEB thickening is observed, however, randomly
distributed white patch formation in the vicinity of DERs can be seen. Besides tempering,
the static loading analysis is conducted to understand the effect of non-proportional loading
histories. The bearing ball samples are subjected to the static load RCF test to evaluate
the effect of lattice widening at extreme loading conditions. The mechanistic approach
employed in current research is highly significant in evaluating RCF-induced subsurface
alterations and can be employed for improving efficiency of analytical modelling applied
in DER and WEB formation.

2. Experimental Setup
RCF tests were conducted with the help of a four-ball configuration in a high-speed
PLINT rotary tribometer, where three lower balls were kept (free to roll) in a cup which
worked like an outer race of a bearing. In a four-ball test configuration, the upper ball
(test sample) rotates on top of lower three balls and experiences 2.25 stress cycles [18]. A
synthetic industrial ISO-Vg-320 oil was used in bath lubrication in order to avoid asperities
interaction. Standard martensitic AISI 52100 (also known as 100Cr6) bearing balls of
12.7 mm diameter were used as test samples. Once the test was suspended, the upper ball
in the collet was demounted, conditioned in an ultrasonic conditioning ultrasonic bath, and
then sectioned into extremely thin slices in both axial and circumferential directions. A cubic
boron nitride (CBN) cutter in a precision sectioning machine was employed with extensive
cooling water to avoid the machining effects on the sample surface. After machining, the
samples were cleaned, preserved via cold mounting, and polished with a 1 μm diamond
suspension. Figure 1 demonstrates the sample preparation and resulting micrographs
after Nital etching. Further details on the RCF experimentation, four-ball test setup, and
respective sample preparation can be found in the authors’ previous research [18,19]. The
acquired subsurface micrographs reveal the microstructure features as a semi-elliptical or
as a ring-shaped geometrical entity via bright field microscopy after etching with Nital
solution (3% nitric acid in ethanol).

264
Materials 2022, 15, 8072

Figure 1. Rolling contact fatigue testing in a rotary tribometer.

3. Tempering Analysis
Standard AISI 52100 bearing ball samples were tested at 6 GPa peak contact pressure
with 100 ◦ C operational temperature, and 4051 rpm speed. Previous research reported [18]
that the formation of microstructure features, especially the Lower Angle Bands (LABs),
starts to form after 37 million cycles at 6 GPa and 100 ◦ C. Therefore, it can be suggested
that the initiating point for white bands formation exists near 37 million rolling cycles.
The initial plastic shakedown of bearing elements during rolling cyclic fatigue is believed
to occur in the very early stage of RCF, usually about 0.1% of total bearing life cycle [20].
Therefore, the bearing ball samples were RCF tested till 37 million and 0.02 million rolling
cycles. The full details of conducted tests for tempering analysis are mentioned in Table 1.
Figure 2 shows the optical micrographs of RCF-tested samples in axial and circumferential
orientations. It can be seen that the bearing sample exhibits formation of LABs, nearly
10–20 μm in length, along with the formation of DER as observed in the circumferential
orientation; refer to Figure 2a.

Table 1. Experimentation conducted for tempering analysis.

Max Contact Lubricant


Sample Spindle Speed Elapsed Cycles Tempering Condition
Pressure Temperature
1 6 GPa 100 ◦ C 4051 rpm 37 million 380 ◦ C for 240 min after RCF test
2 6 GPa 100 ◦ C 500 rpm 0.02 million 380 ◦ C for 240 min after RCF test
3 4 GPa 100 ◦ C 4051 rpm 37 million No heat treatment
4 4 GPa 100 ◦ C 4051 rpm 37 million 380 ◦ C for 240 min before RCF test
5 5 GPa 100 ◦ C 4051 rpm 37 million No heat treatment
6 5 GPa 100 ◦ C 4051 rpm 37 million 380 ◦ C for 240 min before RCF test

265
Materials 2022, 15, 8072

Figure 2. (a) Subsurface micrographs obtained at 6 GPa, 100 ◦ C, after 37 million rolling cycles.
(b) Tempering of RCF-tested sample shown in (a). (c) Micrographs obtained at 6 GPa, 100 ◦ C, after
0.02 × 106 rolling cycles. (d) Tempering of RCF-tested sample shown in (c). ROI:1 and ROI:2 represent
regions of interest for parent structure and RCF damaged area respectively.

To understand the effects of thermal diffusion, the RCF-tested samples (1 and 2) were
tempered in a muffle furnace at a temperature of 380 ◦ C for 240 min. Standard martensitic
bearing steel at such elevated temperatures is expected to undergo thermal tempering

266
Materials 2022, 15, 8072

resulting in more carbon atoms available in the solute leading to enhanced formation of
microstructural alterations. After tempering in a muffle furnace, the same sample was
again polished and etched to remove any oxidation layer resulting from heat treatment.
Figure 2b shows the formation of randomly distributed white patches in the vicinity of
DERs. This white etching matter can be characterised as WEB. The inset of Figure 2a,b
shows no difference in WEBs formation before and after tempering of bearing samples.
The tempering temperature, as high as 380 ◦ C resulted in the reprecipitation of the free
carbon in the ferrite matrix which was segregated during the martensite–ferrite phase
transformation. Additionally, the tempering of the RCF-tested sample also caused the
overall reduction of DER which can only be seen in the upper and lower boundaries of
white etching matter. However, no thickening of LABs or formation of higher angle bands
(HABs) can be observed after heat treatment.
Figure 2c represents RCF investigated sample at 6 GPa maximum Hertzian contact
pressure with 100 ◦ C operational temperature till 0.02 million rolling cycles. In order to
achieve such low stress cycles in a high-speed rotary tribometer, the spindle speed was
reduced from 4051 rpm to 500 rpm. It should be noted that no microstructure features have
been observed under such conditions as seen in the optical micrograph. It is difficult to
justify the shakedown limit of certain bearing steel, however, it is speculated that under
severe operating conditions, i.e., 6 GPa contact pressure, substantial plastic deformation will
occur, which in turn will lead to a high dislocation density after several thousand rolling
cycles. A similar sample was tempered to investigate the influence of carbon diffusion
within areas of high dislocation density. It can be observed in Figure 2d that the heat
treatment of RCF-tested samples does not influence the formation of DER which is a result
of an over-tempered form of martensite, as suggested previously [21,22]. The regions of
interest (ROI:1 and ROI:2) from Figure 2c are also displayed for parent martensitic structure
and RCF-induced damaged area. It is evident from scanning electron micrographs that
the original needle-like microstructure has been lost during the early shakedown stage
of rolling fatigue cycles. The tempered martensite exists in the form of high dislocation
regions due to subsurface plastic deformation.
The bearing ball samples were also tempered prior to RCF testing. For heat treatment,
the ball samples were initially heat treated in a muffle furnace at 380 ◦ C temperature for
240 min which resulted in a hardness drop of nearly 250 HV. The baseline hardness of the
through-hardened bearing samples decreased from 782–810 HV (10 Kgf) to 525–560 HV
(10 Kgf) after the heat treatment process. The tempered and untempered bearing ball
samples were then subjected to rolling cyclic loading in the rotary tribometer discretely
at 4 GPa and 5 GPa maximum contact pressure with 100 ◦ C working temperature and a
constant cyclic frequency of 4051 rpm.
It can be observed in Figure 3a,b that the bearing ball sample, after RCF testing at 4 GPa
contact pressure, shows little to no formation of microstructural alterations, whereas the pre-
tempered RCF test exhibits the development of microstructural features under the contact
track. Correspondingly, the untempered bearing ball tested at 5 GPa contact pressure
shows formation of DER in a very localised area under the centreline contact track, refer
to Figure 3c, whereas the pre-tempered RCF-tested sample exhibits enhanced formation
of white bands which can be observed in axial orientation as well as in circumferential
orientation in Figure 3d,e. Scanning electron micrographs show these elongated band
structures making steeper angles with the rolling direction (RD) and are described as
premature HABs. The formation of premature HABs before LABs has been reported in
the authors’ previous work [19] where stress as high as 6 GPa contact pressure has been
reported to be a cause of WEBs sequence reversal. The development of HABs before LABs
can be explained by considering white bands formation to be yield controlled rather than
stress controlled as suggested by previous researchers [23]. It has been suggested that a
threshold yield limit exists, beyond which the formation of white bands takes place in a
preferable orientation, which is steeper to the RD.

267
Materials 2022, 15, 8072

ȱ
Figure 3. (a) Untempered and (b) tempered RCF-tested sample at 4 GPa, 100 ◦ C, 4051 rpm.
(c) Untempered and (d) tempered RCF-tested samples at 5 GPa, 100 ◦ C, and 4051 rpm. The in-
set of circumferentially sliced sample (e) is shown in ROI:1.

The accelerated growth of white bands for pre-tempered RCF-tested samples suggests
a significant effect of lowering of yield stress on microstructural alterations, resulting in
subsequent subsurface plastic deformation. Moreover, the white bands are considered
to be a softer structure as compared to the parent matrix [9], which can help promote
accumulation of plastic strains. The change in yield stress and evolution of localised flow
curves during the material phase transformation is presented in Section 6.

4. Effect of Thermal Diffusion


Figure 4a represents a schematic for carbon diffusion in a 5 μm thick ferrite plate. The
initial carbon concentration of the matrix is assumed to be 6.52 × 10−6 mol/um3 as shown
and the boundaries of the plate are assumed to be carbon-free, i.e., 0% carbon concentration.
Equation (1) represents the generalised form of carbon flux J expressed as [24]

∂ϕ ∂ ∂p
J = −sD = + ks (ln(θ − θz )) + kp (1)
∂x ∂x ∂x

268
Materials 2022, 15, 8072

where D and s represent diffusivity and solubility, θ and p represent temperature and stress,
where ks and kp denote the temperature-driven and pressure-driven carbon diffusion
factors, respectively. Assuming isotropic behaviour of carbon diffusion in alpha ferrite
and neglecting the temperature and stress gradients in Equation (1), the change in carbon
concentration of ferrite bands can be obtained relative to time. The diffusivity of carbon in
α ferrite is defined in Equation (2) [25], where T represents temperature in kelvin.


2
104 104
log D = −0.9064 − 0.5199 + 1.61 × 10−3 (2)
T T

Figure 4b,c represent the estimated diffusion of carbon at 100◦ and 160 ◦ C respectively.
Complete depletion of a 5μm thick ferrite band is obtained at 160 ◦ C temperature just after
69 h. It can be noted that the initial formation of WEBs (LABs/HABs) is observed after
~69 h of RCF testing under given conditions. The enhanced formation of HABs at 160 ◦ C
operating temperature is an indication of high carbon flux due to thermally activated
carbon along with a large number of mobile dislocations. On the other hand, at 100 ◦ C
operating temperature, it takes more than 720 h (1 month) to diffuse out carbon in a 5 μm
thick ferrite plate. It can be understood that at such slow rates of carbon diffusion, the
transportation of carbon by dislocation glide becomes decisive. However, a typical run
time for bearing components during service can reach tens of thousands of hours, where
thermal diffusion of carbon should be accounted for in the formation of WEBs.

Figure 4. (a) Schematic for carbon diffusion in a 5 μm thick ferrite plate. Change in carbon concentra-
tion due to diffusion at (b) 100 ◦ C and (c) 160 ◦ C.

269
Materials 2022, 15, 8072

5. Static Load Analysis


To refute the migration of carbon as a result of dislocation gliding and understanding
the carbon diffusion process during cyclic loading, static load analyses were carried out.
The static load analysis follows a sequence where bearing balls are initially subjected to
rolling cyclic loadings in a rotary tribometer followed by a static loading condition. During
static loading, the upper ball in contact with the lower three balls was held static with
the driver motor turned off. The contact track formed on the upper ball during rolling
cyclic loading was marked precisely at the points of interactions. The time required to
load the sample under static loading conditions was estimated from the numerical analysis
discussed in Section 4. After the RCF test, the bearing ball sample was serially sectioned to
investigate the statically loaded and offloaded subsurface regions. The testing parameters
employed in the static load analysis are presented in Table 2.

Table 2. Test conducted for static load analysis.

Loading Peak Contact Operating Time for Static


Test ID Spindle Speed Stress Cycles
Condition Pressure Temperature Load
Rolling cyclic 500 rpm - 20,034
ST0.02M
Static - 69 h -
Rolling cyclic 4051 rpm - 22,001,035
ST22M-1 6 GPa 160 ◦ C
Static - 69 h -
Rolling cyclic 4051 rpm - 22,067,262
ST22M-2
Static - 100 h -

Figure 5 represents the optical micrographs obtained at 6 GPa maximum contact


pressure with 160 ◦ C operational temperature. It was suggested previously [18] that DERs
start to form at extreme testing conditions well before one million cycles. However, no
feature microstructure can be observed till 20,000 rolling cycles. Having said that, at such
extreme testing conditions, the bearing material undergoes subsurface plastic deformation
with high dislocation density, as demonstrated previously in Figure 2. The RCF-tested
bearing ball sample was then subjected to static loading conditions for 69 h, as illustrated
in Table 2. As suggested by numerical simulation of the carbon flux across a thin ferrite
plate, diffusion of carbon is highly likely to occur at 160 ◦ C temperature after 69 h. Despite
the numerical prediction and high dislocation density, no feature microstructure can be
observed in the loaded and offloaded regions of ST0.02M sample. It should be noted that
the optical micrograph for RCF testing and subsequent static loading analysis is acquired
from different bearing ball samples.
Correspondingly, the optical micrographs obtained for 6 GPa, 160 ◦ C, and 22 million
rolling cycles are represented in Figure 6 which shows the formation of early HABs in the
rolling direction. These initially developed HABs span nearly 100 μm in length and their
microstructure contains deformed shear bands as shown in the inset image. These shear
bands are formed due to the shear localisation and plastic deformation resulting in the
martensite–ferrite phase transformation. As the maximum carbon solubility in the ferrite
phase is nearly 0.02% [24], it is considered that the free carbon from such microstructural
phase transformation can segregate at the boundaries and form carbon-rich plates termed
as lenticular carbides (LCs). The transport of excess carbon atoms from the deformed
ferrite matrix towards LCs can be accelerated with the help of carbon diffusion where
thermal effects and lattice widening during static loading could lead to the thickening of
pre-existing LCs. ST22M-1 and ST22M-2 represent the static loading analysis of RCF-tested
samples till 69 h* and 100 h of testing time (* 69 h represents the testing time where initial
WEBs can be observed under given loading conditions). It can be seen that despite the
numerical prediction from Section 4, no thickening of LCs can be observed for either of the
samples despite the prolonged tempering of the RCF-tested sample under static loading.

270
Materials 2022, 15, 8072

Moreover, the static loaded regions do not show any notable difference in the growth of
LCs as compared to the offload regions when observed in the high-contrast images from
cross-polarised light microscopy.

Figure 5. Circumferentially sliced micrograph acquired at 6 GPa, 160 ◦ C after 0.02M rolling cycles.
The static loaded and offload regions are also shown for ST0.02M.

Figure 6. Circumferentially sliced micrographs obtained after 6 GPa, 160 ◦ C, 22 M rolling cycles test,
where the inset of figure shows microstructure of an early HAB. The loaded and offload regions
of ST22M-1 and ST22M-2 are also shown after static load analysis. The high-contrast images are
acquired from cross-polarised light.

271
Materials 2022, 15, 8072

The additional RCF tests were carried out at 6 GPa, 100 ◦ C till 0.75 million rolling
cycles with variation in spindle speed, i.e., 4051 rpm and 289 rpm. Figure 7a,b representing
the subsurface micrograph indicate that a higher cyclic speed results in decreased DERs
formation. DERs formation can be quantified by a global DER%, characterising the frac-
tional area of overall dark patches in the region of RCF-induced subsurface damage [19].
The global DER% calculated for slow cyclic speed gives an overall DER percentage of 95%
which reduces to 40% at a higher cyclic speed. Likewise, at 5 GPa and 100 ◦ C, a DER%
of 50% can be observed at 289 rpm, whereas no DER can be seen at 4051 rpm speed. At
low-stress frequency, more time is available for carbon to diffuse from ferrite to surrounding
boundaries. In addition, the low-stress frequency promotes more carbon to be captured by
dislocation during the gliding motion as suggested previously [8].

ȱ
Figure 7. Axially sliced micrographs acquired at 6 and 5 GPa, 100 ◦ C till 0.75 million cycles with
(a,c) 4051 rpm (b,d) 289 rpm speed, respectively.

6. Discussions
It is believed that the development of microstructure features involves degeneration
of the parent structure and the microstructural phase transformation during plastic defor-
mation. The subsurface plastic deformation takes place as a result of stress localisation
and forms shear bands in the preferred planes due to non-proportional loading histories.
Figure 8a represents the subsurface micrograph of feature microstructures containing HABs
and LABs alongside the martensitic matrix. In order to evaluate the micromechanical
properties of feature microstructure, a spherical conical indenter of 9 μm in diameter was
employed. The loading was applied in a cyclic loading-partial unloading manner with
a total penetration depth of 400 nm (assuming homogenous depth distribution). Each
loading cycle yielded an elastic/plastic response of the target surface depending upon
penetration depth. The semi-empirical approach to computing flow stress and represen-
tative strain from spherical indentation is described in Appendix A and further details
are available in [26]. The localised flow curves are acquired from the indentation targets,
shown in Figure 8a, and are plotted in Figure 8b). It is apparent from flow curves that
bearing steel exhibits considerable strain hardening under cyclic loadings. The spherical

272
Materials 2022, 15, 8072

indents Ind1, Ind3, and Ind4 demonstrate the martensitic structure’s stress–strain curve.
Ind2 and Ind5 show LCs stress–strain curves, whereas Ind 6 and Ind7 depict stress–strain
curves of LABs and HABs respectively. The martensitic matrix exhibits a baseline flow
curve of bearing steel while significant hardening and softening can be observed in the
regions of LCs and WEBs, respectively. A decrease in hardness in the carbon-depleted
areas (ferritic grains) and an increase in hardness have been reported for carbon-rich areas
(LCs) [9]. Moreover, a localised hardness map of WEBs has revealed hardness variation
within a white band [19]. The variation in localised hardness arises due to grain size devi-
ation, crystallography, change in chemical composition, and indenter penetration depth.
The constitutive stress–strain curves obtained from spherical indentation suggest that the
localised hardness can vary within a WEB, however, the global response of a WEB exhibit
overall softening of the microstructure. In addition, the lowering of yield stress (softening),
as suggested by the constitutive flow curves in the localised regions, indicates the ferritic
nature of WEBs, which is the manifestation of RCF-induced accumulated plastic strain. The
accumulation of plasticity within WEBs results in the formation of ferrite bands comprising
dislocation cells.

ȱ
Figure 8. (a) Circumferentially sliced micrograph representing white etching bands (WEBs) and spherical
nanoindentation performed. (b) Flow curves obtained from nanoindentation analysis of WEBs.

The martensite–ferrite transformation and the formation of dislocation cells are an


indication of high dislocation density in the areas of WEBs. With progressive loading
cycles, the subsurface plasticity accumulates, and forms elongated ferrite bands, refer to
Figure 6. These ferrite bands are oriented in the preferred slip system due to complex
stress histories during rolling cyclic loading. The preferential planes of elongated ferrite

273
Materials 2022, 15, 8072

bands have been described as the maximisation of normal relative stress across the contact
track and have been explained in previous research work [19]. The carbon transport from
ferrite to the grain edge becomes favourable. The free carbon in the ferrite region diffuses
towards boundaries and precipitates as LCs. With continuous cyclic loadings, more and
more carbon atoms become available in the ferrite matrix due to the dissolution of primary
carbides and its transport to thicken the pre-existing carbide plates. To understand the
transport of carbon, tempering and static load analysis have been conducted.
The heat treatment of bearing ball samples results in the reprecipitation of carbides.
The post-tempering results also showed that the thermally activated carbon during ther-
mal tempering cannot lead to formation of WEBs without external stress histories. It
is noticeable that the pre-tempered RCF-tested sample showed the formation of white
bands after 37 million rolling cycles along with DERs, whereas no white band is visible
for the untempered RCF-tested sample. The poorly bonded carbides to the martensitic
matrix during thermal tempering favour the formation of WEBs instead of DERs. The
RCF test conducted with variation in stress frequency illustrated that a lower cyclic speed
increases the time available for carbon to diffuse from ferrite to surrounding boundaries.
Nevertheless, the low stress time also promotes more carbon to be captured by dislocation
during the gliding motion. From Section 4, the numerical results of the carbon concentra-
tion gradient within supersaturated regions of ferrite predicted the diffusion of carbon
from ferrite bands towards the surrounding matrix. However, the tempering results and
static load analysis of bearing ball samples did not reveal the thickening of pre-existing
LCs despite the prolonged tempering of RCF-tested samples under static loading. The
static load analysis confirms that the WEBs development mechanism is determined by the
non-proportional stress histories during cyclic loading which results in accumulated plastic
strains to forming elongated ferrite bands.
It is evident from the comparative analyses that microstructural alterations are formed
as a result of carbon flux, which is dislocation-assisted, in consort with thermally activated
carbon diffusion. However, it is quite difficult to differentiate the individual effects of
carbon diffusion and plastic deformation on the development of a feature microstructure.
It is recommended to employ further material characterisation techniques to quantify the
grain misorientations and lattice deformations within microstructural alterations. The
mechanistic approach employed in the current study is highly significant in terms of
evaluating the microstructural response of RCF-affected regions and can be used to improve
the efficiency of analytical models for DERs/WEBs formation.

7. Conclusions
Bearing balls are tested systematically in a rotary tribometer for further evaluation
of microstructural alterations. After RCF testing and subsequent parametric analyses, the
following points are highlighted,
• The formation mechanism of WEBs accelerates at a temperature above the tempering
condition of bearing steel. The excessive undissolved carbon arising due to thermal
tempering is suggested to enhance the growth of WEBs rather than form DERs.
• The static load RCF test has validated that a complete rolling mechanism with a non-
proportional stress history is essential for the development and unique orientation of WEBs.
• The localised hardness within WEBs can vary, however, the global response of white
bands indicates softening of bearing material, as illustrated by localised flow curves.
• The development of a feature microstructure is a holistic effect of carbon diffusion
and plasticity; and cannot be described solely with dislocation gliding or the carbon
diffusion phenomenon, as presented previously in various research studies.

Author Contributions: M.U.A. contributed to planning, scheduling of experimental work, data


acquisition, analysis, and manuscript writing. Z.A.K. led conception of idea in terms of ordinality
and significance, organised funding, resources, and project management. All authors have read and
agreed to the published version of the manuscript.

274
Materials 2022, 15, 8072

Funding: The work was funded by Schaeffler Technologies AG & Co., KG, Germany (grant ID: 10187)
to conduct this research at Bournemouth University, United Kingdom.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Not applicable.
Acknowledgments: Authors would like to acknowledge Schaeffler Technologies AG & Co., KG,
Germany, and in particular Wolfram Kruhoeffer, for in-kind support to conduct this research.
Conflicts of Interest: The authors declare no conflict of interest.

Appendix A
Nanoindentation analysis is performed at a single point with continuously increasing
multiple cyclic loading and unloading. For this, a 9 μm spherical indenter is used with
NanoTest VANTAGE Micromaterials. The contact area is determined by using the data from
spherical indentation analysis and the given radius of diamond indenter, as demonstrated in
Figure A1. Subsequently, hardness H, Young’s modulus E, stress Y, and strain are calculated.
This cyclic loading–unloading procedure is expressed as the Field & Swain analysis [27].
The spherical indentation has been employed previously [28,29] for mechanical property
evaluation of small volumes of materials.

Figure A1. Schematic for spherical indentation.

With multiple loading cycles during spherical indentation, the total penetration depth
ht (elastic + plastic component) depends on total applied load Pt . During partial unloading,
the recovered depth hs depends upon the reduced load Ps . Once the entire load is removed,
there will be an associated residual depth hr which can be expressed as [29],

r · hs − ht
hr = (A1)
r−1
Whereas r is a factor described as,

2
Pt 3
r= (A2)
Ps

275
Materials 2022, 15, 8072

The elastic depth (he ) and plastic depth (hp ) during an individual loading cycle can be
computed as,
he = ht − hr (A3)
ht + hr
hp = (A4)
2
Plastic depth and radius of spherical indenter R are used to calculate contact circle
radius ‘a’ as given below, which is in the original surface plane at every indentation cycle.

A= 2 R hp − h2p (A5)

Hardness H, the yield/flow stress Yr , and subsequent representative strain εr can be


computed for load step i = 1ton as follow,

Pti Pt
Hi = = i2 (A6)
Ai πai

Hi
Y ri = (A7)
3
a
εri = 0.2 i (A8)
R
Equations (A1)–(A8) [27,29] can be used to extract micromechanical properties from the
indented surface with the help of cyclic loading–unloadings using a conical spherical indenter.

References
1. Arshad, W.; Hanif, M.A.; Bhutta, M.U.; Mufti, R.A.; Shah, S.R.; Abdullah, M.U.; Najeeb, M.H. Technique developed to study
camshaft and tappet wear on real production engine. Ind. Lubr. Tribol. 2017, 69, 174–181. [CrossRef]
2. Ţălu, Ş. Coperta Thread Rolling Technology 2019; Napoca Star Publishing House: Cluj-Napoca, Romania, 2019.
3. Abdullah, M.U.; Shah, S.R.; Bhutta, M.U.; Mufti, R.A.; Khurram, M.; Najeeb, M.H.; Arshad, W.; Ogawa, K. Benefits of wonder
process craft on engine valve train performance. Proc. Inst. Mech. Eng. Part D J. Automob. Eng. 2018, 233, 1125–1135. [CrossRef]
4. Abdullah, M.U.; Khan, Z.A. A Multiscale Overview of Modelling Rolling Cyclic Fatigue in Bearing Elements. Materials 2022,
15, 5885. [CrossRef] [PubMed]
5. El Laithy, M.; Wang, L.; Harvey, T.J.; Vierneusel, B.; Correns, M.; Blass, T. Further understanding of rolling contact fatigue in
rolling element bearings—A review. Tribol. Int. 2019, 140, 105849. [CrossRef]
6. Yin, H.; Wu, Y.; Liu, D.; Zhang, P.; Zhang, G.; Fu, H. Rolling Contact Fatigue-Related Microstructural Alterations in Bearing Steels:
A Brief Review. Metals 2022, 12, 910. [CrossRef]
7. Voskamp, A.P.; Österlund, R.; Becker, P.C.; Vingsbo, O. Gradual changes in residual stress and microstructure during contact
fatigue in ball bearings. Met. Technol. 1980, 7, 14–21. [CrossRef]
8. Fu, H.; Song, W.; Galindo-Nava, E.I.; Rivera-Díaz-Del-Castillo, P.E. Strain-induced martensite decay in bearing steels under
rolling contact fatigue: Modelling and atomic-scale characterisation. Acta Mater. 2017, 139, 163–173. [CrossRef]
9. Šmel, ova, V.; Schwedt, A.; Wang, L.; Holweger, W.; Mayer, J. Electron microscopy investigations of microstructural alterations due
to classical Rolling Contact Fatigue (RCF) in martensitic AISI 52100 bearing steel. Int. J. Fatigue 2017, 98, 142–154. [CrossRef]
10. Martin, J.; Borgese, S.F.; Eberhardt, A.D. Eberhardt, Microstructural alterations of rolling—Bearing steel undergoing cyclic
stressing. ASME J. Basic Eng. 1966, 88, 555–565. [CrossRef]
11. Bhargava, V.; Hahn, G.T.; Rubin, C.A. Rolling contact deformation, etching effects, and failure of high-strength bearing steel. Met.
Mater. Trans. A 1990, 21, 1921–1931. [CrossRef]
12. Schlicht, H. Material Properties Adapted to the Actual Stressing in a Rolling Bearing. In Ball and Roller Bearing Engineering; SKF
Industries Inc.: Philadelphia, PA, USA, 1981; pp. 24–29.
13. Mitamura, N.; Hidaka, H.; Takaki, S. Microstructural Development in Bearing Steel During Rolling Contact Fatigue. In Materials
Science Forum; Trans Tech Publication: Stafa-Zurich, Switzerland, 2007.
14. Warhadpande, A.; Sadeghi, F.; Evans, R.D. Microstructural Alterations in Bearing Steels under Rolling Contact Fatigue: Part
2—Diffusion-Based Modeling Approach. Tribol. Trans. 2014, 57, 66–76. [CrossRef]
15. Polonsky, I.; Keer, L.M. On white etching band formation in rolling bearings. J. Mech. Phys. Solids 1995, 43, 637–669. [CrossRef]
16. Fu, H.; Galindo-Nava, E.; Rivera-Díaz-Del-Castillo, P. Modelling and characterisation of stress-induced carbide precipitation in
bearing steels under rolling contact fatigue. Acta Mater. 2017, 128, 176–187. [CrossRef]
17. Abdullah, M.U. Finite Element Modelling of Deep Zone Residual Stresses in Rolling Contact Bearing Elements. Ph.D. Thesis,
Bournemouth University, Bournemouth, UK, 2022.

276
Materials 2022, 15, 8072

18. Abdullah, M.U.; Khan, Z.A.; Kruhoeffer, W. Evaluation of Dark Etching Regions for Standard Bearing Steel under Accelerated
Rolling Contact Fatigue. Tribol. Int. 2020, 152, 106579. [CrossRef]
19. Abdullah, M.U.; Khan, Z.A.; Kruhoeffer, W.; Blass, T.; Vierneusel, B. Development of white etching bands under accelerated
rolling contact fatigue. Tribol. Int. 2021, 164, 107240. [CrossRef]
20. Abdullah, M.U.; Khan, Z.A.; Kruhoeffer, W.; Blass, T. A 3D Finite Element Model of Rolling Contact Fatigue for Evolved Material
Response and Residual Stress Estimation. Tribol. Lett. 2020, 68, 122. [CrossRef]
21. Kang, J.H.; Hosseinkhani, B.; Rivera-Díaz-del-Castillo, P.E. Rolling contact fatigue in bearings: Multiscale overview. Mater. Sci.
Technol. 2012, 28, 44–49. [CrossRef]
22. Bush, J.J.; Grube, W.L.; Robinson, G.H. Microstructural and residual stress changes in hardened steel due to rolling contact. Trans.
ASM 1961, 54, e412.
23. Warhadpande, A.; Sadeghi, F.; Evans, R.D. Microstructural alterations in bearing steels under rolling contact fatigue part
1—Historical overview. Tribol. Trans. 2013, 56, 349–358. [CrossRef]
24. Bhadeshia, H.; Honeycombe, R. Steels: Microstructure and Properties; Butterworth-Heinemann: Oxford, UK, 2017.
25. Shabalin, I.L. Ultra-High Temperature Materials. I, Carbon (Graphene/Graphite) and Refractory Metals; Springer: Dordrecht, The
Netherlands, 2014.
26. Abdullah, M.U.; Khan, Z.A.; Kruhoeffer, W. Nanoindentation Analysis of Evolved Bearing Steel under Rolling Contact Fatigue
(RCF). In Proceedings of the 2020 STLE Tribology Frontier Virtual Conference Co-sponsored by ASME Tribology Division,
Cleveland, OH, USA, 9–13 November 2020.
27. Field, J.; Swain, M. Determining the mechanical properties of small volumes of material from submicrometer spherical indenta-
tions. J. Mater. Res. 1995, 10, 101–112. [CrossRef]
28. He, L.H.; Fujisawa, N.; Swain, M.V. Elastic modulus and stress–strain response of human enamel by nano-indentation. Biomaterials
2006, 27, 4388–4398. [CrossRef] [PubMed]
29. Kalidindi, S.R.; Pathak, S. Determination of the effective zero-point and the extraction of spherical nanoindentation stress–strain
curves. Acta Mater. 2008, 56, 3523–3532. [CrossRef]

277
materials
Article
Thermo-Mechanical Fatigue Behavior and Resultant
Microstructure Evolution in Al-Si 319 and 356 Cast Alloys
Kun Liu 1 , Shuai Wang 1 , Lei Pan 2 and X.-Grant Chen 1, *

1 Department of Applied Science, University of Quebec at Chicoutimi, Saguenay, QC G7H 2B1, Canada
2 Arvida Research and Development Centre, Rio Tinto Aluminum, Saguenay, QC G7S 4K8, Canada
* Correspondence: [email protected]

Abstract: The out-of-phase thermo-mechanical fatigue (TMF) behavior of the two Al-Si cast alloys
most widely used for engine applications (319 and 356) were investigated under temperature cycling
(60–300 ◦ C) and various strain amplitudes (0.1–0.6%). The relationship between the microstructural
evolution and TMF behavior was closely studied. Both alloys exhibited asymmetric hysteresis loops
with a higher portion in the tensile mode during TMF cycling. The two alloys showed cyclic softening
behavior with regard to the maximum stress, but an earlier inflection of cyclic stress was found in the
356 alloy. The TMF lifetime of the 319 alloy was generally longer than that of the 356 alloy, especially
at higher strain amplitudes. All the precipitates (β -MgSi in 356 and θ -Al2 Cu in 319) coarsened
during the TMF tests; however, the coarsening rate per cycle in the 356 alloy was significantly
higher than that in the 319 alloy. An energy-based model was applied to predict the fatigue lifetime,
which corresponded well with the experimental data. However, the parameters in the model varied
with the alloys, and the 356 alloy exhibited a lower fatigue damage capacity and a higher fatigue
damage exponent.

Keywords: Al-Si casting alloys; thermo-mechanical fatigue; precipitates; fatigue life prediction; fracture

Citation: Liu, K.; Wang, S.; Pan, L.;


Chen, X.-G. Thermo-Mechanical 1. Introduction
Fatigue Behavior and Resultant Al-Si cast alloys are widely used in modern automotive industries to replace cast iron-
Microstructure Evolution in Al-Si 319
fabricated engine components (engine blocks and cylinder heads) due to their high strength-
and 356 Cast Alloys. Materials 2023,
to-weight ratio, excellent castability, and thermal conductivity [1–4]. In combustion engine
16, 829. https://doi.org/10.3390/
applications, components undergo complex loading changes and dramatic temperature
ma16020829
gradients during start-up and shutdown cycles. For example, engines are often heated
Academic Editors: Lucjan Śnieżek, from ambient temperature—or even lower in cold winters (−25 ◦ C)—to 250–300 ◦ C and,
Jaroslaw Galkiewicz and vice versa, rapidly cooled down in start–stop cycles [5]. Due to the large thermal gradient,
Sebastian Lipiec significant thermal and mechanical stresses occur in the engine components because of
Received: 22 December 2022
the changing thermal expansion/contraction of various engine components [6]. As a
Revised: 9 January 2023
result, cyclic stress and temperature changes can result in thermo-mechanical fatigue
Accepted: 12 January 2023
(TMF) [4,6], which can lead to serious failures, significantly limiting components’ service
Published: 15 January 2023 life. Therefore, understanding TMF behavior is increasingly important for the design,
reliability assessment, and lifecycle management of critical engine components during
their industrial application. It has been reported [4–6] that the most significant damage
mechanism in engine components is out-of-phase (OP) TMF cycling, where the maximum
Copyright: © 2023 by the authors. mechanical strain occurs at the minimum temperature.
Licensee MDPI, Basel, Switzerland. Several studies have evaluated the TMF behavior in Al-Si cast alloys [7–13]. Huter et al. [11]
This article is an open access article performed TMF tests on Al-Si-Cu and Al-Si-Mg cast alloys with varying Si, Cu, Fe, and Sr
distributed under the terms and contents and reported that the nucleation and propagation of cracks were predominantly
conditions of the Creative Commons influenced by the eutectic Si and Fe-rich intermetallics, which lowered the fatigue resistance.
Attribution (CC BY) license (https://
Sehitoglu et al. [14] reported similar results for the Al-Si 319 alloy, finding that the alloy
creativecommons.org/licenses/by/
presented worse TMF resistance at higher Fe contents due to a lower stress level and
4.0/).

Materials 2023, 16, 829. https://doi.org/10.3390/ma16020829 279 https://www.mdpi.com/journal/materials


Materials 2023, 16, 829

higher softening speed. It was also found that the rapidly solidified material with a small
secondary dendrite arm space (SDAS) presented the highest cyclic stress–strain amplitude,
whereas the stress levels decreased gradually as the SDAS levels increased [14]. It has
also been reported that a finer SDAS, lower porosity level, and lower brittle intermetallic
content are beneficial for better thermal resistance and, hence, the fatigue life of defect-free
A356 alloys can be improved by at least one order of magnitude [15,16].
A recent study employed ultrasonic melt treatment for Al-Si cast alloys to refine the
grain size, the eutectic Si, and the Fe-rich intermetallics, which significantly enhanced the
fatigue life [17]. However, contradictory results regarding the influence of precipitates
on the TMF resistance of Al-Si cast alloys can be found in the literature. For example,
Azadi et al. [8,9] studied the effect of heat treatment and strain rate on the TMF cyclic
behavior of the A356 alloy and found that the TMF life was not significantly affected by the
heat treatment in the as-cast and T6 conditions. In contrast, Moizumi et al. [18] reported
that the thermal fatigue life of an A356 + 1%Cu alloy in the over-aged condition was
doubled compared to the peak-aged condition. Hunter et al. [11] also reported that the
TMF resistance of an Al-Si-Cu cast alloy was improved compared to that of a Cu-free Al-Si
alloy due to the Cu-containing precipitates. Beck et al. [19] studied the influence of high-
cycle fatigue (HCF) loading on the TMF life of Al-Si-Mg/Cu cast alloys in the T6 condition
and concluded that TMF was not significantly affected by HCF loading at low strain
amplitudes and that precipitation in the T6 condition was effected by HCF. Toda et al. [13]
reported that precipitates preferentially oriented perpendicular to the loading direction can
effectively prolong the in-phase TMF life, but the study did not comment on the OP-TMF
life. Toyoda et al. [20] investigated the microstructural change in an Al-Si-Cu-Mg cast
alloy under thermo-mechanical loading conditions and reported that the precipitates were
observed to align perpendicularly to the compressive stress during heating, but they did
not further discuss how these precipitates with preferential orientations affected the TMF
behavior. Therefore, the systematic evolution of precipitates during TMF under various
strain amplitudes, which is one of the significant factors influencing the TMF resistance, is
still not well-documented in Al-Si cast alloys.
Contemporary engine components are designed to increase the maximum operating
temperature and pressure in order to further reduce greenhouse gas emissions and improve
their fuel efficiency [5,19,21,22]. However, the TMF behaviors reported in the literature are
limited to relatively low maximum temperatures of 200 or 250 ◦ C [9,11,16,18,19], and the
TMF behavior at higher maximum temperatures (approximately 300–350 ◦ C) is seldom
reported. Moreover, the microstructural evolution during TMF over a wider temperature
range, especially precipitation in age-hardening aluminum alloys, is still not fully under-
stood. This study selected two of the most widely used Al-Si cast alloys—namely, Al-Si 356
and Al-Si-Cu 319—to perform OP-TMF under temperature cycling (60–300 ◦ C) and various
strain amplitudes (0.1–0.6%). The TMF behaviors were evaluated using the strain and stress
cyclic response, softening rate, and fatigue life assessment. The evolution of the precipitates
during TMF cycling was characterized using transmission electron microscopy (TEM) to
establish the relationship between the microstructure and TMF behavior. Additionally, the
TMF damage mechanism is discussed and an energy-based model for the prediction of the
TMF lifetimes of the two alloys is introduced.

2. Materials and Methods


Two Al-Si cast alloys (319 and 356) were prepared using commercially pure aluminum
(99.7%), pure Mg (99.9%), and Al-50%Si, Al-25%Mn, Al-50%Cu, Al-10%Sr, and Al-5%Ti-
1%B master alloys. The raw materials were batched in clay-graphite crucibles and melted
using an electric resistance furnace. After reaching 780 ◦ C, the melt was held for 30 min
and then degassed with high purity argon gas for 15 min. The melt was then cast into
a permanent wedge mold pre-heated at 250 ◦ C. The chemical compositions of the two
alloys were analyzed using optical emission spectroscopy, and the results are listed in
Table 1. All the as-cast samples were subjected to a two-step solution treatment followed

280
Materials 2023, 16, 829

by water quenching. Artificial aging was then performed at 200 ◦ C for 5 h to achieve the T7
temper condition. Table 2 summarizes the parameters of the heat treatment applied to the
alloy samples.

Table 1. Chemical compositions of the alloys studied (wt.%).

Alloy Si Cu Mg Mn Fe Ti Sr Al
319 5.93 3.34 0.12 0.284 0.307 0.11 0.0106 Bal.
356 7.27 0.60 0.34 0.206 0.109 0.21 0.0113 Bal.

Table 2. Heat treatment parameters for the alloys.

Alloy Solution Treatment T7 Aging Treatment


319 495 ◦ C/4 h + 515 ◦ C/2 h
200 ◦ C/5 h
356 500 ◦ C/4 h + 540 ◦ C/2 h

After the T7 heat treatment, the TMF specimens were machined to a parallel gauge
length of 75 mm and diameter of 10 mm and a 5 mm diameter hole was machined through
the specimens for air circulation, as schematically shown in Figure 1. In the present work,
OP-TMF, where the highest mechanical strain is reached at the lowest temperature, was
applied to the experimental alloys (as shown in Figure 1a). OP-TMF tests were performed
using a Gleeble 3800 thermal–mechanical simulator system with a strain-controlled loading
mode. The samples were heated using a Joule heating system and cooling was achieved by
using compressed air delivered through a hollow tube specimen. The TMF temperature
was varied between 60 and 300 ◦ C for a cycle using heating and cooling rates of 5 ◦ C/s.
The temperature was measured and controlled using thermocouples attached to the center
of the gauge length. The applied mechanical strain amplitudes were 0.1, 0.2, 0.4, and 0.6%.
The TMF tests were automatically terminated when either a total of 2000 fatigue cycles
was reached or a decrease of 30% in the initial maximum tensile/compression stress was
detected. When the TMF test was terminated without full fracture of the sample, the sample
was pulled to full fracture in the Gleeble machine at room temperature for a later fracture
analysis. More details on the TMF test procedure, such as the process to reach zero stress
at the beginning of the test and the isolation of the mechanical strain from the total strain
during the TMF test, can be found in our previous study [23].
The microstructures of the TMF samples were characterized using optical and electron
microscopies. Optical microscopy was used to observe the as-cast and T7 grain structures
and intermetallics, whereas a scanning electron microscope (SEM) equipped with an energy-
dispersive X-ray spectroscope was used to identify the intermetallic phases and to observe
the fracture faces after the TMF tests. TEM was used to observe the evolution of the
precipitates during TMF testing. TEM samples were cut from the cast ingot after the T7
treatment and from the gauge zone of the samples (close to fracture) after the TMF tests at
various strain amplitudes. The TEM samples were initially punched into 3 mm diameter
disks and then ground and polished to approximately 30 μm. The samples were then
twin-jet electrochemically polished using a solution of 75 mL HNO3 in 250 mL methanol
at −30 ◦ C to create TEM observation regions. Image analysis was used to quantify the
characteristics of the intermetallics and precipitates, such as their area fraction and number
density, under various conditions.

281
Materials 2023, 16, 829

Figure 1. Schematic illustration of OP-TMF at 0.4% strain amplitude (a), TMF sample dimensions (b),
and the setup with gripping blocks (c).

3. Results
3.1. TMF Behaviors
Figure 2 shows the stress–strain hysteresis loops of the initial, mid-life, and end-life
cycles of the two alloys at the mechanical strain amplitudes of 0.1, 0.2, 0.4, and 0.6%. Both
alloys exhibited some common phenomena. First, both alloys exhibited asymmetrical
hysteresis loops at all strain amplitudes, and the tensile portions were higher than the
compression portions, indicating a higher tensile stress than compression stress at the same
tensile/compression strain, which has also been reported in the literature [8,11,14,18,23].
This can be attributed to the characteristics of OP-TMF, where the maximum tensile strain
is reached at the lowest temperature (60 ◦ C) and the compression strain increases to a
maximum at the highest temperature (300 ◦ C). The mechanical properties of Al-Si cast
alloys worsen with increasing temperature [5,24,25]. Therefore, the stress needed to reach
the desired strain was higher in the tensile mode due to its low temperature compared to
the compression mode at high temperature, leading to asymmetrical hysteresis loops with
higher portions in the tensile mode (stress > 0 MPa in Figure 2). Second, the loop areas
were significantly larger and the peak tensile/compression stress increased with increasing
strain amplitude. As shown in Figure 2a,c,e,g, the loop area of the 319 alloy gradually
increased with an increase in the strain amplitude from 0.1 to 0.6%, whereas the mid-life
peak tensile stress increased from 100 MPa at 0.1% to 124, 198, and 211 MPa at 0.2, 0.4, and
0.6%, respectively. A comparison of the mid- and end-life cycles with the first cycle showed
that the shape of the hysteresis loop became flatter with increasing fatigue life at a fixed
strain amplitude, indicating an increasing plastic strain, which can be calculated using the
strain range at zero stress. Additionally, the maximum tensile and compression stresses
decreased at all strain amplitudes, indicating cyclic softening behavior [10,11,23]. Taking
the TMF cycles of alloy 319 at 0.4% strain amplitude as an example (Figure 2e), the plastic
strain increased from 0.36% after the second cycle to 0.45 and 0.51% at the mid-life and

282
Materials 2023, 16, 829

final cycles, respectively, while the maximum tensile stress decreased from 236 MPa after
the second cycle to 198 and 156 MPa at the mid-life and final cycles, respectively.

Figure 2. OP-TMF stress–strain hysteresis loops for experimental alloys at different strain amplitudes
of 0.1, 0.2%, 0.4%, and 0.6%: (a,c,e,g) 319 alloy; (b,d,f,h) 356 alloy.

283
Materials 2023, 16, 829

On the other hand, differences between the two experimental alloys were also ob-
served. Figure 2 shows that the 319 alloy always exhibited higher maximum/minimum
stress than the 356 alloy, whereas the extent of the strain range at zero stress (plastic strain)
was always greater for the 356 alloy than for the 319 alloy. For example, the maximum and
minimum stresses of the 319 alloy during the mid-life cycle under a 0.6% strain amplitude
(Figure 2g,h) were 206 and 121 MPa, respectively, which were significantly higher than
the respective values of 96 and 86 MPa obtained for the 356 alloy. In contrast, the plastic
strain in the 356 alloy was 0.8%, higher than the value of 0.6% in the 319 alloy. These
differences can be attributed to the different mechanical properties of the two alloys, with
the 319 alloy with a high Cu content exhibiting higher strength but lower elongation than
the 356 alloy [5,24–26].
Figure 3 shows the evolution of the maximum tensile and compression stresses of
the 319 and 356 alloys as a function of the fatigue cycles at different strain amplitudes.
The maximum tensile stress for both alloys was always higher than the compression
stress, which was due to the higher strength at lower temperatures in OP-TMF [11,23].
Additionally, the maximum tensile/compression stress generally decreased with increasing
numbers of cycles, indicating cyclic softening for both alloys, which is also reflected in
Figure 2.

Figure 3. Evolution of the maximum tensile and compression stresses as a function of the cycle
number at different strain amplitudes: (a) 319 alloy; (b) 356 alloy.

In addition to the common phenomena in the two alloys, Figure 3 also shows some
differences between the two. First, similarly to Figure 2, the maximum tensile/compression
stresses of the 319 alloy were always higher than those of the 356 alloy. For example, the
initial maximum tensile and compression stresses of the 319 alloy at a strain amplitude of
0.6% were 245 and 110 MPa, respectively, which were higher than the respective values of
190 and 97 MPa obtained for the 356 alloy. Second, the decrease in stress as a function of
the number of cycles differed for the two alloys. As indicated by the arrows in Figure 3, the
point where the softening rate exhibited a sharp change was defined as the inflection point
at which the TMF could generally be divided into two different stages. Stage I spanned
from the initial fatigue cycle to the inflection point. In this stage, the stress level initially
increased during the first few cycles and then gradually decreased with a low softening rate.
Stage II spanned from the inflection point to the point where the specimen failed, at which
point the softening rate significantly increased. In this stage, the stress sharply decreased
with increasing numbers of TMF cycles, which can be attributed to the occurrence of large
cracks that exceeded the critical crack length [11]. The 356 alloy always exhibited an earlier
inflection point than the 319 alloy, especially under higher strain amplitudes (0.4 and 0.6%),
which resulted in early fatigue damage in the 356 alloy and, hence, reduced total fatigue
life. For an improved comparison, the proportions of stage I in the total lifetimes of the two

284
Materials 2023, 16, 829

alloys were calculated, and the results are listed in Table 3. The table shows that stage I
accounted for more than 90% of the total lifetime of the 319 alloy at all the tested strain
amplitudes, even reaching 100% at the strain amplitude of 0.6%, which may indicate that
there was little macro-damage before the end of the TMF test. However, the proportion of
stage I in the 356 alloy was considerably lower, especially at higher strain amplitudes. As
shown in Table 3, stage I accounted for only 22 and 27% of the total lifetime of the 356 alloy
at the 0.4 and 0.6% strain amplitudes, respectively, indicating that the 356 alloy may have
been subjected to earlier severe macro-damage during TMF cycling.

Table 3. Proportion of stage I in total TMF lifetimes of experimental alloys.

Strain Amplitude 319 Alloy 356 Alloy


0.2% 90% 85%
0.4% 96% 22%
0.6% 100% 27%

Figure 4 summarizes the OP-TMF lifetimes of the 319 and 356 alloys under different
strain amplitudes. Generally, the fatigue life decreased with increasing strain amplitude,
and both alloys exhibited approximately linear relations when plotted on a log–log scale.
Both alloys reached the limit of 2000 cycles at a strain amplitude of 0.1%; however, the
319 alloy exhibited increased fatigue life starting at the 0.2% strain amplitude, especially
at higher strain amplitudes (0.4 and 0.6%). As shown in Figure 4, the average fatigue life
of the 319 alloy at a 0.2% strain amplitude was 700 cycles, which was slightly higher than
the 580 cycles obtained for the 356 alloy. The 319 alloy exhibited a significantly longer
average fatigue life than the 356 alloy at higher strain amplitudes, demonstrating two- and
threefold longer average TMF lifetimes at the 0.4 (183 versus 90) and 0.6% strain amplitudes
(62 versus 20), respectively. Therefore, the TMF resistance of the 319 alloy was generally
higher than that of the 356 alloy, especially when exposed to higher strain amplitudes.

Figure 4. OP-TMF lifetimes of two experimental alloys under various strain amplitudes.

3.2. Fracture Analysis


Figure 5 shows overall views of the fracture surfaces of the 319 and 356 alloys after
TMF cycling at 0.2% strain amplitude. Porosity was observed on both fracture surfaces,
as marked by the red circles in Figure 5. The porosity in the 319 alloy was relatively low

285
Materials 2023, 16, 829

and evenly spread across the sample. Crack ridges were found on the fracture surface of
the 319 alloy, indicating that the cracks may have nucleated at multiple initiation sites and
merged together during fatigue failure. Large and continuous pores were observed on the
fracture surface of the 356 alloy, and these could have served as crack nucleation sites and
provided easier propagation paths for the crack [11,15–17], which was one of the likely
reasons for the earlier inflection point in the 356 alloy (see Figure 3). It is worth noting that
the left part in Figure 5b (356 alloy, marked by two red lines) was not the original TMF
fracture surface but the tensile fracture surface from the pulling process after the TMF test,
as the sample was not fully fractured when the TMF was terminated. As explained in the
Materials and Methods section, the sample that remained unfractured upon termination of
the test was pulled to full fracture.

Figure 5. Fracture surface after 0.2% TMF test: (a) 319 alloy; (b) 356 alloy.

Figure 6 shows the fracture faces at the crack propagation areas in the 319 and
356 alloys at 0.2% strain amplitude. As shown in Figure 6a,d, both fracture faces appeared
as ductile fractures with dimples in the matrix. However, the dimples in the 319 alloy were
smaller than those in the 356 alloy, indicating its lower plastic strain. The SEM backscatter
images in Figure 6b,e show that a number of bright phases were found on the fracture sur-
face, as marked by the arrows, which were identified as Fe-rich intermetallics and eutectic
Si particles [24,25]. The 319 alloy (Figure 6b) exhibited more Fe-rich intermetallics than the
356 alloy (Figure 6e), which was attributed to its higher Fe content (0.3 and 0.1% in the 319
and 356 alloys, respectively (Table 1)). As reported in [5,11,17], these hard intermetallic
phases could serve as nucleation sites for cracks as a result of their detachment from the
aluminum matrix or by rupturing themselves. This was confirmed by the observation of
the cross-sectional fracture surface in Figure 7, in which several broken eutectic Si particles
and Fe-rich intermetallic phases were observed in the 319 and 356 alloys after the TMF
tests. Table 4 summarizes the characteristics of the Fe-rich intermetallics and eutectic Si
particles in both alloys in the T7 condition (before TMF). The eutectic Si in the 356 alloy
was more globular and had a higher area fraction, whereas the area fraction of the Fe-rich
intermetallics in the 319 alloy was double that in the 356 alloy. However, the total area
fractions of the eutectic Si and Fe-rich intermetallics in both alloys were similar, indicat-
ing their similar contributions to the nucleation and propagation of cracks during TMF
deformation. Additionally, striations were observed in both the 319 (Figure 6c) and 356
(Figure 6e) alloys. The interspacing of the striations in the 356 alloy was larger than that in
the 319 alloy, indicating faster fatigue propagation during TMF cycling [27].

286
Materials 2023, 16, 829

Figure 6. Fractography of fractured specimens after 0.2% TMF test: (a–c) 319 alloy; (d,e) 356 alloy;
(a,c,d,f) SEM secondary electron images; (b,e) SEM backscatter images.

Figure 7. Optical cross-section images of the fracture surfaces showing the cracks in the (a) 319 alloy
and (b) 356 alloy after the 0.2% TMF test.

Table 4. Microstructure characteristics of experimental alloys under T7 condition (before TMF).

319 Alloy 356 Alloy


Morphology of eutectic Si Lamellar Globular
Diameter of eutectic Si 4.0 μm 3.5 μm
Area fraction of eutectic Si 6.6% 8.8%
Area fraction of intermetallics (total) 1.6% 0.8%

287
Materials 2023, 16, 829

4. Discussion
4.1. Microstructural Evolution during TMF
As shown in Figure 3, the maximum tensile and compression stresses for both al-
loys gradually decreased with the increasing numbers of cycles at all strain amplitudes.
This cyclic softening behavior during TMF was likely related to the evolution of the pre-
cipitates during the TMF cycling [12,16,18,20,23]. As typical precipitation-strengthening
alloys, the precipitates in both alloys play an important role in their mechanical properties.
Figure 8 shows the initial precipitate microstructures of the two alloys in the T7 condition
(before TMF). A large number of nano-sized precipitates were visible in the aluminum
matrixes of both alloys, which were identified as mainly the plate-like θ -Al2 Cu phase in
the 319 alloy and the needle-like β -MgSi phase in the 356 alloy, consistent with reports
from the literature [5,16,25].

Figure 8. Bright-field TEM images showing precipitate microstructure after T7 (before TMF) for
(a) θ -Al2 Cu in the 319 alloy and (b) β -MgSi in the 356 alloy.

Figure 9 presents the precipitate microstructures of the two alloys after TMF cycling
at various strain amplitudes. In general, all the precipitates in the two alloys coarsened
during TMF cycling. Figure 9 shows that, after TMF cycling, the plate-like θ -Al2 Cu in the
319 alloy became rod-like θ -Al2 Cu, whereas the needle-like β -MgSi coarsened to rod-like
β -MgSi. However, their number densities decreased significantly as their size increased.
As shown in Table 5, the number density of the precipitates in the 319 alloy after TMF
cycling at 0.2% strain amplitude decreased from 4451 to 501 μm−1 and the length increased
from 48 to 115 nm compared to the T7 condition (before TMF). A similar phenomenon was
also observed in the 356 alloy. Though both the θ -Al2 Cu and β -MgSi coarsened during
the TMF, different coarsening behaviors were observed for the two alloys. Compared to
the evolution of θ -Al2 Cu in the 319 alloy shown in Figure 9, the size and number-density
changes for the β -MgSi during TMF were significantly larger, especially at the higher
strain amplitude (0.6%) (compare Figure 9e,f). As shown in Table 5, after TMF cycling at
the 0.6% strain amplitude, the length of the θ -Al2 Cu in the 319 alloy increased from 48 to
80 nm and the number density decreased from 4451 to 691 μm−1 . On the other hand, the
size of the β -MgSi in the 356 alloy increased from 70 to 205 nm and the number density
significantly decreased from 19,144 to 394 μm−1 . Additionally, a small part of the β -MgSi
in the 356 alloy transformed into equilibrium β-Mg2 Si after TMF cycling, as indicated by the
red arrows in Figure 9b,d,f. The rod-like θ -Al2 Cu remained the predominant precipitate
in the 319 alloy after TMF cycling, which indicated that the precipitates in the 356 alloy
exhibited a higher coarsening rate than those in the 319 alloy.

288
Materials 2023, 16, 829

Figure 9. Bright-field TEM images showing precipitate coarsening after TMF at various strain
amplitudes: (a,c,e) 319 alloy and (b,d,f) 356 alloy.

289
Materials 2023, 16, 829

Table 5. Characteristics of precipitates of the alloys before and after TMF.

Length Width Number Density


Alloys Conditions
(nm) (nm) (μm−1 )
T7 before TMF 48.1 4.3 4451.2

319 After 0.2% 115.9 12.1 501.4


(θ -Al2 Cu) After 0.4% 83.9 10.0 789.8
After 0.6% 80.1 18.1 690.6
T7 before TMF 70.1 2.8 19144.8

356 After 0.2% 331.1 14.9 191.1


(β -MgSi) After 0.4% 204.7 14.6 253.7
After 0.6% 204.8 12.6 394.0

The coarsening of precipitates during TMF cycling was mainly related to their TMF
behavior. In this study, the coarsening rates of both the precipitates were evaluated using
the classical Lifshitz–Slyozov–Wagner model, which is expressed by Equation (1) [28]:

Ln − L0n = k · t (1)

where L is the average half-length of the precipitates after TMF cycling, L0 is the average
half-length of the precipitates in the T7 condition (before TMF), k is the coarsening rate
constant, t is the time, and n is the temporal exponent. It has been reported [29,30] that the
growth of θ -Al2 Cu and β -MgSi precipitates is prone to obeying the t1/2 law (n = 2) under
the dominant coarsening of the interface reaction. Thus, n = 2 was adopted in this study.
Additionally, the value of t is normally taken as the time elapsed at constant temperature.
However, in this study, the temperature in each fatigue cycle changed, meaning that the
coarsening rate per second could not be applied here due to the temperature variation
between 60 and 300 ◦ C in each cycle. For simplification, the value of t was modified to the
number of TMF cycles in this study, and the value of k in Equation (1) was calculated as the
coarsening rate for the precipitates per cycle. The k values of the precipitates in the two
alloys were calculated using the data in Table 5, and the results are summarized in Table 6.

Table 6. Coarsening rate during TMF under various strain amplitudes.

k Value
Strain Amplitude
319 Alloy 356 Alloy
0.2% 3.45 43.85
0.4% 6.22 87.32
0.6% 16.05 463.01

As shown in Table 6, k generally increased with an increasing strain amplitude, indi-


cating that a higher strain amplitude resulted in a higher precipitate coarsening rate. This
was attributed to the heavy dislocations in the matrix, which were generated under the
high-strain-amplitude conditions, accelerating the coarsening of the precipitates [31,32].
However, the 319 alloy exhibited a significantly lower coarsening rate than the 356 alloy
at each strain amplitude, and the differences in k between the two alloys also increased
with the strain amplitude. For example, k in the 319 alloy at a 0.2% strain amplitude was
3.45, which was only one tenth of the value of 43.85 obtained for the 356 alloy. It was even
found that k in the 319 alloy at the 0.4 and 0.6% strain amplitudes decreased to levels 1/5th
(6.2 versus 87.3) and 1/30th (16 versus 463), respectively, of those in the 356 alloy.
The lower coarsening rate for θ -Al2 Cu in the 319 alloy resulted in a slower decrease
in the maximum stress during TMF (Figure 3), which has also been reported in the litera-

290
Materials 2023, 16, 829

ture [11,18]. The lower coarsening of the precipitates and slower decrease in the maximum
stress could stabilize both the matrix and stress level and, hence, enhance the TMF en-
durance of the 319 alloy (Figure 4). The coarsening rate of the precipitates per cycle in the
319 alloy, especially at the high strain amplitude (0.6%), was significantly lower than that
in the 356 alloy, leading to a longer TMF life for the 319 alloy. In contrast, the plastic strain
in the 356 alloy during TMF cycling was higher than that in the 319 alloy, and the rate of
decrease in the maximum stress was also higher (Figure 3), which can result in a greater
portion of the sample deforming during TMF loading, as well as the buckling effect in the
gauge area, which further produces early failures [11]. This phenomenon occurred more
frequently at the higher strain amplitudes. As shown in Figure 3, the stress inflection point
of the 356 alloy at the strain amplitude of 0.6% occurred after only a few cycles, resulting in
a significantly shorter TMF life relative to that of the 319 alloy.

4.2. TMF Lifetime Prediction


Thermo-mechanical fatigue lifetimes are difficult to predict due to the complex defor-
mation under both strain and temperature cycling during TMF. Various models have been
developed based on different mechanisms, including the Neu–Sehitoglu model based on
creep, fatigue, and oxidation damage [33]; the Miller model based on the accumulation
of the damage rate [34]; the J-integral model based on fracture mechanics [35]; and the
energy-based model based on the dissipated energy per cycle [8,36–38]. Among these
models, the energy-based model is widely accepted as a more suitable approach to predict
TMF lifetimes [8,36–40]. The hysteresis energy in the energy-based model involves both the
strain and stress amplitudes and can be approximated as the product of the plastic strain
range (Δεp ) and the stress range (Δσ) using Equation (2) [8,40]:
"
Wi = σdε ≈ Δε p · Δσ (2)

where Wi represents the hysteresis energy variable of the ith cycle. The evolution of the
hysteresis energy (W, calculated using the hysteresis loop area) with increasing numbers of
cycles under various strain amplitudes in the two alloys, as determined using Equation (2),
is shown in Figure 10. The hysteresis energy for both alloys involved both the strain and
stress amplitudes and exhibited cyclic stability after a short period of adaptation for most
of the strain amplitudes applied, confirming the greater rationality of the energy-based
model. However, an early energy loss occurred in the 356 alloy after the first few cycles at
the 0.6% strain amplitude, which may have been related to early crack formation during
TMF loading.

Figure 10. Evolution of hysteresis loop area (W) as a function of cycles for (a) 319 alloy and
(b) 356 alloy.

291
Materials 2023, 16, 829

On the other hand, the hysteresis energy (equal to the plastic strain energy) is also
related to the fatigue life (Nf ), as shown in Equation (3) [8,36]:

−1/β
Ws = W0 · N f (3)

where Ws is the saturation hysteresis plastic strain energy. As shown in Figure 10, the plastic
strain energy reached the saturation stage after only a few cycles, which thus allowed for Ws
to be determined from the fatigue data after that point, and it was generally obtained at the
mid-life cycle of the completed TMF tests. W 0 and β are material parameters representing
the fatigue damage capacity and fatigue damage exponent, respectively. They can be
calculated from the logWs − logNf relationship [36,39]. In this way, the W 0 and β calculated
from the known fatigue life and plastic strain energy under a selection of strain amplitudes
can be used with Equation (3) to estimate the TMF life under other strain amplitudes after
only implementing a few cycles, such as 10; thus, the saturation plastic strain energy can be
obtained, which is more time-efficient than running a full TMF test at each strain amplitude.
Due to the limited data in this study, the TMF behaviors at the 0.2 and 0.6% strain
amplitudes were used to calculate W 0 and β. The TMF life at the 0.4% strain amplitude was
then estimated using Equation (3) and compared with the experimental value to validate
the model. The data for both alloys after 2000 cycles at the 0.1% strain amplitude were
obtained by triggering the settled limit of the test and could not be used to determine the
W 0 and β. Table 7 lists the calculated W 0 and β values, which are similar to the reported
values for Al-Si cast alloys [36]. The 319 alloy exhibited a higher fatigue damage capacity
value (W 0 ) than the 356 alloy. The exponent value (β) represents the relation between the
TMF lifetime and hysteresis energy, and the higher β value for the 356 alloy indicates that
the TMF lifetime decreased significantly with increasing hysteresis energy.

Table 7. Material parameters calculated with energy-based model.

W0 β
319 alloy 107.7 1.09
356 alloy 5.27 2.28

The predicted fatigue lifetime at the 0.4% strain amplitude was calculated using
Equations (2) and (3), as well as the data in Table 7, and compared with the experimental
fatigue life, as listed in Table 8. The results illustrate that the predicted lifetime corresponded
well with the experimentally measured lifetime. Figure 11 presents the experimental and
predicted TMF lifetimes at 0.2–0.6% strain amplitudes. All the predicted TMF lifetimes
correlated with the low life prediction factor (LPF; approximately 1.3) obtained from
the experimental life [36], confirming that the energy-based model was suitable for the
prediction of the TMF lifetimes of the two alloys and had acceptable accuracy. Using the
parameters from Table 7 and the TMF behavior shown in Figure 3, the TMF lifetimes of the
319 and 356 alloys at the strain amplitude of 0.1% were estimated to be 4152 and 4730 cycles,
respectively, both of which were significantly higher than 2000 cycles.

Table 8. Experimental and predicted TMF lifetimes under 0.4% strain amplitude for the two alloys.

319 356
Test 1 Test 2 Test 1 Test 2
Experimental 189 175 80 106
Predicted 163 155 61 92

292
Materials 2023, 16, 829

Figure 11. Comparison of predicted and experimental fatigue lifetimes for 356 and 319 alloys.

5. Conclusions
This study investigated and compared the OP-TMF behaviors of two typical Al-Si cast
alloys (319 and 356) and the following conclusions were drawn:
1. During TMF loading, asymmetrical hysteresis loops were observed in both alloys
and the tensile portion was higher than the compression portion due to out-of-
phase cycling;
2. Cyclic softening behavior was observed for the maximum stress in both alloys, but
the rate of decrease in the cyclic stress in the 356 alloy was greater than that in the
319 alloy, especially at higher strain amplitudes. Moreover, the 356 alloy presented an
earlier inflection point in the cyclic evolution of the cyclic stress;
3. The TMF resistance decreased with increasing strain amplitudes. The 319 alloy exhib-
ited a longer TMF lifetime than the 356 alloy, especially at higher strain amplitudes;
4. The precipitates in both alloys coarsened during TMF cycling. The coarsening rate
of the β -MgSi in the 356 alloy per cycle was higher than that of the θ -Al2 Cu in the
319 alloy and increased at higher strain amplitudes;
5. The fatigue lifetime predicted using the energy-based model corresponded well with
the experimental results, exhibiting a low life prediction factor of 1.3. However, the
material parameters varied with the alloys, and the 356 alloy exhibited a lower fatigue
damage capacity (W 0 ) and a higher fatigue damage exponent (β).

Author Contributions: K.L.: conceptualization, methodology, validation, writing—original draft


preparation; S.W.: investigation, formal analysis; L.P.: validation, review and editing; X.-G.C.:
conceptualization, methodology, validation, review and editing, supervision, funding acquisition.
All authors have read and agreed to the published version of the manuscript.
Funding: This research was funded by the Natural Sciences and Engineering Research Council of
Canada (NSERC) under grant no. CRDPJ 514651-17.
Institutional Review Board Statement: Not Appliable.
Informed Consent Statement: Not Appliable.
Data Availability Statement: Supporting data are available upon reasonable request.

293
Materials 2023, 16, 829

Acknowledgments: The authors would like to acknowledge the financial support of the Natural
Sciences and Engineering Research Council of Canada (NSERC) under grant no. CRDPJ 514651-17
and Rio Tinto Aluminum through the Research Chair in the Metallurgy of Aluminum Transformation
at the University of Quebec in Chicoutimi. Special thanks are given to Jian Qin for performing part of
the TMF tests.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Ye, H. An overview of the development of Al-Si-Alloy based material for engine applications. JMEP 2003, 12, 288–297. [CrossRef]
2. Cole, G.; Sherman, A. Light weight materials for automotive applications. Mater. Charact. 1995, 35, 3–9. [CrossRef]
3. Garat, M.; Laslaz, G. Improved Aluminium Alloys for Common Rail Diesel Cylinder Heads. 2007. Available online: https://
www.researchgate.net/publication/284890980_Improved_aluminum_alloys_for_common_rail_diesel_cylinder_headsJ (accessed
on 30 June 2022).
4. Javidani, M.; Larouche, D. Application of cast Al–Si alloys in internal combustion engine components. Int. Mater. Rev. 2014, 59,
132–158. [CrossRef]
5. Grieb, M.; Christ, H.-J.; Plege, B. Thermomechanical fatigue of cast aluminium alloys for cylinder head applications–experimental
characterization and life prediction. Procedia Eng. 2010, 2, 1767–1776. [CrossRef]
6. Zhang, S.; Wang, Z.; Han, Y.; Zheng, Y.; Zhang, D. Experimental and theoretical studies on thermo-mechanical fatigue test for
aluminium cast alloy. Fatigue Fract. Eng. Mater. Struct. 2020, 43, 110–118. [CrossRef]
7. Mattos, J.; Uehara, A.; Sato, M.; Ferreira, I. Fatigue properties and micromechanism of fracture of an alsimg0.6 cast alloy used in
diesel engine cylinder head. Procedia Eng. 2010, 2, 759–765. [CrossRef]
8. Azadi, M. Effects of strain rate and mean strain on cyclic behavior of aluminum alloys under isothermal and thermo-mechanical
fatigue loadings. Int. J. Fatigue 2013, 47, 148–153. [CrossRef]
9. Azadi, M.; Shirazabad, M. Heat treatment effect on thermo-mechanical fatigue and low cycle fatigue behaviors of A356.0
aluminum alloy. Mater. Des. 2013, 45, 279–285. [CrossRef]
10. Engler-Pinto, C.; Sehitoglu, H.; Maier, H.; Foglesong, T. Thermo-Mechanical Fatigue Behavior of Cast 319 Aluminum Alloys. In
European Structural Integrity Society; Rémy, L., Petit, J., Eds.; Elsevier: Amsterdam, The Netherlands, 2002; pp. 3–13.
11. Huter, P.; Oberfrank, S.; Grün, F.; Stauder, B. Thermo-mechanical fatigue influence of copper and silicon on hypo-eutectic
Al–Si–Cu and Al–Si–Mg cast alloys used in cylinder heads. Int. J. Fatigue 2016, 88, 142–155. [CrossRef]
12. May, A.; Belouchrani, M.; Taharboucht, S.; Boudras, A. Influence of heat treatment on the fatigue behaviour of two aluminium
alloys 2024 and 2024 plated. Procedia Eng. 2010, 2, 1795–1804. [CrossRef]
13. Toda, H.; Fukunaga, T.; Kobayashi, M. Improvement of thermomechanical fatigue life in an age-hardened aluminum alloy. Scr.
Mater. 2009, 60, 385–387. [CrossRef]
14. Sehitoglu, H.; Emgler-Pinto, C.; Maier, H.; Foglesong, T. Thermomechanical deformation of AL319 Alloys with different iron
contents. In Proceedings of the CAMP 2002—High Temperature Fatigue, Bad Lippspringe, Germany, 3–4 April 2002; pp. 76–83.
15. Wang, Q.; Apelian, D.; Lados, D. Fatigue behavior of A356-T6 aluminum cast alloys. Part I. Effect of casting defects. J. Light Met.
2001, 1, 73–84. [CrossRef]
16. Firouzdor, V.; Rajabi, M.; Nejati, E.; Khomamizadeh, F. Effect of microstructural constituents on the thermal fatigue life of A319
aluminum alloy. Mater. Sci. Eng. A 2007, 454–455, 528–535. [CrossRef]
17. Wang, M.; Pang, J.; Liu, X.; Wang, J.; Liu, Y.; Li, S.; Zhang, Z. Optimization of Thermo-Mechanical Fatigue Life for Eutectic
Al&ndash;Si Alloy by the Ultrasonic Melt Treatment. Materials 2022, 15, 7113.
18. Moizumi, K.; Mine, K.; Tezuka, H.; Sato, T. Influence of Precipitate Microstructures on Thermal Fatigue Properties of Al-Si-Mg
Cast Alloys. Mater. Sci. Forum 2002, 396–402, 1371–1376. [CrossRef]
19. Beck, T.; Löhe, D.; Luft, J.; Henne, I. Damage mechanisms of cast Al–Si–Mg alloys under superimposed thermal–mechanical
fatigue and high-cycle fatigue loading. Mater. Sci. Eng. A 2007, 468–470, 184–192. [CrossRef]
20. Toyoda, M.; Toda, H.; Ikuno, H.; Kobayashi, T.; Kobayashi, M.; Matsuda, K. Preferential orientation of precipitates during
thermomechanical cyclic loading in an aluminum alloy. Scr. Mater. 2007, 56, 377–380. [CrossRef]
21. Tabibian, S.; Charkaluk, E.; Constantinescu, A.; Szmytka, F.; Oudin, A. TMF–LCF life assessment of a Lost Foam Casting A319
aluminum alloy. Int. J. Fatigue 2013, 53, 75–81. [CrossRef]
22. Tsuyoshi, T.; Sasaki, K. Low cycle thermal fatigue of aluminum alloy cylinder head in consideration of changing metrology
microstructure. Procedia Eng. 2010, 2, 767–776. [CrossRef]
23. Qin, J.; Racine, D.; Liu, K.; Chen, X.-G. Strain-controlled thermo-mechanical fatigue testing of aluminum alloys using the gleeble
3800 system. In Proceedings of the 16th International Aluminum Alloys Conference (ICAA 16), Canadian Institute of Mining,
Metallurgy & Petroleum, Montreal, QC, Canada, 17–21 June 2018; pp. 1–9, Paper no. 401895.
24. Chen, S.; Liu, K.; Chen, X. Precipitation behavior of dispersoids and elevated-temperature properties in Al–Si–Mg foundry alloy
with Mo addition. J. Mater. Res. 2019, 34, 3071–3081. [CrossRef]
25. Jin, L.; Liu, K.; Chen, X. Evolution of dispersoids and their effects on elevated-temperature strength and creep resistance in
Al-Si-Cu 319 cast alloys with Mn and Mo additions. Mater. Sci. Eng. A 2020, 770, 138554. [CrossRef]

294
Materials 2023, 16, 829

26. Farkoosh, A.; Chen, X.G.; Pekguleryuz, M. Dispersoid strengthening of a high temperature Al–Si–Cu–Mg alloy via Mo addition.
Mater. Sci. Eng. A 2015, 620, 181–189. [CrossRef]
27. Liu, K.; Mirza, F.; Chen, X. Effect of Overaging on the Cyclic Deformation Behavior of an AA6061 Aluminum Alloy. Metals
2018, 8, 528. [CrossRef]
28. Baldan, A. Review Progress in Ostwald ripening theories and their applications to nickel-base superalloys Part I: Ostwald
ripening theories. J. Mater. Sci. 2002, 37, 2171–2202. [CrossRef]
29. Chen, Y.; Doherty, R. On the growth kinetics of plate-shaped precipitates in aluminium-copper and aluminium-gold alloys. Scr.
Metall. 1977, 11, 725–729. [CrossRef]
30. Liu, G.; Zhang, G.; Ding, X.; Sun, J.; Chen, K. Modeling the strengthening response to aging process of heat-treatable aluminum
alloys containing plate/disc- or rod/needle-shaped precipitates. Mater. Sci. Eng. A 2003, 344, 113–124. [CrossRef]
31. Nakajima, T.; Takeda, M.; Endo, T. Accelerated coarsening of precipitates in crept Al–Cu alloys. Mater. Sci. Eng. A 2004, 387–389,
670–673. [CrossRef]
32. Yassar, R.; Field, D.; Weiland, H. The effect of cold deformation on the kinetics of the β” precipitates in an Al-Mg-Si alloy. Metall.
Mat. Trans. A 2005, 36, 2059–2065. [CrossRef]
33. Neu, R.; Sehitoglu, H. Thermomechanical fatigue, oxidation, and Creep: Part II. Life prediction. MTA 1989, 20, 1769–1783.
[CrossRef]
34. Miller, M.; McDowell, D.; Oehmke, R. A Creep-Fatigue-Oxidation Microcrack Propagation Model for Thermomechanical Fatigue.
J. Eng. Mater. Technol. 1992, 114, 282–288. [CrossRef]
35. Eichlseder, W.; Winter, G.; Farrahi, G.; Azadi, M. The Effect of Various Parameters on Out-of-phase Thermo-mechanical Fatigue
Lifetime of A356.0 Cast Aluminum Alloy. Int. J. Eng. 2013, 26, 1461–1470.
36. Wang, M.; Pang, J.; Zhang, M.; Liu, H.; Li, S.; Zhang, Z. Thermo-mechanical fatigue behavior and life prediction of the Al-Si
piston alloy. Mater. Sci. Eng. A 2018, 715, 62–72. [CrossRef]
37. Gocmez, T.; Awarke, A.; Pischinger, S. A new low cycle fatigue criterion for isothermal and out-of-phase thermomechanical
loading. Int. J. Fatigue 2010, 32, 769–779. [CrossRef]
38. Riedler, M.; Leitner, H.; Prillhofer, B.; Winter, G.; Eichlseder, W. Lifetime simulation of thermo-mechanically loaded components.
Meccanica 2007, 42, 47–59. [CrossRef]
39. Wang, M.; Pang, J.; Liu, H.; Li, S.; Zhang, M.; Zhang, Z. Effect of constraint factor on the thermo-mechanical fatigue behavior of
an Al-Si eutectic alloy. Mater. Sci. Eng. A 2020, 783, 139279. [CrossRef]
40. Farrahi, G.; Azadi, M.; Winter, G.; Eichlseder, W. A new energy-based isothermal and thermo-mechanical fatigue lifetime
prediction model for aluminium–silicon–magnesium alloy. Fatigue Fract. Eng. Mater. Struct. 2013, 36, 1323–1335. [CrossRef]

Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.

295
materials
Article
Mechanical Behavior of Austenitic Steel under Multi-Axial
Cyclic Loading
Abhishek Biswas 1,† , Dzhem Kurtulan 2 , Timothy Ngeru 2 , Abril Azócar Guzmán 1,‡ , Stefanie Hanke 2
and Alexander Hartmaier 1, *

1 Interdisciplinary Centre for Advanced Materials Simulation (ICAMS), Ruhr-Universität Bochum,


Universitätstraße 150, 44801 Bochum, Germany
2 Chair of Materials Science and Engineering, Universität Duisburg-Essen, 47057 Duisburg, Germany
* Correspondence: [email protected]
† Current address: VTT Technical Research Centre of Finland Ltd., Vuorimiehentie 2, FI-02150 Espoo, Finland.
‡ Current address: Institute for Advanced Simulations, Forschungszentrum Jülich GmbH, 42525 Jülich, Germany.

Abstract: Low-nickel austenitic steel is subjected to high-pressure torsion fatigue (HPTF) loading,
where a constant axial compression is overlaid with a cyclic torsion. The focus of this work lies on
investigating whether isotropic J2 plasticity or crystal plasticity can describe the mechanical behavior
during HPTF loading, particularly focusing on the axial creep deformation seen in the experiment.
The results indicate that a J2 plasticity model with an associated flow rule fails to describe the axial
creep behavior. In contrast, a micromechanical model based on an empirical crystal plasticity law
with kinematic hardening described by the Ohno–Wang rule can match the HPTF experiments
quite accurately. Hence, our results confirm the versatility of crystal plasticity in combination with
microstructural models to describe the mechanical behavior of materials under reversing multiaxial
loading situations.

Keywords: multi-axial fatigue; crystal plasticity; micromechanical modeling; austenitic steel

Citation: Biswas, A.; Kurtulan, D.;


1. Introduction
Ngeru, T.; Azócar Guzmán, A.;
Hanke, S.; Hartmaier, A. Mechanical Austenitic stainless steels, like 316L and Rex 734, exhibit remarkable corrosion resis-
Behavior of Austenitic Steel under tance and formability, which makes them suitable for medical implants. However, the
Multi-Axial Cyclic Loading. Materials leaching of alloying metals like chromium, cobalt, and nickel ion in the human body en-
2023, 16, 1367. https://doi.org/ vironment can cause a strong allergic response and is even found to be carcinogenic in
10.3390/ma16041367 rats [1–3]. This led to the development of a low nickel and high nitrogen austenitic stain-
less steel 1.3808 with the commercial name P558, which exhibits remarkable mechanical
Academic Editor: Carlos
Garcia-Mateo
properties and good bio-compatibility [4–7] with targeted applications in medical implants.
In view of the excellent properties of P558, many advanced manufacturing techniques like
Received: 10 December 2022 liquid-phase sintering [8] and metal powder injection molding [9] have been explored to
Revised: 31 January 2023 manufacture complex components. However, the existing research of austenitic stainless
Accepted: 2 February 2023 steel mechanical properties is mainly focused on stainless steel grades, like 316L or 304L,
Published: 6 February 2023 and few publications report even very basic mechanical properties like yield stress, surface
roughness, or hardness of P558 [9,10].
The understanding of material behavior under multiaxial mechanical loading is of
great importance for the application of structural materials because components are of-
Copyright: © 2023 by the authors.
tentimes exposed to multiaxial stress states. In addition, the loading is typically time-
Licensee MDPI, Basel, Switzerland.
This article is an open access article
dependent and reversing. Experimental investigations under such complex loading con-
distributed under the terms and
ditions are very laborious, such that our understanding of the mechanisms leading to
conditions of the Creative Commons plastic deformation and failure of materials under multiaxial reversing loading is quite
Attribution (CC BY) license (https:// insufficient. This is particularly important for a newly developed steel, like P558, and the
creativecommons.org/licenses/by/ present work aims at closing this gap in our understanding of the mechanical behavior of
4.0/). this new austenitic steel under complex mechanical loading.

Materials 2023, 16, 1367. https://doi.org/10.3390/ma16041367 297 https://www.mdpi.com/journal/materials


Materials 2023, 16, 1367

To investigate the kinematic hardening behavior of P558 austenitic stainless steel under
multiaxial reversing loading, the steel specimens are subjected to combined compression
and cyclic torsional load. In preliminary investigations, it has been observed experimentally
that, under combined compression and cyclic torsional loading, the specimens undergo
plastic axial strain although the compressive load is well below the yield stress of the
material. This time-dependent deformation under small stresses can be considered as
room-temperature creep behavior, triggered by the multiaxial loading. This compressive
creep occurs as soon as a critical angle for the cyclic torsion is exceeded, such that it
remains unclear whether an equivalent stress of the applied stress tensor can be defined
that describes this deformation mode with sufficient accuracy. Ngeru et al. [11] reported
similar behavior in a high nitrogen stainless steel subjected to similar loading conditions. It
is noted here, that this kind of creep deformation is purely mediated by dislocation motion,
whereas diffusive or other thermally activated mechanisms do not play a role. The term
creep is merely used here to describe time-dependent plastic deformation occurring at a
constant stress below the yield stress of the material.
Numerical simulations of the mechanical behavior of metals under cyclic loading
have been reported frequently in the literature, e.g., Cruzado et al. [12] performed crystal
plasticity finite element model (CPFEM) simulations of a polycrystalline microstructure
and showed a good agreement between simulation and experiments. In other work,
the effects of monotonic-cyclic loading on the behavior of high-alloyed steel and pure
copper were investigated and an axial stress-drop was reported for combined cyclic torsion
and monotonic axial tensile loading [13].
To incorporate kinematic hardening in constitutive models, mathematical descriptions
are required that can predict both the stress–strain hysteresis as well as the relaxation
of the mean stress that occurs under cyclic loading [14]. During strain-controlled cyclic
loading, the mean stress is defined as the average between the maximum and the minimum
stress resulting from the cyclic material response. If the upper and the lower value of
the applied strain are not equal with respect to their absolute value, i.e. for asymmetric
loading with amplitude ratios different from R = −1, the mean stress typically takes a
rather high positive or negative values at the beginning of the cyclic experiment, but then
it usually relaxes towards zero. The importance of the correct description of kinematic
hardening in constitutive models for multiaxial cyclic loading has been demonstrated for
investigations on thin-walled tubular specimens under axial tension, and cyclic torsion [15].
To describe kinematic hardening in general, the empirical kinematic hardening models
of Armstrong–Fredrick [16], Chaboche [17], and Ohno–Wang [18] are most widely used.
Several works implying J2 plasticity [18,19] or crystal plasticity [14] have noted that the
Armstrong–Fredrick kinematic hardening model failed to predict the mean stress relaxation
that occurs under non-symmetric loading conditions, where the amplitude ratio R
= −1.
Schäfer et al. [20] showed a comparison between different kinematic hardening models in
CPFEM with loading cases having both symmetric loading R = −1 and non-symmetric
loading R = 0. The results indicate that the Ohno–Wang model is most adept at predicting
cyclic mechanical properties. Sajjad et al. [21] used both, J2 as well as CPFEM, with the
Chaboche kinematic hardening model for predicting the hysteresis loop and the mean
stress relaxation under uniaxial cyclic loading. Based on the aforementioned publications,
in this work, the Chaboche kinematic hardening model is used with J2 plasticity, and the
Ohno–Wang kinematic hardening model is selected for the CPFEM simulations due to their
ability to predict complex cyclic properties for different loading cases and R-values.
The outline of the present work is as follows: Section 2 describes the experimental
setup for the EBSD and fatigue experiments. Section 3 explains the modeling strategy
for the estimation of elastic stiffness constants, the micromechanical modeling based on
experimental input, and the material models used to mimic the material behavior in the nu-
merical model. Section 4 shows the comparison of the results from the numerical model and
experiments and provides a discussion of the results and observations. Finally, in Section 5,

298
Materials 2023, 16, 1367

the conclusions of this work are laid out. To support the reading of the mathematical
equations, a list of variables/symbols is provided in the Nomenclature section.

2. Material Characterization
2.1. Specimen Preparation and Setup of EBSD
P558 is the commercial name for the austenitic stainless steel 1.3808 sold by voestalpine
BÖHLER Edelstahl GmbH & Co KG (Kapfenberg, Austria). The raw material was delivered
in the form of cylindrical bars from which the specimens were produced as specified below.
The chemical analysis of the tested material is given in Table 1. Due to the absence of nickel
in the alloy, a combination of excellent mechanical properties and high corrosion resistance
is achieved, as well as good bio-compatibility. To bring the material into a well-defined state
and to remove all cold-deformation from the sample manufacturing process, the material
was solution annealed at 1150 ◦ C for 45 min. After that, tensile tests of the material resulted
in an ultimate tensile strength of 937 MPa and a yield stress of 595 MPa, where the results
of three standard tensile tests have been averaged.

Table 1. Chemical composition of P558.

C Si Mn P S Cr Ni Mo Cu V Nb N
0.19 0.41 12.20 0.01 0.00 17.15 0.02 2.89 0.02 0.01 0.03 0.54

For a microscopic analysis, specimens from bar material were cut both in longitudinal
and cross-sectional directions according to Figure 1. Consequently, the material was
embedded with conductive resin Technotherm 3000 (Kulzer GmbH, Hanau, Germany),
ground up to 800 grit, and polished with silica on RotoPol-31 (Struers GmbH, Willich,
Germany). The final preparation step for EBSD was vibration polishing of specimens at a
frequency of 90 Hz for 3 h on Saphir Vibro (ATM Qness GmbH, Mammelzen, Germany).
Longitudinal cut Cross-sectional cut

Measuring area Measuring area

Figure 1. Schematic view of the cutting directions of the specimens for EBSD Analysis.

A Field Emission Gun Scanning Electron Microscope (LEO Gemini 1530 from Carl
Zeiss AG) was used to capture micrographs with the electron backscatter detector (EBSD)
at a voltage of 20 kV and an aperture size of 120 μm. The working distance was 15 mm, and
the spot size varied between 0.5 μm and 2 μm. The EBSD data sets were initially analyzed
with the help of the analysis software OIM 6.2 from EDAX (Mahwah, NJ, USA), due to the
large number of EBSD data sets, MATLAB scripting (with MTEX toolbox) was used for the
post-processing of data sets.

2.2. Microstructure Analysis Using EBSD


In this section, we describe the microstructural features observed in the EBSD data
sets, which provide the key inputs for the generation of the virtual microstructure used for

299
Materials 2023, 16, 1367

CPFEM simulations. To properly characterize the microstructure across the bulk material,
23 EBSD measurements were performed at different locations of the longitudinal and
cross-sectional cuts and for different magnifications, resulting in the characterization of
≈19,000 grains in the underlying microstructure.
Figure 2 shows the (Inverse pole figure) IPF-Z plot of one representative EBSD map and
the contour plot of the pole figure of the orientation distribution function (ODF) estimated
by combining 23 EBSD data sets. The ODF is estimated using a kernel density estimate [22].
Ng
f (q) = ∑ wi ψΩ (qi ) (1)
i =1

where the function f is the ODF, q is the crystallographic orientation, Ng is the number
of grains, and wi = ai / ∑ ai is the weighting factor for each grain estimated considering
the area ai of the grain. The grains are estimated using a disorientation threshold of 5◦ ,
and to avoid experimental/estimation errors the grains smaller than 3 μm are discarded.
The grain size is represented
 using the equivalent diameter, which is the diameter of a circle
with the same area (= 2 ( ai /π )).

100m

Figure 2. (top left) IPF-Z plot of one of the EBSD data sets, (bottom left) corresponding IPF key,
(top right) pole figure contour plot of the ODF estimated by combining all EBSD data sets (i.e.,
≈19,000 grains), (bottom right) the grain equivalent diameter histogram.

The microstructure exhibits mostly equiaxed grains containing annealing twins, which
are formed due to the re-crystallization during the annealing process and which show a
density that is proportional to the strain level achieved before the annealing process [23].
Usually, the annealing twins are of Σ3 type with both a symmetrical and asymmetrical
grain boundary tilt [24].

300
Materials 2023, 16, 1367

2.3. Fatigue Test


Both uniaxial and high-pressure torsion fatigue (HPTF) experiments were conducted
on cylindrical double-cone specimens, presented in Figure 3a,b. While uniaxial fatigue
specimens have a shorter gauge length, the HPTF specimens possess a thicker gauge
section to prevent the buckling of the specimens due to the multiaxial loading case. Before
experiments, the gauge section of all specimens was ground with increasingly finer grades
of sandpaper and polished with 1 μm diamond paste to remove turning marks. After
polishing, the arithmetic mean roughness (Ra ) of the specimen surface was 46 ± 6 nm.
A Bionix 858 (MTS Corporation, Eden Prairie, MN, USA) servo-hydraulic universal
testing machine was used for all fatigue tests. With the stated testing setup, the specimen
can be loaded with 50 kN of maximum axial force and 300 Nm of maximum torsion
moment. Uniaxial fatigue tests were performed at 20 ± 2 ◦ C with 40–60% relative humidity
under total strain-controlled tension-compression with sinusoidal loading at R = −1. The
frequency of the experiments was 2 Hz. All experiments were conducted until failure or
run out at 2 × 106 cycles.
The frequency of experiments was also controlled, limiting the temperature increase
of specimens to 50 K for uniaxial tests (2.5 Hz for  < 1.5% and 0.5 Hz for  = 1.5%).
During the HPTF tests, however, the temperature increase may reach values over 200 K
and saturated after 400–500 cycles.
Multiaxial loading of the specimens was rotation angle controlled torsion (0◦ to 5–20◦ )
with sinusoidal loading at R = 0 or −1 and a frequency of 2.5 Hz, combined with a
monotonous constant compressive stress 250 or 350 MPa. Figure 3 visualizes the special
multiaxial loading of the specimen during experiments. The axial force necessary to apply the
stated axial stress is related to the smallest cross-section of specimens, measured with a laser
thickness gauge with a precision of 10 μm. The experiments are terminated with the observa-
tion of clearly visible cracks on the surface of the specimen. During the HPTF experiments, the
specimens exhibit compressive axial strains under constant load below the yield stress, which
is referred to as axial creep deformation in this work.
To have a proper database for the modeling of the material behavior under HPTF
loading for different R values and strains, three different loading conditions were selected:
(R = −1, θ = 7.5◦ ), (R = 0, θ = 10◦ ), and (R = 0, θ = 15◦ ) where θ is the amplitude of the
cyclic torsion. Figure 4 shows the experimental results for the stress–strain hysteresis loops
after the saturation of the mean stress and the axial creep for the selected HPTF loading
cases. The equivalent true strain shown in Figure 4 is estimated using

true = log (1 + eng ), (2)

where the engineering strain eng for torsion is defined as


rg
eng = θ , (3)
lg

with the radius r g and the length l g of the gauge section. This equivalent true strain,
thus, corresponds to the strain reached at the surface of the gauge section. Similarly,
the engineering shear stress τeng at the surface of the gauge section is defined in a linear
approximation as
rg
τ eng = T , (4)
J
where J = 0.5πr4g is the polar moment of inertia for the circular cross-section and T is the
experimentally measured torque. The general expression for true stress is defined as

σtrue = σeng (1 + eng ). (5)

301
Materials 2023, 16, 1367

(a)
SectionB-B
Secon B-B

Section A-A
(b)
A

(c)
Force [N]

Angle [°]
Axial

Cycle Cycle

Figure 3. Figure showing the dimensions and surface finish for (a) HPTF specimen, (b) uniaxial
fatigue specimen, and (c) HPTF loading of the specimen, indicating the constant axial force and the
cyclic torsional angles.

Figure 4. (left) Experimental stress–strain hysteresis loops recorded for uniaxial fatigue after the
saturation of the mean stress and (right) axial creep for the HPTF loading cases given in the legend,
where θ is the amplitude of the cyclic torsion loading, and R is the ratio of the minimum vs. maximum
torsion angle.

Table 2 summarizes the details of the fatigue experiments performed in this work,
where θ denotes the amplitude of torsion cycles for HPTF tests and E33 is the loading
amplitude in the axial direction (z-axis) for uniaxial fatigue tests.

302
Materials 2023, 16, 1367

Table 2. Details of uniaxial fatigue and high pressure torsion fatigue (HPTF) experiments. Axial
stress and strain components are denoted by indices “33”.

Type R Amplitude Frequency Axial Pressure


Uniaxial −1 E33 = 0.55% 2.5 Hz -
HPTF −1 θ = 7.5◦ 2.5 Hz σ33 = 250 MPa
HPTF 0 θ = 10◦ 2.5 Hz σ33 = 250 MPa
HPTF 0 θ = 15◦ 2.5 Hz σ33 = 250 MPa

3. Micromechanical Model
In this section, the workflow of the micromechanical modeling is described. Due to the
recent development of the P558 austenitic stainless steel, the stiffness constants essential
for numerical modeling are unavailable in the published literature. Therefore, density
functional theory (DFT) calculations are employed to estimate the elastic constants, as
detailed in Appendix B. To incorporate the influence of various microstructural features
such as inclusions, defects, etc., the ideal stiffness tensor estimated by the aforementioned
DFT method using a pristine molecular structure is scaled by a scalar constant λ, such that
Ceff = λCDFT ). The parameter λ is calibrated to match the elastic part of the experimental
hysteresis loop with corresponding FE simulations. Ceff is used to describe anisotropic
elastic behavior in CPFEM and, in the case of the J2 model, isotropic elastic constants are
determined by homogenizing the elastic stiffness tensor according to the crystallographic
texture as quantified by f (q) [25]. Table 3 shows the ideal elastic stiffness parameters
obtained with DFT calculations, homogenized Young’s modulus (Y) (without scaling), and
the scaling factor λ. The values obtained from DFT simulations are comparable to the
elastic modulus of 155–220 GPa reported for similar nickel-free stainless steels [26,27].

Table 3. Ideal elastic stiffness parameters C11 , C12 , and C44 obtained by DFT calculations, resulting
homogenized Young’s modulus Y, and scaling factor λ.

C11 C12 C44 Y λ


(GPa) (GPa) (GPa) (GPa) (-)
263.159 122.644 76.506 185.184 0.63

The microstructure characterization of the present material with EBSD reveals the
presence of annealing twins. Castelluccio and McDowell [28] indicate that the annealing
twins can influence the fatigue crack initiation. Furthermore, Pande et al. [29] demonstrate
that size effects can occur due to annealing twins in the microstructure. However, this
work focuses on cyclic property prediction rather than aiming at predicting fatigue failure.
Therefore, for the sake of simplicity, annealing twins in the micromechanical model and
application of gradient-based plasticity models to describe size effects are disregarded in
the present work.

3.1. Representative Volume Element


Due to the high computational effort required for CPFEM simulations with kinematic
hardening, a simple representative volume element (RVE) is selected in which each grain is
represented by a cubic shape. To optimize the number of elements in the model, a parametric
study of the RVE focusing on the number of elements used for the discretization of each
grain is performed. This study indicates that an RVE with 512 grains and 8 elements per
grain, see Figure 5, provides the optimum balance between mechanical property prediction
and computational effort. The edge length of each cubic grain is taken as 0.05 mm, which is
estimated as the mean equivalent grain diameter from the EBSD analysis, see Figure 2.
The required number of 512 discrete orientations for the grains constituting the RVE
are systematically sampled from the EBSD data for a total of 19,000 grains, using the method
in Biswas et al. [30]. The results of this sampling method are shown in Figure 5, and the

303
Materials 2023, 16, 1367

quality of the sampling is judged using the error function f (q) − fˆ(q) 1 , where · 1 is
the L1 norm of the difference of experimental ODF f (q) and the reduced ODF fˆ(q) used
for the RVE simulations. It is seen that sampling of 512 discrete orientations is achieved
within an error margin of 10%. Since this work focuses on the macroscopic properties, the
grain boundary disorientation distribution is neglected [31].

0.04

Figure 5. (top left) Graphical representation of the used RVE with 512 cubic grains and 8 finite elements
per grain, (bottom left) grains are colored according to their orientation and the corresponding IPF key.
Contour plots of the ODF pole figures from the orientation set obtained by combining all the EBSD
data sets of about 19,000 grains (top right) and 512 discrete orientations in the RVE (bottom right).

3.2. Material Model


3.2.1. J2/Von Mises Plasticity
In this section, the J2 model with the isotropic hardening/softening and the Chaboche
kinematic hardening model [17] are described briefly. For a detailed description please
refer to ABAQUS documentation [32]. The yield function
'
3 dev
f (σ ) = (σ − κdev ) : (σ dev − κdev ) − σm , (6)
2
depends on the stress tensor σ and incorporates kinematic hardening via the back-stress κ.
The scalar yield stress of the material is denoted as σm , and the superscript “dev” indicates
that the corresponding deviatoric tensor is addressed. Plastic yielding sets in when the
yield function vanishes, i.e., f (σ ) = 0 is used as the yield criterion. If the value of the yield
function is larger than zero, a plastic strain increment is calculated that reduces the elastic
strain and, thus, relaxes the stress to a value where the yield function is again zero. In this
so-called return mapping algorithm, the plastic strain increments are calculated as

∂f 3 σ dev − κdev
dpl = dλ = dλ , (7)
∂σ 2
2 (σ − κdev ) : (σ dev − κdev )
3 dev

where the direction of the plastic strain is given by the gradient of the yield function with
respect to the stress tensor and the magnitude of the plastic strain is quantified by the
plastic strain multiplier dλ. It is noted here, that for this associated flow rule the direction
of the plastic strain increment corresponds to the deviatoric stress tensor reduced by the
back stress tensor. For details, the reader is referred to the basic literature on continuum
plasticity, e.g. the textbook of de Borst et al. [33].

304
Materials 2023, 16, 1367

The isotropic hardening/softening is modeled using an exponential law as

σm = σm + Q(1 − e−beq ), (8)


where Q and b are material
parameters that can be estimated by matching experimental
results, and eq = (2/3)pl : pl is the equivalent plastic strain. The evolution of κ is
given by

1
κ̇i = Ci ˙ eq
(σ − κi ) − gi ˙ eq κi , (9)
σm
where Ci , gi are the material parameters, and ˙ eq is the equivalent plastic strain rate. The
index i ∈ {1, 2} addresses two different back-stresses terms considered here. The total
back-stress is computed as κ = ∑i κi .

3.2.2. Crystal Plasticity


This section presents a brief description of the crystal plasticity model used in FE
simulations of the mechanical behavior of the RVE. For the sake of brevity, we mainly focus
on the modeling of kinematic hardening, please refer to [34] and references therein for a
detailed description of the CPFEM method. The total deformation gradient is described
using multiplicative decomposition F = Fe Fp , where Fe is the elastic part representing
a reversible lattice deformation and rotation, and Fp is the plastic part representing an
irreversible lattice deformation.
The plastic deformation is assumed to result from the motion of dislocations on
crystallographic slip systems described by the slip direction dα and slip plane normal
nα where α = 1, ..., Ns is the slip system index, and Ns is the number of slip systems.
The driving force for the motion of dislocations is given by the resolved shear stress (τ α ),
which is calculated by mapping the second Piola-Kirchhoff stress tensor S = C2eff (Fe T Fe − I)
on the slip system α. I is second rank unit tensor.
In this work, a phenomenological constitutive model is used, in which the kinematic
hardening is modeled by including an additional back-stress τbα in the flow rule
 α 
 τ − τbα  p1
γ̇α = γ̇0   sign(τ α − τ α ) (10)
τα 
c
b

where γ̇α is the shearing rate, τcα is the slip resistance for isotropic hardening. In this work,
the Ohno–Wang model is used, which is adapted from the original formulation to the
crystal plasticity model by defining the evolution of the back-stress as

|τbα | α α
τ̇bα = η γ̇α − μ τ |γ̇ | (11)
(η/μ)m b
following the work of McDowell [35]. Hennessey et al. [14] used two components for τbα
using sets of {ηi , μi , mi } with (i = 1, 2). For the sake of simplicity, in this work τbα is limited
to a single component with only one set of {η, μ, m} as parameters. The isotropic hardening
is given by the evolution of the slip resistance, described as

Ns
β p2
τc
τ̇cα = ∑ h0 ξ αβ 1− f
|γ̇ β | (12)
β =1 τc
where τcα is the saturation slip resistance and ξ αβ is the cross-hardening matrix in which the
diagonal elements representing the coplanar slip systems are set to 1.0, and off-diagonal
elements representing the non-coplanar slip systems are set to 1.4.

4. Results and Discussion


In this section, the results obtained from numerical models are compared with ex-
perimental findings. Due to the multiaxial and heterogeneous stress and strain fields in

305
Materials 2023, 16, 1367

the gauge section of the specimens, the material parameters cannot be assessed directly
from the results of the HPTF test, but the constitutive parameters for J2 and CPFE model
must be obtained using an inverse procedure instead, as shown in Appendix C. In the
first step, a FE model of the gauge section of the test specimens is applied to characterize
the material response under HPTF loading using a J2/von Mises plasticity model with
Chaboche kinematic hardening and isotropic softening. In the second step, a representative
volume element (RVE) is introduced that considers the mean grain size and crystallographic
texture resulting from the EBSD analysis of the tested specimens, as described above. In FE
simulations of this RVE, boundary conditions are applied that mimic the HPTF loading on a
volume element close to the surface of the gauge section of the experimental specimens. The
plastic deformation of each grain in the RVE is described by the CP model introduced above.
It is noted here that in the present study, we do not attempt to model the grain shapes or
even microstructural details as annealing twins, but we focus on the determination of the
material parameters, which requires the use of a numerically efficient model. Furthermore,
a parametric study of the CP kinematic hardening parameters is shown in Appendix A,
emphasizing the importance of axial creep experiments for the calibration of parameters.

4.1. J2 Plasticity
The parameters for the J2 model are calibrated using data from uniaxial fatigue
experiments. The parameters controlling the stress–strain hysteresis loop, i.e., C1 , C2 , g1 ,
and g2 , are obtained in an iterative process in which the difference between experimental
and numerical hysteresis loops are minimized by systematically varying those material
parameters. For this purpose, a representative hysteresis loop is chosen from the regime
of load cycles in which the mean stress has already saturated. In contrast, the isotropic
softening parameters, i.e., σm , Q, and b, are gained from the comparison of the maximum
stress obtained during each load cycle between simulation and experiment. To obtain the
hysteresis loop in the regime where the amplitude is constant, the experimental hysteresis
loop from the 5000th cycle is used for calibration of the kinematic hardening parameters.
The isotropic softening parameters are calibrated by simulating the initial 300 cycles.
Figure 6 shows the comparison of the saturated hysteresis, i.e., the hysteresis loop in
the regime where the stress amplitude is constant, and the maximum stress obtained from
simulation and experiments; the corresponding calibrated J2 material parameters are given
in Table 4.

Figure 6. (left) Stress–Strain hysteresis loop and (right) reduction of the maximum stress in uniaxial
fatigue tests with a strain amplitude of E33 = 0.55% and an amplitude ratio of R = −1.

Table 4. Calibrated J2 model parameters.

C1 C2 g1 g2 σm Q b
(MPa) (MPa) (-) (-) (MPa) (MPa) (-)
43,106.44 4192.81 513.95 0.0 545.82 −300 3.17

306
Materials 2023, 16, 1367

Keeping the J2 material parameters constant, HPTF simulations are performed with
the same FE model, see Figure 7, in which the resulting axial creep strain E33 obtained from
the experiment and the HPTF simulations also are shown. It is seen that the simulations
result in a similar axial creep behavior under different HPTF loading conditions, which
stands in stark contrast to the experiment where the axial creep behavior is very sensitive to
the torsional loading. Furthermore, the axial creep strains obtained from the FE simulations
with material parameters fitted to uniaxial fatigue experiments grossly overestimate the
experimentally found creep strains under HPTF loading, which points out the limitations
of the J2 plasticity model. The reason for these deficits in the J2 model in predicting the
axial creep behavior correctly can be attributed to the associated flow rule used here, see
Equation (7), where the plastic strain increment is proportional to the deviatoric stress
reduced by the back stress. Hence, as long as constant stress is applied along the axial
direction, the plastic strain in this direction will not be sensitive to the amplitude of the
cyclic torsional load, which is clearly seen in the simulation results for axial creep in
Figure 7. Since this is a fundamental restriction, no attempt is made to test other modeling
strategies for J2 plasticity, although in the literature, it has been demonstrated that an
associated flow rule in combination with a three-surface hardening-recovery model for
kinematic hardening could be successfully parameterized to describe the material behavior
of thin-walled tubular specimens under monotonic-cyclic loading [15].

33
(a)

(b) (c)
Figure 7. (top left) FE model used to simulate HTPF loading with a J2 plasticity model and resulting
axial creep strain (E33 ) of simulation and experiment vs. cycle number. The axial pressure amounted
to σ33 = 250 MPa and three different combinations of amplitude ratios R and cyclic torsion amplitudes
θ are shown: (a) R = −1, θ = 7.5◦ , (b) R = 0, θ = 10◦ , and (c) R = 0, θ = 15◦ .

To understand the distributions of axial stress and axial plastic strain in the J2 model,
these quantities are shown along a cut through the gauge section in Figure 8. The contour
plots indicate that both, axial stress and plastic strain, are rather constant in each cross-
section of the model. The axial stress, however, exhibits elevated compressive stresses close
to the symmetry axis, where high hydrostatic stress components are observed.

307
Materials 2023, 16, 1367

Figure 8. Contour plot of the axial stress in MPa (top) and the axial plastic strain (bottom) along a
cut through the J2 model.

4.2. Crystal Plasticity


Since the RVE with 512 grains is much smaller than the dimension of the experimental
specimens, we cannot directly apply torsional loading on this RVE. Instead, we consider
a small volume element close to the surface of the cylindrical FE model representing the
gauge section of the experimental specimens, which has been used for the J2 plasticity
calculations, see Figure 9. The distortions experienced by this volume element during the
different HPTF loading cases are applied as boundary conditions to the CPFEM model to
achieve a comparable loading situation on the length scale of the RVE. The major distortion
occurs in the X–Z and Y–Z shear directions (tensor components E13 and E23 , respectively).
The RVE is subject to displacement boundary conditions resulting in X–Z shear, and a
distributed force boundary condition in the Z direction, the magnitude of the force is
adjusted such that the homogenized stress in the RVE matches the required axial pressure
for the given HPTF load, and similarly, the shear displacement is adjusted to match the
strain obtained from the saturated hysteresis loop of the HPTF experiment. Furthermore, as
shown in the J2 model, the axial plastic strain and the axial stress are rather homogeneous
throughout the gauge section, which justifies the application of the distortions of a volume
element close to the surface as boundary conditions to the RVE, as the plastic strains
observed in the RVE will be representative for the entire gauge section.
It is noted here, that the RVE is exposed to periodic boundary conditions to avoid
free surfaces, which would severely influence the stress field within the RVE and produce

308
Materials 2023, 16, 1367

finite size effects. This simplification is justified because we do not attempt to study crack
initiation, for which free surface effects might play a crucial role.

Ʌ
‫ܧ‬ଵଷ

ߪଷଷ

ߪଷଷ

ߪଷଷ ‫ܧ‬
ଵଷ

Figure 9. To mimic HPTF loading conditions on the length scale of the RVE used for CPFEM sim-
ulations, the distortions of a small volume element close to the surface of the cylindrical FE model,
representing the gauge section of the HPTF specimens, are applied as boundary conditions to the RVE.

In the first step to calibrate the CP parameters, the HPTF experiment with R = −1 and
θ = 7.5◦ is used. Only the kinematic hardening parameters η and μ can be calibrated in
f
this way, whereas the isotropic hardening parameters τc ,h0 , and p2 are temporarily set to
zero and will be calibrated in a second step as described below. To validate the correctness
and generality of the CP parameters obtained by this procedure, the calibrated model is
verified by comparing the simulation results with experiments for R = 0, θ = 10◦ and
R = 0, θ = 15◦ . Figure 10 shows that the stress–strain hysteresis loops obtained from
HPTF simulations using the RVE model closely match the experiments for all loading cases,
which confirms the predictive capability of the parameterized model.

309
Materials 2023, 16, 1367

(a)

(b) (c)
Figure 10. Comparison of saturated HPTF stress–strain hysteresis loops between simulation and
experiment; the constant axial pressure amounts to σ33 = 250 MPa in all cases and cyclic shear
loading E13 is applied to mimic the cyclic torsion in the experiment. (a) The result of calibration
for HPTF loading with R = −1, θ = 7.5◦ , E13 = 0.02. The two subfigures in the bottom represent
predictions of the model for (b) R = 0, θ = 10◦ , E13 = 0.028, and (c) R = 0, θ = 15◦ , E13 = 0.04.
f
The calibration of the isotropic hardening parameters (τc ,h0 , and p2 ) and the remaining
kinematic hardening parameter m, requires matching the evolution of maximum and
minimum stress recorded during all cycles between simulation and experiment. Again,
this is done for loading with (R = −1, θ = 7.5◦ ), and the result is validated by comparison
with the other cases. As seen in Figure 11, a satisfying agreement between simulation and
experiment could be achieved in this way. All resulting CP parameters are given in Table 5
and referred to as set 1.

Table 5. Calibrated CPFE model parameters (set 1).

f
λ γ̇0 p1 τ0 τc h0 p2 η μ m
(-) (1/s) (-) (MPa) (MPa) (MPa) (-) (MPa) (-) (-)
0.63 0.001 25 235 184.2 −25.0 2 16,888.0 125.0 2.0

As indicated in the parametric study shown in Appendix A, the axial creep behavior
is also employed for the calibration of the CP parameters. Figure 12 shows the axial creep
in the RVE during HPTF simulations. The results indicate that the CPFE model does not
only predict the cyclic properties but also the axial creep behavior for the different HPTF
loading cases very well.

310
Materials 2023, 16, 1367

(a)

(c) (b)
Figure 11. Comparison of maximum and minimum stress obtained in experiment and simulation
during torsional cycles with (a) R = −1, θ = 7.5◦ , comparison of cyclic softening behavior for the
loading cases (b) R = 0, θ = 10◦ , and (c) R = 0, θ = 15◦ .

(a)

(b) (c)
Figure 12. Axial creep strain (E33 ) vs. the number of cycles obtained from HPTF simulations with the
CPFE model and from experiment for a constant axial pressure of σ33 = 250 MPa and cyclic torsional
loading: (a) R = −1, θ = 7.5◦ ) used for calibration, (b) R = 0, θ = 10◦ , and (c) R = 0, θ = 15◦ . Cases
(b,c) represent predictions of the model.

311
Materials 2023, 16, 1367

In the next step, the CPFE model calibrated with HPTF experiments (set 1) is used to
predict material behavior under uniaxial fatigue loading. It is pointed out here that it is
expected that the material parameters fitted to more complex loading situations should be
able to describe the material behavior under uniaxial loading (see, for example, the study
of de Castro e Susa [36], where the authors tested various protocols to determine material
parameters for cyclic plasticity using inverse procedures. However, the comparison of the
simulation and experiments shown in Figure 13a exhibits the limitations of this approach,
as the predicted and measured stress–strain hysteresis loops for uniaxial loading differ
significantly from the parameters of set 1 that yield good comparability between experiment
and simulation for the HPFT case, as seen in Figure 13b.

exp't uniaxial exp't HPTF


sim. uniaxial, set 1 400 sim. HPTF, set 1
400 sim. uniaxial, set 2 sim. HPTF, set 2

200
200
Stress (MPa)

Stress (MPa)
0 0

200
200

400

400

0.006 0.004 0.002 0.000 0.002 0.004 0.006 0.02 0.01 0.00 0.01 0.02
Strain (-) Strain (-)
(a) (b)
Figure 13. Comparison of experimental and simulated stress–strain hysteresis loops (a) for uniaxial
fatigue with the strain amplitude E33 = 0.55% and (b) under HPTF loading with (R = −1, θ = 7.5◦ ,
E13 = 0.02). The material parameters for set 1 (Table 5) were obtained using an inverse analysis of the
HPTF experiments, whereas the parameters of set 2 (Table 6) were fitted to uniaxial experiments.

To understand the reason for this discrepancy in the prediction of the CPFE model, the
experimental stress–strain hysteresis loops for HPTF and uniaxial fatigue tests are analyzed
more closely. Figure 14 shows the comparison of the experimental results for saturated
hysteresis loops under HPTF loading (θ = 7.5◦ , R = −1) and uniaxial fatigue loading
(E33 = 0.55%, R = −1). Since the elastic stiffness in the former case is dominated by the
materials’ shear modulus and in the latter case by Young’s modulus, the stress is normalized
by the respective elastic quantities, which are estimated from the stiffness parameters given
in Table 3 as Young’s modulus Y = 185.2 GPa and shear modulus G = 76.5 GPa.
It is seen that the effective slopes of the elastic branches of the HPTF and uniaxial
hysteresis loops are significantly different, even after the proper scaling is applied, and
that the elastic response of the material under uniaxial fatigue is approximately 1.46 times
stiffer than under HPTF loading.
Due to this difference in the elastic slope, the effective elastic parameters required for
the uniaxial fatigue simulations are different from those used for the HPTF simulation, which
can be controlled by the scaling factor λ in this work. A similar change in the stress–strain
relationship due to pre-straining has been reported in the experimental work of Vrh et al. [37].
A degradation of Young’s modulus by more than 20% during fatigue tests was reported
for dual-phase steels in [38]. It can be interpreted from the experimental work of Wilshire
and Willis [39], Stout et al. [40], and Schneider et al. [41] that the magnified effect of defects
(especially dislocation structures) due to pre-straining in HPTF experiments can lead to the
change in the stress–strain behavior. It is also pointed out here that the temperature increase
during the HPTF experiments is significantly higher than that during the uniaxial tests (200 K
vs. 50 K), which might also lead to a reduction in the elastic stiffness and the yield stress of
the austenitic steel.

312
Materials 2023, 16, 1367

Figure 14. Comparison of the hysteresis loops in the regime where the stress amplitude is constant for
the uniaxial fatigue experiment (E33 = 0.55%, R = −1) and the HPTF loading case (θ = 7.5◦ , R = −1),
where θ and E33 are the amplitudes of the cyclic deformation. The values of all stresses in the plot are
normalized with the corresponding stiffness parameters, i.e., Young’s Modulus Y = 185.2 GPa for
uniaxial fatigue and shear modulus G = 76.506 GPa for the torsional loading during HPTF.

To demonstrate the possibility of calibrating the material parameters used for CPFE
simulations to describe uniaxial fatigue behavior, the scaling parameter for elastic properties
f
λ, the critical resolved shear stress τ0 , and the isotropic hardening parameters ({τc , h0 })
are re-calibrated to match uniaxial experiments, while keeping the kinematic hardening
parameters unchanged. The re-calibrated uniaxial parameters, referred to as set 2, are given
in Table 6. The CPFE results obtained for both loading cases are plotted in Figure 13 as set 2.
It is seen that this data set describes the stress–strain hysteresis under uniaxial loading very
well, whereas the results for HPTF are not represented in an acceptable manner for set 2. A
closer analysis of the material parameters of set 1 and set 2 reveals that the ratio of the fitted
elastic constants in the CP models of both loading cases amounts to 1.51, which is close to
the ratio of the effective stiffness values read from the experimental curves, which takes
a value of 1.46. Furthermore, the scaling parameter λ used to adapt the elastic constants
from DFT calculations to fit the uniaxial experiments is 0.95 and thus close to unity, which
demonstrates the predictive capabilities of DFT methods concerning the calculation of
elastic parameters. Generally, this finding that the transfer of crystal parameters obtained
by inverse analysis of data referring to one experiment cannot be simply transferred to
predict another experiment indicates a limitation in our current constitutive models and
their parameters.
However, this analysis also reveals, that the kinematic hardening parameters obtained
from uniaxial fatigue or HPTF loading are consistent, whereas the effective elastic param-
eters and, with them, some basic plastic properties, such as critical resolved shear stress
and isotropic hardening parameters, take different values for the two different loading
cases. The difference in these parameters can possibly be attributed to the higher tem-
peratures that occur during HPTF testing or to different dislocation structures that result
from multiaxial vs. uniaxial deformations. However, more research is required to draw a
firm conclusion and to derive a model that describes these differences quantitatively. It is
also pointed out here, that care needs to be taken when transferring material parameters
obtained for particular loading cases to different scenarios.

313
Materials 2023, 16, 1367

Table 6. Re-calibrated CPFE model parameters for uniaxial fatigue loading (set 2). The remaining
material parameters are unchanged with respect to set 1 and given Table 5.

f
λ τ0 τc h0
(-) (MPa) (MPa) (MPa)
0.95 188 92 −130.0

5. Conclusions
In this work, we investigated the mechanical behavior of austenitic stainless steel
1.3808 (commercial name P558), exposed to high-pressure torsion fatigue (HPTF), i.e.,
overlaid constant axial and cyclic torsional loading. Experimental tests revealed that under
such multiaxial reversed loading, the material undergoes plastic deformation in the axial
direction if the amplitude of the cyclic torsion exceeds a critical value, even if the axial
stress is well below the yield stress of the tested material. Since this time-dependent axial
deformation occurs under a constant axial stress that is smaller than the yield stress, it
can be considered a low temperature creep phenomenon, although its mechanisms are the
same as those for plastic deformation and do not involve diffusive processes. Such material
behavior has not yet been widely described in the literature. However, as multiaxial loading
occurs frequently in technical systems or medical implants, it is essential to understand
the mechanisms leading to this kind of plastic deformation. In order to accomplish this,
two different numerical models, J2 and crystal plasticity (CP), were employed in finite
element simulations of the HPTF experiments. The results indicate that the J2 model with
an associated flow rule is not suitable to describe material behavior under such multi-axial
cyclic loading cases. In contrast, the CP model with an Ohno–Wang kinematic hardening
model reveals a good agreement with the experiment and even was able to predict the
material response for different HPTF loading cases after fitting the material parameters to
experimental data from one individual test.
A comparison between uniaxial fatigue experiments and HPTF loading demonstrated
that there was a significant difference in the elastic branches of the respective stress–strain
hysteresis loops. The origin of this discrepancy might lie either in the significantly higher
temperature that builds up in the material during HPTF testing, in different dislocation
structures that develop during the different test conditions, or in a combination of both.
Consequently, the CP model calibrated by an inverse procedure based on HPTF data
is incapable of matching the results of uniaxial fatigue experiments. However, it was
demonstrated that a re-calibration of the material stiffness tensor and isotropic hardening
parameters by a scalar factor, while leaving the kinematic hardening parameters unchanged,
leads to a good description of the data obtained from uniaxial fatigue experiments. To
understand the correlation between the constitutive parameters and the resulting material
behavior in more detail, a parametric study of the CP kinematic hardening model under
HPTF loading was performed, and the resulting stress–strain hysteresis loops and axial
creep deformation were analyzed. The most important result of this case study is that a
reliable parameter identification by an inverse modeling approach based on experimental
data is possible, and that including axial deformation into the parameter fitting procedure
improves the calibration of the parameters for kinematic hardening. Generally, our results
indicate that material parameters that are identified by an inverse analysis of experimental
data from more complex situations can be used to describe simpler loading conditions,
whereas material parameters identified from simpler tests typically cannot be used to
describe more complex loading situations. However, it must also be pointed out here
that there are cases where material parameters obtained by inverse analysis of some test
conditions simply cannot be generalized to other situations, which is a clear limitation of
current constitutive models and the parameters they rely on.
The finite element model used for the simulation of the material response based on
crystal plasticity employed a simple representative volume element (RVE) that mimics the
crystallographic texture obtained from the EBSD analysis of the P558 steel specimens used in

314
Materials 2023, 16, 1367

the experiment. This RVE contained 512 cubic grains, thus ignoring microstructural features
like grain shape or annealing twins—a simplification that was necessary to minimize the
numerical effort in this work. In forthcoming studies, it is planned to investigate the
influence of grain shapes, annealing twins, and crystallographic texture on the material
behavior under HPTF loading in more detail by considering more realistic grain geometries
in the RVE simulations.

Author Contributions: Conceptualization, all authors; methodology, A.B., A.H.; software, A.B.,
A.A.G. and A.H.; validation, A.B., D.K. and T.N.; formal analysis, A.B., D.K., T.N. and A.A.G.;
investigation, D.K., T.N. and S.H.; resources, S.H., A.H.; writing—original draft, all authors; writing—
review & editing, all authors; visualization, A.B., D.K., T.N. and A.H.; supervision, S.H., A.H.; project
administration, S.H., A.H.; funding acquisition, S.H., A.H. All authors have read and agreed to the
published version of the manuscript.
Funding: This research was funded by the Deutsche Forschungsgemeinschaft (DFG, German Research
Foundation) under grant number 441180620.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Data will be made available upon request.
Acknowledgments: The authors thank voestalpine BÖHLER Edelstahl GmbH & Co KG, Kapfen-
berg, Austria, and in particular Rainer Fluch, for supporting this project and providing the steel
investigated here.
Conflicts of Interest: The authors declare no conflict of interest.

Nomenclature
Variable Description
σ, σij Cauchy stress
S Second Piola Kirchhoff stress
E, Eij Green-Lagrange strain

pl ,
eq Plastic strain tensor, equivalent plastic strain
eng Engineering strain
γ̇α Shear rate of dislocation on slip system α
γ̇0 Initial shear rate
τα Resolved shear stress for slip system α
τ0 Initial guess for the slip resistance
{σm , Q, b} Isotropic hardening material parameter for J2 plasticity model
{Ci , gi } Chaboche kinematic hardening material parameter for J2 plasticity model
f
{τc , h0 , p2 } Isotropic hardening parameters for CP model
{η, μ, m} Onho-Wang Kinematic hardening parameters for CP model
Y, G Young’s modulus and shear modulus
eq Equivalent strain
C Stiffness tensor
F Deformation gradient
Fe Elastic deformation gradient
Fp Plastic deformation gradient
I Unit tensor
q Crystallographic orientation
wi Weight of ith crystallographic orientation
f Orientation distribution function from experimental data
fˆ Orientation distribution function from reduced discrete orientations
ψΩ Kernel density function with half-width Ω
R Ratio of maximum and minimum strains in cyclic hysteresis loop
θ Amplitude of torsional cyclic deformation

315
Materials 2023, 16, 1367

Appendix A. Parametric Study of CPFE Parameters


In this section, a parametric study of the CP parameters for kinematic hardening is
performed, we focus particularly on the kinematic hardening parameters {η, μ, m}, while
f
keeping other parameters { γ̇0 , p1 , p2 ,h0 , τ0 , τc } constant. As reported by Schäfer et al. [20]
the CP kinematic hardening parameters η, μ mainly influence the stress–strain hysteresis loop,
and m controls the mean stress relaxation. Therefore the parametric study is divided into
two parts: (i) characterize the influence of η and μ on saturated stress–strain hysteresis loop,
and (ii) quantify the influence m on the mean stress relaxation in combination with the axial
creep behavior during HPTF loading. This study indicates that several sets of CP parameters
for kinematic hardening can produce a similar stress–strain hysteresis loop with minor
differences, however, parameters η and μ have a strong effect on the axial creep behavior.
For the parametric study, the hysteresis stress–strain loop (without isotropic hardening)
and the axial creep values are compared for the loading case θ = 7.5◦ , R = −1. To ensure
the saturation of the mean stress value, the hysteresis loop is taken after 25 cycles, and the
isotropic hardening is neglected. In the case of the axial creep simulation, the isotropic
hardening parameters are kept constant (as reported in Table 5). Figure A1 shows the
effect of the η and μ on the stress–strain hysteresis loop and the axial creep behavior. The
results indicate that both the stress–strain hysteresis loop and the axial creep behavior are
important for the calibration of the CP kinematic hardening parameters η and μ.

0.000
400

0.005
200
(MPa)

0.010
= 11000 MPa
E33 (-)

0
= 13000 MPa
= 15000 MPa
13

0.015

200
0.020

400 0.025
0.02 0.01 0.00 0.01 0.02 0 5 10 15 20 25
E13 (-) Cycle

0.000
400

0.005
200
(MPa)

= 100 MPa
0.010
E33 (-)

0
= 115 MPa
= 130 MPa
13

0.015

200
0.020

400
0.025
0.02 0.01 0.00 0.01 0.02 0 5 10 15 20 25
E13 (-) Cycle

Figure A1. Effect of CP kinematic hardening parameters η and μ on the stress–strain hysteresis loop
and the axial creep behavior.

To study the effect of the kinematic hardening parameter m on the mean stress relax-
ation and the axial creep deformation during HPTF the loading case θ = 10◦ , R = 0 is
chosen. Figure A2 shows the results from the simulation and illustrates that a reduction
of m leads to lower values of the mean stress and increases the increments of axial creep
per torsion cycle. Although the general trend in the change of the mean stress and the
axial deformation with respect to the change of the parameter m is similar, the mean stress
changes more uniformly (rather linearly) with the respect to m in comparison to axial
deformation, which has a progressive effect for smaller values of m.

316
Materials 2023, 16, 1367

Figure A2. Effect of the kinematic hardening parameter m on the axial creep behavior and the mean
stress relaxation during HPTF loading.

Appendix B. Elastic Constants Estimation


Density functional theory (DFT) calculations were employed to estimate the elastic
constants using the VASP [42–44] software. The calculations were performed with the pro-
jector augmented-wave (PAW) method [45], as implemented in VASP, and the generalized
gradient approximation, which used a Perdew-Burke-Ernzerhof (PBE) parametrization [46].
The structure of face-centered cubic Fe was constructed with repetitions of 4 × 2 × 2 of the
four-atom unit cell. Correspondingly, a 4 × 8 × 8 k-point mesh of the Monkhorst-Pack [47]
type was created. The substitutional alloying elements were added randomly, assuming
a disordered structure. The final composition of the supercell was Fe44 Cr11 Mn7 Mo2 C1 ,
where carbon occupies octahedral lattice sites. Spin-polarized calculations were carried out
to account for a ferromagnetic Fe state. The plane–wave basis cut-off energy was 500 eV.
In a cubic system, the independent elastic constants are C11 , C12 , and C44 . Applying
uniform isotropic straining to the lattice and fitting the obtained data to the Murnaghan
equation of state allows the calculation of the bulk modulus, B.

1
B= C + 2C12 (A1)
3 11
In an isotropic material C  = C11 - C12 . The values of C  and C44 can be found consider-
ing volume-conserving strains applying distortions, δ, in a ±2% range. To obtain C  , an
orthorhombic strain as follows was used,
⎛ ⎞
δ 0 0
ortho = ⎝0 −δ 0 ⎠ (A2)
0 0 δ2 / (1 − δ2 )
In the case of C44 , a monoclinic strain was considered:
⎛ ⎞
0 δ/2 0

mono = δ/2 0 0 ⎠ (A3)
0 0 δ2 / (4 − δ2 )
Respectively, the elastic constants’ values can be obtained from:

Eortho = E0 + C  Vδ2 + O[δ4 ] (A4)

Emono = E0 + C44 Vδ2 + O[δ4 ] , (A5)


where E0 is the energy of the system in its equilibrium state.
The numerical values of the elastic constants determined by this procedure are shown
in Table 3.

317
Materials 2023, 16, 1367

Appendix C. Material Parameter Optimization Algorithm


The optimization algorithm is implemented using python script (refer Figure A3),
the minimize function (with Nelder-Mead [48] method) is used from the python Scipy
optimize library. The function used for minimization takes in material parameters as input,
performs FEM simulations with the parameters, homogenizes the solution obtained from
FEM simulation, and provides the error between the simulation and experiments. The
sim | − σ exp | )2 , where n is the
mean square difference is used as the error function n1 ∑in=1 (σ13 i 13 i
number of time step in simulation, σ13 |i is the 13 (or xz) component of the result from the
sim
exp
simulation, and σ13 |i is the 13 (or xz) component of the result from the experiment.

Virtual
Mechanical test
‫ܧ‬ଵଷ

ߪଷଷ ‹ ݁‫݂ݎ‬


݁‫݂ݎ‬

ൌ ௡ଵ σ೙೔సభ ೞ೔೘ ȁ ିఙ
ఙభయ ೔
೐ೣ೛
భయ ȁ೔

Update
material
parameters

Figure A3. Flow diagram of the optimization algorithm used for optimizing the material parameters.

References
1. Thomann, U.I.; Uggowitzer, P.J. Wear-corrosion behavior of biocompatible austenitic stainless steels. Wear 2000, 239, 48–58. .
[CrossRef]
2. Uggowitzer, P.J.; Magdowski, R.; Speidel, M.O. Nickel free high nitrogen austenitic steels. ISIJ Int. 1996, 36, 901–908. [CrossRef]
3. Williams, D. Tissue-biomaterial interactions. J. Mater. Sci. 1987, 22, 3421–3445. [CrossRef]
4. Fini, M.; Giavaresi, G.; Giardino, R.; Lenger, H.; Bernauer, J.; Rimondini, L.; Torricelli, P.; Borsari, V.; Chiusoli, L.; Chiesa, R.;
et al. A new austenitic stainless steel with a negligible amount of nickel: An in vitro study in view of its clinical application in
osteoporotic bone. J. Biomed. Mater. Res. Part B Appl. Biomater. 2004, 71, 30–37. [CrossRef] [PubMed]
5. Montanaro, L.; Cervellati, M.; Campoccia, D.; Prati, C.; Breschi, L.; Arciola, C.R. No genotoxicity of a new nickel-free stainless
steel. Int. J. Artif. Organs 2005, 28, 58–65. [CrossRef] [PubMed]
6. Torricelli, P.; Fini, M.; Borsari, V.; Lenger, H.; Bernauer, J.; Tschon, M.; Bonazzi, V.; Giardino, R. Biomaterials in orthopedic surgery:
Effects of a nickel-reduced stainless steel on in vitro proliferation and activation of human osteoblasts. Int. J. Artif. Organs 2003,
26, 952–957. [CrossRef]
7. Niinomi, M.; Nakai, M.; Hieda, J. Development of new metallic alloys for biomedical applications. Acta Biomater. 2012,
8, 3888–3903. [CrossRef]
8. Salahinejad, E.; Hadianfard, M.J.; Ghaffari, M.; Mashhadi, S.B.; Okyay, A.K. Fabrication of nanostructured medical-grade stainless
steel by mechanical alloying and subsequent liquid-phase sintering. Metall. Mater. Trans. A 2012, 43, 2994–2998. [CrossRef]
9. Carreno-Morelli, E.; Zinn, M.; Rodriguez-Arbaizar, M.; Bassas, M. Nickel-free P558 stainless steel processed from metal
powder–PHA biopolymer feedstocks. Eur. Cells Mater. 2016, 32, 32.
10. Fini, M.; Aldini, N.N.; Torricelli, P.; Giavaresi, G.; Borsari, V.; Lenger, H.; Bernauer, J.; Giardino, R.; Chiesa, R.; Cigada, A. A
new austenitic stainless steel with negligible nickel content: an in vitro and in vivo comparative investigation. Biomaterials 2003,
24, 4929–4939. [CrossRef]
11. Ngeru, T.; Kurtulan, D.; Karkar, A.; Hanke, S. Mechanical Behaviour and Failure Mode of High Interstitially Alloyed Austenite
under Combined Compression and Cyclic Torsion. Metals 2022, 12, 157. [CrossRef]
12. Cruzado, A.; LLorca, J.; Segurado, J. Modeling cyclic deformation of inconel 718 superalloy by means of crystal plasticity and
computational homogenization. Int. J. Solids Struct. 2017, 122, 148–161. [CrossRef]

318
Materials 2023, 16, 1367

13. Kowalewski, Z.; Szymczak, T.; Maciejewski, J. Material effects during monotonic-cyclic loading. Int. J. Solids Struct. 2014,
51, 740–753. [CrossRef]
14. Hennessey, C.; Castelluccio, G.M.; McDowell, D.L. Sensitivity of polycrystal plasticity to slip system kinematic hardening laws
for Al 7075-T6. Mater. Sci. Eng. A 2017, 687, 241–248. [CrossRef]
15. Mróz, Z.; Maciejewski, J. Constitutive modeling of cyclic deformation of metals under strain controlled axial extension and cyclic
torsion. Acta Mech. 2018, 229, 475–496. [CrossRef]
16. Armstrong, P.J.; Frederick, C.O. A Mathematical Representation of the Multiaxial Bauschinger Effect; Berkeley Nuclear Laboratories:
Berkeley, CA, USA, 1966; Volume 731.
17. Lemaitre, J.; Chaboche, J.L. Mechanics of Solid Materials; Cambridge University Press: Cambridge, UK , 1994.
18. Ohno, N.; Wang, J.D. Kinematic hardening rules with critical state of dynamic recovery, part I: Formulation and basic features for
ratchetting behavior. Int. J. Plast. 1993, 9, 375–390. [CrossRef]
19. Bari, S.; Hassan, T. Anatomy of coupled constitutive models for ratcheting simulation. Int. J. Plast. 2000, 16, 381–409. [CrossRef]
20. Schäfer, B.J.; Song, X.; Sonnweber-Ribic, P.; Hartmaier, A. Micromechanical modelling of the cyclic deformation behavior of
martensitic SAE 4150—A comparison of different kinematic hardening models. Metals 2019, 9, 368. [CrossRef]
21. Sajjad, H.M.; Hanke, S.; Güler, S.; Fischer, A.; Hartmaier, A. Modelling cyclic behaviour of martensitic steel with J2 plasticity and
crystal plasticity. Materials 2019, 12, 1767. [CrossRef]
22. Hielscher, R. Kernel density estimation on the rotation group and its application to crystallographic texture analysis. J. Multivar.
Anal. 2013, 119, 119–143. [CrossRef]
23. Jin, Y.; Bernacki, M.; Rohrer, G.S.; Rollett, A.D.; Lin, B.; Bozzolo, N. Formation of annealing twins during recrystallization and
grain growth in 304L austenitic stainless steel. In Materials Science Forum; Trans Tech Publications Ltd.: Baech, Switzerland , 2013;
Volume 753, pp. 113–116.
24. Ren, S.; Sun, Z.; Xu, Z.; Xin, R.; Yao, J.; Lv, D.; Chang, J. Effects of twins and precipitates at twin boundaries on Hall–Petch relation
in high nitrogen stainless steel. J. Mater. Res. 2018, 33, 1764–1772. [CrossRef]
25. Mainprice, D.; Hielscher, R.; Schaeben, H. Calculating anisotropic physical properties from texture data using the MTEX
open-source package. Geol. Soc. 2011, 360, 175–192. [CrossRef]
26. Buhagiar, J.; Qian, L.; Dong, H. Surface property enhancement of Ni-free medical grade austenitic stainless steel by low-
temperature plasma carburising. Surf. Coatings Technol. 2010, 205, 388–395. [CrossRef]
27. Heidari, L.; Tangestani, A.; Hadianfard, M.; Vashaee, D.; Tayebi, L. Effect of fabrication method on the structure and properties of
a nanostructured nickel-free stainless steel. Adv. Powder Technol. 2020, 31, 3408–3419. [CrossRef]
28. Castelluccio, G.M.; McDowell, D.L. Effect of annealing twins on crack initiation under high cycle fatigue conditions. J. Mater. Sci.
2013, 48, 2376–2387. [CrossRef]
29. Pande, C.S.; Rath, B.; Imam, M. Effect of annealing twins on Hall–Petch relation in polycrystalline materials. Mater. Sci. Eng. A
2004, 367, 171–175. [CrossRef]
30. Biswas, A.; Vajragupta, N.; Hielscher, R.; Hartmaier, A. Optimized reconstruction of the crystallographic orientation density
function based on a reduced set of orientations. J. Appl. Crystallogr. 2020, 53, 178–187. [CrossRef] [PubMed]
31. Biswas, A.; Prasad, M.R.; Vajragupta, N.; ul Hassan, H.; Brenne, F.; Niendorf, T.; Hartmaier, A. Influence of microstructural
features on the strain hardening behavior of additively manufactured metallic components. Adv. Eng. Mater. 2019, 21, 1900275.
[CrossRef]
32. Smith, M. ABAQUS/Standard User’s Manual, Version 6.14; Dassault Systèmes Simulia Corp: Johnston, RI, USA, 2014.
33. de Borst, R.; Crisfield, M.A.; Remmers, J.J.C.; Verhoosel, C.V. Non-Linear Finite Element Analysis of Solids and Structures; Wiley:
Chichester, UK, 2012.
34. Roters, F.; Eisenlohr, P.; Hantcherli, L.; Tjahjanto, D.D.; Bieler, T.R.; Raabe, D. Overview of constitutive laws, kinematics,
homogenization and multiscale methods in crystal plasticity finite-element modeling: Theory, experiments, applications. Acta
Mater. 2010, 58, 1152–1211. [CrossRef]
35. McDowell, D. Stress state dependence of cyclic ratchetting behavior of two rail steels. Int. J. Plast. 1995, 11, 397–421. [CrossRef]
36. de Castro e Sousa, A.; Suzuki, Y.; Lignos, D. Consistency in solving the inverse problem of the Voce-Chaboche constitutive model
for plastic straining. J. Eng. Mech. 2020, 146, 04020097. [CrossRef]
37. Vrh, M.; Halilovič, M.; Štok, B. The evolution of effective elastic properties of a cold formed stainless steel sheet. Exp. Mech. 2011,
51, 677–695. [CrossRef]
38. ul Hassan, H.; Maqbool, F.; Güner, A.; Hartmaier, A.; Ben Khalifa, N.; Tekkaya, A.E. Springback prediction and reduction in deep
drawing under influence of unloading modulus degradation. Int. J. Mater. Form. 2016, 9, 619–633. [CrossRef]
39. Wilshire, B.; Willis, M. Mechanisms of strain accumulation and damage development during creep of prestrained 316 stainless
steels. Metall. Mater. Trans. A 2004, 35, 563–571. [CrossRef]
40. Stout, M.; Martin, P.; Helling, D.; Canova, G. Multiaxial yield behavior of 1100 aluminum following various magnitudes of
prestrain. Int. J. Plast. 1985, 1, 163–174. [CrossRef]
41. Schneider, A.; Kiener, D.; Yakacki, C.; Maier, H.; Gruber, P.; Tamura, N.; Kunz, M.; Minor, A.; Frick, C. Influence of bulk
pre-straining on the size effect in nickel compression pillars. Mater. Sci. Eng. A 2013, 559, 147–158. [CrossRef]
42. Kresse, G.; Hafner, J. Ab initio molecular dynamics for open-shell transition metals. Phys. Rev. B 1993, 48, 13115–13118. [CrossRef]
[PubMed]

319
Materials 2023, 16, 1367

43. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys.
Rev. B 1996, 54, 11169–11186. [CrossRef]
44. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775.
[CrossRef]
45. Blöchl, P.E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953–17979. [CrossRef]
46. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868.
[CrossRef] [PubMed]
47. Pack, J.D.; Monkhorst, H.J. Special points for Brillouin-zone integrations. Phys. Rev. B 1977, 16, 1748–1749. .
PhysRevB.16.1748. [CrossRef]
48. Gao, F.; Han, L. Implementing the Nelder-Mead simplex algorithm with adaptive parameters. Comput. Optim. Appl. 2012,
51, 259–277. [CrossRef]

Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.

320
materials
Article
Effects of Micro-Shot Peening on the Fatigue Strength of
Anodized 7075-T6 Alloy
Chih-Hang Su 1 , Tai-Cheng Chen 2 , Yi-Shiun Ding 3 , Guan-Xun Lu 1 and Leu-Wen Tsay 1, *

1 Department of Optoelectronics and Materials Technology, National Taiwan Ocean University,


Keelung 20224, Taiwan
2 Nuclear Fuels and Materials Division, Institute of Nuclear Energy Research, Taoyuan 32546, Taiwan
3 Material Research Group, Asia Development Center, SRAM LLC, Taichung 40765, Taiwan
* Correspondence: [email protected]; Tel.: +886-2-24622192 (ext. 6405)

Abstract: Micro-shot peening under two Almen intensities was performed to increase the fatigue
endurance limit of anodized AA 7075 alloy in T6 condition. Compressive residual stress (CRS) and a
nano-grained structure were present in the outermost as-peened layer. Microcracks in the anodized
layer obviously abbreviated the fatigue strength/life of the substrate. The endurance limit of the
anodized AA 7075 was lowered to less than 200 MPa. By contrast, micro-shot peening increased the
endurance limit of the anodized AA 7075 to above that of the substrate (about 300 MPa). Without
anodization, the fatigue strength of the high peened (HP) specimen fluctuated; this was the result of
high surface roughness of the specimen, as compared to that of the low peened (LP) one. Pickling
before anodizing was found to erode the outermost peened layer, which caused a decrease in the
positive effect of peening. After anodization, the HP sample had a greater fatigue strength/endurance
limit than that of the LP one. The fracture appearance of an anodized fatigued sample showed an
observable ring of brittle fracture. Fatigue cracks present in the brittle coating propagated directly
into the substrate, significantly damaging the fatigue performance of the anodized sample. The CRS
and the nano-grained structure beneath the anodized layer accounted for a noticeable increase in
resistance to fatigue failure of the anodized micro-shot peened specimen.

Keywords: AA 7075-T6 alloy; anodizing; micro-shot peening; nanograin; rotating bending fatigue
Citation: Su, C.-H.; Chen, T.-C.; Ding,
Y.-S.; Lu, G.-X.; Tsay, L.-W. Effects of
Micro-Shot Peening on the Fatigue
Strength of Anodized 7075-T6 Alloy.
Materials 2023, 16, 1160. https://
1. Introduction
doi.org/10.3390/ma16031160 Surface technologies are extensively applied to components in the aerospace, automo-
bile, and power industries to protect against corrosion and wear. Thermal spray coating,
Academic Editors: Lucjan Śnieżek,
electroplating, and chemical and physical vapor deposition are employed to modify the
Jaroslaw Galkiewicz and
Sebastian Lipiec
surface characteristics of the components. The most typical approach to increase the wear
and corrosion resistance of Al alloys is to coat them with anodic oxide film [1,2]. An oxide
Received: 31 December 2022 film formed on the surface of anodized Al alloys consists of a thin compact inner layer
Revised: 18 January 2023 and a porous outer layer [2]. A major concern about the use of AA 7075 Al alloy is the
Accepted: 27 January 2023 possible fatigue failure of structural components [3–5]. The anodization process is reported
Published: 29 January 2023 to obviously degrade the fatigue performance of high-strength Al alloys [6–11].
The effect of anodization on the fatigue strength/life of Al alloys is counted on several
process variables. Moreover, the fatigue strength of a soft anodized coating is superior
to that of a hard anodized one [12]. Chromic acid anodizing is less harmful to fatigue
Copyright: © 2023 by the authors.
strength than sulfuric acid anodizing [13,14]. Increasing the thickness of the anodizing
Licensee MDPI, Basel, Switzerland.
This article is an open access article
coating lowers the fatigue strength/life of AA 7075 alloy [10]. Moreover, a coating on
distributed under the terms and
an ultrafine-grained substrate has shown enhanced resistance to anodizing-induced and
conditions of the Creative Commons fatigue-induced cracking of AA 6682 alloy [12]. It is reported that the microcracks in the
Attribution (CC BY) license (https:// anodized film and irregularities beneath the film are the reasons for the decrease in fatigue
creativecommons.org/licenses/by/ strength [10]. The loss in fatigue strength of anodized AA 2219 and AA 2024 alloys is
4.0/). related to the induced microcracks at the etch pits, which are formed from the preferential

Materials 2023, 16, 1160. https://doi.org/10.3390/ma16031160 321 https://www.mdpi.com/journal/materials


Materials 2023, 16, 1160

dissolution of intermetallics in pickling performed before anodizing [7]. In one study, Al2 O3
coatings were applied on the AA 6061-T6 alloy by hard anodizing, micro-arc oxidation,
detonation spray, and air plasma spray, and hard anodizing was found to have the most
damaging effect on fatigue endurance [15].
Shot peening is employed to upgrade the fatigue performance of the Al alloys [16–18]. The
resulting compressive residual stress (CRS), strain hardening, and fine-grained microstructure
in the shot-peened layer are the main causes of improved fatigue properties. Successful fatigue
enhancement depends on a compromise between the CRS and the detrimental effect on surface
quality [19]. Advanced shot peening technologies, such as laser [20,21], ultrasonic [22,23],
and waterjet peening [24,25], are employed to reduce the irregularity and upgrade the fatigue
life of Al alloys. It is noted that shot peening can mitigate the harmful effect of an anodized
coating on the fatigue strength of high-strength Al alloys [13,21,26]. With shot peening, the
fatigue strength of anodized AA 6082-T651 alloy is even higher than that of the unanodized
substrate [26]. In 3.5% NaCl solution, the fatigue lives of AA 7075-T73 alloy can be markedly
increased by shot peening treatments as compared to an electropolished sample [27]. To reduce
surface irregularity, the use of fine particles for peening [28,29] is more economical and practical
than other advanced peening processes.
AA 7075-T6 alloy has been widely used in the aerospace industry, which can be heat-
treated to achieve various microstructures and properties [30]. A decrease in the surface
roughness of machined AA 7075 alloy can be achieved by increasing the cutting speed,
increasing the cutting depth, or decreasing the feeding rate [30]. The aim of this work
was to explore the application of micro-shot peening as a pretreatment before anodizing,
which provided a means to restore the fatigue performances of anodized AA 7075 alloy.
The fatigue limit of the specimens was evaluated by a rotating bending machine at room
temperature. The effect of hard anodizing on the fatigue limit/life of the unpeened and
peened AA 7075 alloy was investigated. Plastic deformation of the micro-shot peened
specimen was measured by using the Almen intensity, and the outer roughness was
detected with a 3D contour profiler. The surface morphologies of ruptured specimens
after fatigue tests were examined with a scanning electron microscope (SEM). The refined
structure in the peened samples was identified by the inverse pole figure (IPF) map.

2. Material and Experimental Procedures


2.1. Sample Preparation
The fatigue tests were performed on AA 7075 bar with 10 mm diameter. The chemical
composition (wt.%) of the alloy bar was as follows: 1.462 Cu, 2.504 Mg, 5.854 Zn, 0.029 Mn,
0.224 Cr, 0.07 Si, 0.117 Fe, 0.023 Ti, and the residual Al. Moreover, AA 7075 plate with
a thickness of 8 mm was used for surface metrology examination and residual stress
determination of the samples. The composition in wt.% of the plate was as follows: 1.294 Cu,
2.420 Mg, 5.616 Zn, 0.015 Mn, 0.292 Cr, 0.040 Si, 0.192 Fe, 0.020 Ti, and the balance Al. All the
samples were solution-annealed at 743 K for 1 h, followed by water-quenching, and then
aged at 393 K for 24 h. The solution-annealed and aged AA 7075 alloy was named the base
metal (BM) sample. The BM had a yield strength, tensile strength, and tensile elongation
of 618 MPa, 668 MPa, and 9%, respectively [31]. The tested specimens were ground with
SiC paper before micro-shot peening. Micro-shot peening was performed on the ground
samples using amorphous powders with sizes of 50–80 μm under 200% surface coverage.
The two micro-shot peen intensities employed in this work were determined from the
height of the N-type Almen specimen to be 0.110 mm (low peen, LP) and 0.204 mm (high
peen, HP). The micro-shot peened samples were distinguished as LP and HP according to
the Almen intensity.

2.2. Sulfuric Hard Anodizing


Some of the unpeened and peened samples were subjected to anodizing treatment.
Prior to the hard anodization process, the sample surface was thoroughly cleaned by de-
greasing, acid activation, and chemical polishing to totally remove the surface contaminants.

322
Materials 2023, 16, 1160

Pickling acts chemically to remove the oxides, inclusions, and compounds from the sample
surface. Hard anodizing to a thickness of 30 μm was performed in an electrolyte (20 wt.%
sulfuric electrolyte + 5 wt.% aluminum sulfate) at a bath temperature of −3 to 3 ◦ C under a
varying voltage (25 to 45 V). Sealing was performed in nickel acetate solution at 85–92 ◦ C.
For anodized samples, the symbol A affixed to the specified sample (e.g., HPA denoted the
anodized high peened sample).

2.3. Hardness Measurement and Fatigue Testing


An MVK-G1500 Vickers hardness tester (Mitutoyo, Kawasaki, Japan) was applied to
determine the hardness of the samples. The anodized coating was loaded at 10 gf for
15 s. Moreover, a Hysitron TI 980 nanoindenter (Bruker, Billerica, MA, USA) loaded at
2000 μN was used to determine the change in hardness around the interface between the
anodized and peened layers. Surface metrology of the samples with and without micro-shot
peening was performed with a Contour GT-K 3D optical profiler (Bruker, Billerica, MA,
USA), which provided noncontact surface measurements. Figure 1 displays the dimensions
of the dog-bone samples for fatigue tests. Rotary bending fatigue tests were conducted at
room temperature and a frequency of 33.3 Hz at R = −1 (fully reversed). Fatigue stress (S) vs.
number of cycles (N) to failure of the tested samples was measured, and the results presented
herein are the averages of three samples, although individual values are also reported.

Figure 1. Dimensions of the fatigue test specimen used in this study.

2.4. Microstructural and Fracture Surface Observations


The microstructures of the samples after metallurgical preparation were examined
using an S-4800 SEM (Hitachi, Tokyo, Japan). The samples were also detected using
an SEM equipped with a NordlysMax2 electron backscatter diffraction (EBSD, Oxford
Instruments, Abingdon, UK) detector to identify the refined structure of the inspected
specimens. Moreover, the strain fields around the interface between the coating and
substrate of the peened samples after anodizing were analyzed using the HKL Channel
5 software (Oxford Instruments, Abingdon, UK) to process the original data obtained by
the EBSD. The macro-fracture appearance and the detail of the fracture features of the
fatigue-fractured specimens were examined using an S-3400N SEM (Hitachi, Tokyo, Japan).

2.5. Residual Stress Measurement


The μ-X360s (Pulstec, Hamamatsu, Japan), a residual stress analyzer, was applied to
determine the distribution of residual stress in a micro-shot peened sample. The standard
settings of the X-ray source were using Cr target Kα radiation (wavelength 2.291 Å) at an
X-ray tube voltage of 30 kV with 1.5 mA current. The device for measuring residual stress
was based on the cos α method. The full width at half maximum (FWHM) of the (311)
peak was related to the distortion of the lattice. The distribution of residual stress in the
thickness direction was obtained by removing the surface layer of the sample using an EP-3
electrochemical polisher (Pulstec, Hamamatsu, Japan).

323
Materials 2023, 16, 1160

3. Results
3.1. Micro-Shot Peening and Morphology
The morphology of the amorphous shot and the surface appearance of the micro-shot
peened sample are displayed in Figure 2. The individual pellet of the amorphous shot
ranged from 50 to 80 μm in size (Figure 2a). As reported in previous studies [28], the
amorphous pellet had a Vickers microhardness of about HV 1150. The surface morphology
of the micro-shot peened samples (Figure 2b) had indentations of 10–30 μm in size. AA
7075 BM was ground with SiC paper before micro-shot peening. The surface roughness
of the BM and two micro-shot peened samples (LP and HP), determined by a 3D contour
profiler, are listed in Table 1. The ground sample had Sa, Sp, and Sv values of 0.217, 1.494,
and −1.211 μm, respectively. Continuous bombardments with amorphous shot increased
the surface roughness of the peened sample, especially that of the HP sample. The Sa, Sp,
and Sv roughness values of the LP one were 0.385, 1.396, and −1.506 μm, accordingly. This
revealed that the LP specimen had a slightly higher surface roughness than the unpeened
sample. The HP sample had an obviously higher surface roughness than the other two
samples; the Sa, Sp, and Sv were 1.303, 6.400, and −7.519 μm, respectively. In general,
such a high surface roughness of the HP sample was expected to degrade the resistance to
fatigue cracking.

Figure 2. Typical appearance of the (a) amorphous pellet, (b) surface morphology, and (c) surface
roughness of the micro-shot peened specimen.

Table 1. The surface roughness values of the distinct specimens (unit: μm).

Specimen Sa 1 Sp 2 Sv 3
Ground sample 0.217 1.494 −1.211
LP sample 0.385 1.396 −1.506
HP sample 1.303 6.400 −7.519
1Sa—arithmetical mean height of the surface. 2 Sp—maximum peak height of the surface. 3 Sv—maximum pit
depth of the surface.

Figure 3 displays the top and cross-sectional views of anodized samples with or without
micro-shot peening. The top surface features of the unpeened anodized sample showed
regular rectangular cracks (Figure 3a). In the HPA sample, shallow dents with vague cracks
were visible on the top surface (Figure 3b). It was obvious that the surface topography and
roughness of the anodized samples were affected by the surface texture of the substrate. The
cross-sectional views of the anodized layers of different samples are shown in Figure 3c,d,
revealing that the thickness of the anodized layer was about 30 μm in both samples. Without
peening, the top surface and the interface between the anodized layer and the substrate were
quite flat (Figure 3c). Some deep microcracks and fine defects of uneven sizes were present
in the anodized layer (Figure 3c). Overall, the anodized peened samples (HPA and LPA)
and the BMA had similar microstructures within the anodized zone. The difference between
them was that the straight interfaces of the unpeened sample (Figure 3c) were replaced by a
slightly tortuous profile of the peened one (Figure 3d). It was found that micro-shot peening
did not reduce the number of defects in the anodized zone. Within the anodized zone, the

324
Materials 2023, 16, 1160

open crack width of the peened sample seemed to be narrower than that of the unpeened
sample, but the difference was hard to distinguish.

Figure 3. (a,b) Top view and (c,d) cross-sectional view of the anodized sample: (a,c) the unpeened
sample; (b,d) the micro-shot peened sample.

3.2. Hardness Measurements


Figure 4 shows the hardness value of the anodized film and the hardness distribution
around the interface between the micro-shot peened zone and anodized layer, as deter-
mined by a Vickers hardness tester and nanoindenter, respectively. As shown in Figure 4a,
the anodized layer had an apparently higher hardness (around HV 400) than that of the
substrate (around HV 180). Thus, the anodized layer improved the resistance to corrosion
and wear of the AA 7075 alloy. The hardness indenters were also applied to the specific
sites around the interface of the anodized sample. The hardness values determined using a
nanoindenter on the right side of the interface (Figure 4b), i.e., the substrate side, were 1.74,
1.65, and 1.70 GPa, respectively. On the other side, hardness values of 3.83 and 2.92 GPa
were obtained; these two high values likely belonged to the anodized layer. In prior
work [30], the hardness of the micro-shot peened layer was found to exceed 2.94 GPa (HV
300). Thus, the strain hardening and nanocrystalline structure introduced by micro-shot
peening could be partly removed by pickling before anodizing.

Figure 4. The hardness indentations: (a) the micro-Vickers indentations of the anodized sample;
(b) the nano-hardness distribution around the anodized interface of the micro-shot peened sample.

325
Materials 2023, 16, 1160

3.3. Microstructural Observations


A highly deformed layer should consist of nanograins in the outermost surface and
micron-size elongated grains in the subsurface of the micro-shot peened sample [31].
Figure 5 presents the IPF and strain maps of the HPA and LPA samples, showing the
changes in the granular microstructures and strain fields around the interfaces between the
substrates and anodized layers. It should be noticed that no Kikuchi pattern was detected
within the anodized layer due to its amorphous structure. Because of the limited resolution
of the EBSD analysis, the ultrafine grains could not be distinguished and displayed in
different colors to show the individual grain boundaries or orientations. The IPF map
shows a nanogranular structure present within a depth of 10 μm below the interface of the
HPA sample (Figure 5a); this structure was associated with the original micro-shot peened
zone. As shown in Figure 5b, no fine granular structure was seen in the LPA sample. The
results indicated that only some fine grains occasionally nucleated along the boundaries
of coarse grains in the LPA sample. Those refined grains in the micro-shot peened layer
were caused by the dynamic recrystallization of the deformed surface layer during the
bombardment with fine shot. The original deformed and recrystallized layer disappeared in
the LPA sample (Figure 5b), which was attributed to the results of pickling before anodizing.
Figure 5c,d show the strain distribution maps, which are in cross-sectional view around
the anodizing interfaces of the HPA and LPA specimens. These two figures refer to the
IPF maps shown in Figure 5a,b. The high-strain zone is shown in red; the low-strain zone
appears in blue. The HPA sample exhibited high strain (in red) within a depth of 20 μm
beneath the interface (Figure 5c), which was not seen in the LPA sample (Figure 5d). Those
red zones in the HPA sample (Figure 5c) could be related to the severely deformed zones
during peening. Moreover, the LPA sample had low strain beneath the interface (Figure 5d).
It was confirmed that, in the HPA sample, part of the micro-shot peened zone remained
after pickling.

Figure 5. (a,b) The IPF maps and (c,d) the strain maps of the inspected samples in cross-sectional
view: (a,c) the HPA sample; (b,d) the LPA sample.

326
Materials 2023, 16, 1160

3.4. Fatigue Evaluation


Figure 6 demonstrates the results of fatigue tests conducted in laboratory air for up
to 107 cycles. The fatigue life of the tested samples was sensitive to the loading condi-
tion. Without peening and anodizing, the endurance limit of the AA 7075 BM was about
275 MPa [31] (Figure 6a). As revealed in prior work [31], the endurance limit of the LP sam-
ple was about 500 MPa, which was much greater than that of the AA 7075 BM (Figure 6a).
It was noticed that the fatigue properties of the HP sample were not better than those of
the LP specimen (Figure 6a). Furthermore, the HP sample exhibited high fluctuation in
fatigue strength/life during testing. The fatigue curves of anodized specimens are shown
in Figure 6b. The endurance limit of the anodized specimen (BMA) decreased to about
180 MPa (Figure 6b). It was obvious that hard anodizing had a strongly adverse effect on the
fatigue strength of AA 7075 alloy in the high-cycle region. Peening at a low Almen intensity
improved the fatigue performance of the LPA sample, as compared with that of the BMA
sample (Figure 6b). The endurance limit of the LPA specimen was about 300 MPa, if the
BM sample was pretreated with low peening intensity before anodizing. Peening at a high
Almen intensity raised the endurance limit of the HPA specimen to about 400 MPa, which
was about 100 MPa greater than that of the BM. The results indicated that the peening and
pickling strongly influenced the fatigue performance of anodized AA 7075 alloy. Moreover,
micro-shot peening could improve the fatigue performance of anodized AA 7075-T6 alloy.

Figure 6. Fatigue stress (S) vs. cycle (N) curves of the samples (a) without and (b) with a hard
anodizing coating.

3.5. Fractured Surface Examinations


The fatigue-fractured morphologies of the anodized BM sample (BMA) are shown in
Figure 7. The backscattering electron (BSE) image is more likely to reveal the microcracks
and delamination at the interface, whereas the secondary electron (SE) image can show
the detailed surface feature of the fatigue-cracked sample. Macroscopically, the main crack
grew from the outer surface and into the interior in a radial crack path (Figure 7a). In the
BSE image, a thin brittle case was found to decorate the outer profile of the BMA sample
(Figure 7b). The anodized layer contained some deep fine cracks (Figure 3) and had a flat
fracture appearance. This shows that the secondary cracks grew in the direction normal to
the external surface of the sample. Therefore, the cracking of the anodized layer resulted
in degradation of the fatigue resistance of the anodized AA 7075-T6 alloy. Moreover, the
presence of surface cracks within the anodized layer indicated that the crack-initiation
stage should be unnecessary. Examining the fracture features around the anodizing layer
at higher magnification, the fatigue crack propagated across the anodized coating and
into the substrate, and strong bonding between the anodized coating and substrate was

327
Materials 2023, 16, 1160

observed (Figure 7c). Moreover, a transgranular crack extended through the straight
interface and showed the trace of the crack path (parallel stripes) just beneath the anodized
layer (Figure 7d). This event confirmed the continuous progress of the crack without facing
the barrier. It is obvious that strong bonding between the anodized layer and the substrate
caused the fatigue crack growth rapidly through the interface (Figure 7c,d). In addition,
a typical transgranular cleavage-like fracture was observed as the fatigue crack extended
into the AA 7075 substrate (Figure 7d). Furthermore, a dimple fracture mixed with small
facets was seen in the final fractured zone (not shown here).

Figure 7. Fatigue fracture appearance of the BMA sample: (a,d) SE images; (b,c) BSE images.
(a) Macro-fractured appearance of the crack initiation site; (b) enlarged view of the anodized layer;
(c,d) fracture features around the anodized interface.

SEM photographs of the fractured appearance of the HPA specimen after fatigue
tests are presented in Figure 8. The results indicated that the macro-appearance of the
fractured HPA sample was similar to that of the BMA sample, consisting of a thin brittle
case decorating the outer profile of the specimen (Figure 8a,b). It was noticed that the flat
fracture of the anodized layer of the BMA sample was replaced by a chopped and smashed
layer in the HPA sample (Figure 8c). Moreover, relatively large numbers of fine cracks and
fragmental debris present in the anodized coating could be attributed to the brittle nature
of the coating under high fatigue loading of the test. It was noticed that micro-shot peening
before anodizing assisted the formation of a tortuous interface between the anodized layer
and the AA 7075 substrate. Numerous microcracks assisted in dividing the anodized layer
into many fine fragments after the fatigue test (Figure 8c). Moreover, the interface tended to
delaminate as the crack growth passed it (indicated by the arrows in Figure 8c). In addition,
a rubbed fracture feature was observed beneath the anodized layer (Figure 8d). It was
noticed that the fatigue-fractured feature revealed a microscopic change in crack growth
direction that occurred as the fatigue crack growth passed through the interface (Figure 8d).
It was deduced that the interface separation seemed to deflect the crack growth direction
(Figure 8d), which was beneficial in impeding the fatigue crack growth. The cleavage-like
brittle fracture in the BMA sample was replaced by the squeezed rubbed feature in the HPA
sample. Therefore, the CRS in the micro-shot peened specimen caused crack closure and
had a great effect on retarding the fatigue crack growth of the anodized specimen.

328
Materials 2023, 16, 1160

Figure 8. Fatigue fracture appearance of the HPA sample: (a,d) SE images; (b,c) BSE images. (a) Macro-
fractured appearance of the crack initiation site; (b) enlarged view of the anodized layer; (c,d) fracture
features around the anodized interface.

Figure 9 shows the changes in residual stress and FWHM intensity in the thickness
direction of the LP [31] and HP samples from the surface to the interior. As mentioned
previously, the increase in peening intensity caused an increase in surface roughness but
was expected to increase the strength and depth of the residual stress field. Peak CRS was
found in the subsurface zones of both peened samples. Moreover, the increase in peening
intensity caused an obvious increase in depth of the CRS field. A steep decrease in residual
stress to −48 MPa resulted in the LP specimen at a depth 50 μm below the peened surface.
However, under the same CRS, the depth was increased to about 100 μm in the HP one.
With the increase in peening intensity, the peak CRS increased from about −350 MPa to
over −400 MPa, and the stress field also increased. The increased CRS field was definitely
beneficial to the fatigue resistance to cracking. Moreover, the narrow CRS field meant that
only a very limited deformed depth was introduced by the micro-shot peening.

Figure 9. The changes in the residual stress and FWHM intensity in the X-ray diffraction pattern in
the thickness direction of the LP and HP samples from the surface to the interior.

329
Materials 2023, 16, 1160

4. Discussion
Although the anodized layer consisted of a few deep microcracks and fine defects of
uneven sizes (Figure 3), the anodized layer was much harder (around HV 400) than the
AA 7075-T6 substrate (around HV 180), as revealed in Figure 4. Thus, the anodized layer
improved the resistance to corrosion and wear of AA 7075 alloy. The straight boundary
between the anodized layer and the substrate of the unpeened specimen was replaced
by a slightly tortuous trace of the peened one (Figure 3). It was noticed that the shallow
dents with vague cracks were seen in the top surface of the micro-shot peened specimen.
By contrast, scratches with rectangular microcracks were observed in the top surface of
anodized unpeened specimen (Figure 3). Thus, the surface morphology and roughness
of the anodized sample were affected by the surface texture of the original condition. In
prior work, the micro-shot peened layer of AA 7075-T6 alloy comprised ultrafine grains in
the outermost layer and micron-order elongated grains in the subsurface of the micro-shot
peened specimen [31]. With increased peening intensity, the depth of the residual stress
field and the surface roughness are expected to be increased due to deep deformation.
The IPFs showed a nanogranular structure within a depth 10 μm below the anodized
interface of the HPA sample (Figure 5a), but no such structure was detected in the LPA
specimen (Figure 5b). It was clear that pickling before anodizing completely removed the
fine-grained structure of the LPA sample. The retained fine structure, which was caused
by the recrystallization of the heavily deformed zone during peening, was linked with
the detectable strain within a depth of 20 μm beneath the anodized interface of the HPA
sample (Figure 5c).
The results of fatigue tests in air revealed that the fatigue strengths/lives of the micro-
shot peened (LP and HP) samples were much greater than that of the BM, particularly
that of the LP one (Figure 6a). The high fluctuation in the fatigue strength/life of the HP
sample in air, relative to that of the LP specimen, could be attributed to its inherent high
surface roughness. Similar results have been reported: increasing the machined roughness
of AA 7010-T7451 alloy from 0.6 to 3.2 μm caused a 32% drop in endurance limit in a
low-stress state [11]. It was noted that hard anodizing had a very detrimental effect on the
high-cycle fatigue of AA 7075 alloy shown in Figure 6b. The fatigue strength/life will be
obviously shortened by the defects, especially those of surface defects [32]. A probabilistic
fatigue life prediction model using the calibrated weakest-link theory has been proposed
considering the superiority of the notch and surface defect [32,33]. With the presence of
surface microcracks in the anodized coating, the fatigue limit of the BMA sample decreased
to about 180 MPa. It was interesting that the fatigue strengths/lives of the peened anodized
samples (LPA and HPA), especially that of the HPA, were much higher than that of the
unpeened anodized (BMA) samples, as shown in Figure 6b. With micro-shot peening, the
harmful effects of microcracks present in the anodized coating on the fatigue performance
of AA 7075 alloy were mitigated. The difference in the outermost microstructure between
the LPA and HPA samples was due to the retained nanocrystal structure of the latter
(Figure 5a), which was hard to find in the former (Figure 5b). It was obvious that pickling
almost completely removed the nano-grained structure of the LPA sample. It was deduced
that precise control of the pickling process would highly alter the fatigue performance of
anodized Al alloys, especially that of the micro-shot peened samples.
Micro-cracks initiate at corrosion pits around the surface and etch pits of AA 7050 alloy
under cyclic loading [3]. Furthermore, fatigue failure of anodized Al alloy is often induced
by the nucleation and growth of those micro-cracks formed in the anodic oxide film and
through the interface into the substrate [9,10]. The stress concentration at the microcracks
in the coating is responsible for the decreased fatigue properties of hard anodized 7475-
T6 alloy [9]. It has been reported that pickling deteriorates the fatigue properties due
to localized dissolution around the intermetallic compounds and/or inclusions [7,34,35],
especially in the very-long-life regime [34]. Therefore, pickling was one of the controlling
factors for the fatigue performance of the anodized Al alloys. As shown in Figure 7,
fatigue cracks present in the brittle coating of an anodized sample grew through the

330
Materials 2023, 16, 1160

coating/substrate interface and propagated directly into the substrate without the need
of micro-crack initiation. This event illustrated the poor fatigue resistance of the BMA
sample, as compared with that of the BM sample. Since micro-shot peening only influenced
a thin external layer of the impacted material, the difference in the damaging mechanisms
between the anodized samples with and without peening should only exist in the stage
of retardation of crack growth below the anodized layer. The nano-grained structure and
the CRS beneath the anodized layer were helpful in delaying or retarding fatigue crack
propagation, which accounted for the higher fatigue strengths/lives of the peened anodized
samples than that of the unpeened anodized one.
The macro-fractured appearances of the anodized unpeened and peened specimens
were similar, as shown in Figures 7 and 8. However, differences in the fracture features
could still be distinguished between the samples in detail. Under high loading, the anodized
layer of the peened sample was cracked into small patches (Figure 8c), in contrast to the
overall flat fracture of the anodized layer of the unpeened sample (Figure 7c). Moreover,
a change in the fatigue fracture appearance around the interface between the unpeened
and micro-shot peened samples was observed. The crack propagated directly into the
substrate of the unpeened sample, showing a cleavage-like fracture therein (Figure 7d). As
the crack growth passed the interface of the micro-shot peened sample, delamination of
the interface occurred (Figure 8c,d); this delamination could cause a change in the crack
growth direction. It was noticed that a rubbed fracture feature instead of cleavage-like
fracture was observed just below the anodized interface of the micro-shot peened sample
(Figure 8d). The CRS around the anodized interface enhanced the crack closure, retarding
the fatigue crack growth and leading to the rubbed fracture feature.

5. Conclusions
1. Micro-shot peening under two Almen intensities was performed to increase the fatigue
strength/life of the anodized AA 7075-T6 alloy. Micro-cracks in the anodized layer
significantly deteriorated the fatigue performance of AA 7075 alloy. Under high cycle
fatigue, the endurance limit of the BMA sample was lowered to less than 200 MPa.
Micro-shot peening could improve the fatigue strength/life of the anodized sample
to the level of the unanodized substrate (about 300 MPa). Without anodizing, the
fatigue performance of the HP sample was worse than that of the LP one. Moreover,
fluctuation in the fatigue strength of the HP sample, relative to the LP sample, was
attributed to the inferior effect of high surface roughness.
2. With an anodized layer, the fatigue strength of the HP sample was higher than that of
the LP one. Pickling before anodizing eroded the outermost nanograins in the peened
layer, which degraded the positive effect of micro-shot peening. Therefore, the HPA
sample had higher fatigue resistance than the LPA one did. The fracture appearance
of the anodized fatigued samples consisted of an observable ring of brittle fracture.
Without any need for crack initiation, fatigue cracks present in the anodized layer
propagated directly into the unpeened substrate, significantly reducing its fatigue
strength/life. By contrast, the presence of CRS beneath the anodized layer of the
micro-shot peened sample could retard the fatigue crack growth and was responsible
for a noticeable increase in its fatigue performance. Moreover, interfacial separation
between the anodized layer and peened surface could possibly deflect the crack path,
which also contributed to the increased resistance to fatigue crack growth of the
micro-shot peened sample.

Author Contributions: Conceptualization, L.-W.T.; methodology, L.-W.T.; formal analysis, C.-H.S.,


T.-C.C. and G.-X.L.; investigation, C.-H.S., T.-C.C. and G.-X.L.; resources, L.-W.T.; writing—original
draft preparation, L.-W.T.; writing—review and editing, T.-C.C. and Y.-S.D.; visualization, T.-C.C.;
supervision, L.-W.T.; project administration, L.-W.T.; funding acquisition, L.-W.T. All authors have
read and agreed to the published version of the manuscript.

331
Materials 2023, 16, 1160

Funding: This research was funded by the Ministry of Science and Technology, R.O.C., grant number
MOST 110-2622-E-019-008.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Data available on request due to restrictions eg privacy or ethical.
Acknowledgments: The authors are grateful to the Ministry of Science and Technology (National Tai-
wan University) for the assistance in TEM (EM003800) examinations and EPMA analysis (EPMA000300).
The authors are grateful to the Core Facility Center of National Cheng Kung University (OTHER002200)
for granting access to the 3D optical profiler. The authors also thank the Ministry of Science and Tech-
nology (National Sun Yat-sen University) for the assistance in sample preparations using a focused
ion beam (EM025100). The authors would also like to thank Likuan Technology Corp. for their great
help in determining the residual stress of the micro-shot peened sample and Vincent Vacuum Tech. for
performing the micro-shot peening.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Lerner, L.M. Hard Anodising of Aerospace Aluminum Alloys. Trans. Inst. Met. Finish. 2010, 88, 21–24. [CrossRef]
2. Matinez-Viademonte, M.P.; Abrahami, S.T.; Hack, T.; Burchardt, M.; Terryn, H. A Review on Anodizing of Aerospace Aluminum
Alloys for Corrosion Protection. Coatings 2020, 10, 1106. [CrossRef]
3. Barter, S.A.; Sharp, P.K.; Holden, G.; Clark, G. Initiation and Early Growth of Fatigue Cracks in an Aerospace Aluminum Alloy.
Fatigue Fract. Eng. Mater. Struct. 2002, 25, 111–125. [CrossRef]
4. Ma, J.; Wang, Q.; Yang, Y.; Yang, F.; Dong, B.; Che, X.; Cao, H.; Zhang, T.; Zhang, Z. Anisotropic Low Cycle Behavior of the
Extruded 7075 Al Alloy. Materials 2021, 14, 4506. [CrossRef] [PubMed]
5. Turkmen, H.S.; Loge, R.E.; Dawson, P.R.; Miller, M.P. On the Mechanical Behaviour of AA 7075-T6 during Cyclic Loading. Int. J.
Fatigue 2003, 25, 267–281. [CrossRef]
6. Shiozaw, K.; Kobayash, H.; Terada, M.; Matsui, A. Effect of Anodized Coatings on Fatigue Strength in Aluminum Alloy. Trans.
Eng. Sci. 2001, 33, 398–406. [CrossRef]
7. Rateick, R.G., Jr.; Griffith, R.J.; Hall, D.A.; Thompson, K.A. Relationship of Microstructure to Fatigue Strength Loss in Anodised
Aluminum-Copper Alloys. Mater. Sci. Technol. 2005, 25, 1227–1235. [CrossRef]
8. Sadeler, R. Effect of a commercial hard anodizing on the fatigue property of a 2014-T6 aluminium alloy. J. Mater. Sci. 2006, 41,
5803–5809. [CrossRef]
9. Lonyuk, B.; Apachitei, I.; Duszczyk, J. The Effect of Oxide Coatings on Fatigue Properties of 7475-T6 Aluminium Alloy. Surf. Coat.
Technol. 2007, 201, 8688–8694. [CrossRef]
10. Cirik, E.; Genel, K. Effect of anodic oxidation on fatigue performance of 7075-T6 alloy. Surf. Coat. Technol. 2008, 202, 5190–5201.
[CrossRef]
11. Shahzad, M.; Chaussumier, R.; Chieragatti, R.; Mabru, C.; Rezai-Aria, F. Influence of Anodizing Process on Fatigue Life of
Machined Aluminium Alloy. Procedia Eng. 2010, 2, 1015–1024. [CrossRef]
12. Hockauf, K.; Winter, L.; Händel, M.; Halle, T. The Effect of Anodic Oxide Coating on the Fatigue Behaviour of AA6082 with an
Ultrafine-Grained Microstructure. Mater. Werkst. 2011, 42, 624–631. [CrossRef]
13. Camargo, A.; Voorwald, H. Influence of Anodization on the Fatigue Strength of 7050-T7451 Aluminium Alloy. Fatigue Fract. Eng.
Mater. Struct. 2007, 30, 993–1007. [CrossRef]
14. Inoue, A.K.; Tagawa, T.; Ishikawa, T. The Effect of Anodizing on the Fatigue Life of Aluminum Alloy 7050. J. Jpn. Inst. Met. 2012,
76, 365–374. [CrossRef]
15. Sundararajan, G.; Wasekar, N.P.; Ravi, N. The Influence of the Coating Technique on the High Cycle Fatigue Life of Alumina
Coated Al 6061 Alloy. Trans. Indian Inst. Met. 2010, 63, 203–208. [CrossRef]
16. Oguri, K. Fatigue Life Enhancement of Aluminum Alloy for Aircraft by Fine Particle Shot Peening (FPSP). J. Mater. Process.
Technol. 2011, 21, 1395–1399. [CrossRef]
17. Abdulstaar, M.; Mhaede, M.; Wollmann, M.; Wagner, L. Investigating the Effects of Bulk and Surface Severe Plastic Deformation
on the Fatigue, Corrosion Behaviour and Corrosion Fatigue of AA5083. Surf. Coat. Technol. 2014, 254, 244–251. [CrossRef]
18. Benedetti, M.; Fontanari, V.; Bandini, M.; Savio, E. High- and Very High-Cycle Plain Fatigue Resistance of Shot Peened High-
Strength Aluminum Alloys: The Role of Surface Morphology. Int. J. Fatigue 2015, 70, 451–462. [CrossRef]
19. Trsko, L.; Guagliano, M.; Bokuvka, O.; Novy, F.; Jambor, M.; Florkova, Z. Influence of Severe Shot Peening on the Surface State
and Ultra-High-Cycle Fatigue Behavior of an AW 7075 Aluminum Alloy. J. Mater. Eng. Perform. 2017, 26, 2784–2797. [CrossRef]
20. Gao, Y.K. Improvement of Fatigue Property in 7050-T7451 Aluminum Alloy by Laser Peening and Shot Peening. Mater. Sci. Eng.
A 2011, 528, 3823–3828. [CrossRef]

332
Materials 2023, 16, 1160

21. Luong, H.; Hill, M.R. The Effects of Laser Peening and Shot Peening on High Cycle Fatigue in 7050-T7451 Aluminum Alloy.
Mater. Sci. Eng. A 2010, 527, 699–707. [CrossRef]
22. Pandey, V.; Chattopadhyay, K.; Srinivas, N.C.S.; Singh, V. Role of Ultrasonic Shot Peening on Low Cycle Fatigue Behavior of 7075
Aluminium Alloy. Int. J. Fatigue 2017, 103, 426–435. [CrossRef]
23. Singh, V.; Pandey, V.; Kumar, S.; Srinivas, N.C.S. Effect of Ultrasonic Shot Peening on Surface Microstructure and Fatigue Behavior
of Structural Alloys. Trans. Indian Inst. Met. 2015, 69, 295–301. [CrossRef]
24. Muruganadhan, R.; Mugilvalavan, M.; Thirumavalavan, K.; Yuvaraj, N. Investigation of Water Jet Peening Process Parameters on
AL6061-T6. Surf. Eng. 2018, 34, 330–340. [CrossRef]
25. Mahmoudi, A.H.; Jamali, A.M.; Salahi, F.; Khajeian, A. Effects of Water Jet Peening on Residual Stresses, Roughness, and Fatigue.
Surf. Eng. 2021, 37, 972–981. [CrossRef]
26. Hadzima, B.; Novy, F.; Trško, L.; Pastorek, S. Shot Peening as a Pre-Treatment to Anodic Oxidation Coating Process of AW 6082
Aluminum for Fatigue Life Improvement. Int. J. Adv. Manuf. Technol. 2017, 93, 3315–3323. [CrossRef]
27. Mhaede, M. Influence of Surface Treatments on Surface Layer Properties, Fatigue and Corrosion Fatigue Performance of AA7075
T73. Mater. Des. 2012, 41, 61–66. [CrossRef]
28. Chung, Y.-H.; Chen, T.-C.; Lee, H.-B.; Tsay, L.-W. Effect of Micro-Shot Peening on the Fatigue Performance of AISI 304 Stainless
Steel. Metals 2021, 11, 1408. [CrossRef]
29. Sawada, T.; Yanagitani, A. Properties of Cold Work Tool Steel Shot Peened by 1200 HV-Class Fe-Cr-B Gas Atomized Powder as
Shot Peening Media. Mater. Trans. 2010, 51, 735–739. [CrossRef]
30. Zhang, P.; Liu, J.P.; Gao, Y.R.; Liu, Z.H.; Mai, Q.Q. Effect of heat treatment process on the micro machinability of 7075 aluminum
alloy. Vacuum 2023, 207, 111574. [CrossRef]
31. Su, C.-H.; Chen, T.-C.; Tsay, L.-W. Improved Fatigue Strength of Cr-electroplated 7075 Al by Micro-Shot Peening. Int. J. Fatigue
2023, 167, 107354. [CrossRef]
32. Li, X.-K.; Zhu, S.-P.; Liao, D.; Correia, J.A.F.O.; Berto, F.; Wang, Q.Y. Probabilistic fatigue modelling of metallic materials under
notch and size effect using the weakest link theory. Int. J. Fatigue 2022, 159, 106788. [CrossRef]
33. He, J.C.; Zhu, S.-P.; Luo, C.Q.; Niu, X.P.; Wang, Q.Y. Size effect in fatigue modelling of defective materials: Application of the
calibrated weakest-link theory. Int. J. Fatigue 2022, 165, 107213. [CrossRef]
34. Wang, Y.Q.; Kawagoishi, N.; Chen, Q. Effect of Pitting Corrosion on Very High Cycle Fatigue Behavior. Scr. Mater. 2003, 49,
711–716. [CrossRef]
35. Shahzad, M.; Chaussumier, M.; Chieragatti, R.; Mabru, C.; Rezai-Aria, F. Surface Characterization and Influence of Anodizing
Process on Fatigue Life of Al 7050 Alloy. Mater. Des. 2011, 32, 3328–3335. [CrossRef]

Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.

333
materials
Article
Statistical Characterization of Strain-Controlled Low-Cycle
Fatigue Behavior of Structural Steels and Aluminium Material
Žilvinas Bazaras and Vaidas Lukoševičius *

Faculty of Mechanical Engineering and Design, Kaunas University of Technology, Studentu˛ Str. 56,
51424 Kaunas, Lithuania
* Correspondence: [email protected]

Abstract: Probabilistic evaluation of the resistance to low-cycle deformation and failure of the critical
components in the equipment used in the energy, engineering, metallurgy, chemical, shipbuilding,
and other industries is of primary importance with the view towards their secure operation, in
particular, given the high level of cyclic loading acting on the equipment during its operation. Until
recently, systematic probabilistic evaluation has been generally applied to the results of statistical
and fatigue investigations. Very few investigations applying this approach to the low-cycle domain.
The present study aims to substantiate the use of probabilistic calculation in the low-cycle domain by
systematic probabilistic evaluation of the diagrams of cyclic elastoplastic deformation and durability
of the materials representing the major types of cyclic properties (hardening, softening, stabilization)
and investigation of the correlation relationships between mechanical properties and cyclic defor-
mation and failure parameters. The experimental methodology that includes the calculated design
of the probabilistic fatigue curves is also developed and the curves are compared to the results of
the experiment. Probabilistic values of mechanical characteristics were determined and calculated
low-cycle fatigue curves corresponding to different failure probabilities, to assess them from the
Citation: Bazaras, Ž.; Lukoševičius, V.
probabilistic perspective. A comparison of low-cycle fatigue curves has shown that the durability
Statistical Characterization of
curves generated for some materials using analytical expressions are not accurate. According to
Strain-Controlled Low-Cycle Fatigue
Behavior of Structural Steels and
the analysis of the relative values of experimental probabilities of low-cycle fatigue curves, the use
Aluminium Material. Materials 2022, of analytical expressions to build the curves can lead to a significant error. The results obtained
15, 8808. https://doi.org/10.3390/ allow for the revision of the load bearing capacity and life of the structural elements subjected to
ma15248808 cyclic elastoplastic loading in view of the potential scattering of mechanical properties and resistance
parameters to low-cycle deformation and failure. In addition, the results enable determination of the
Academic Editors: Thomas Niendorf,
scatter tolerances, depending on the criticality of the part or structure.
Jaroslaw Galkiewicz, Lucjan Śnieżek
and Sebastian Lipiec
Keywords: correlation; durability; low-cycle fatigue; regression; probability; strain-controlled loading
Received: 14 November 2022
Accepted: 7 December 2022
Published: 9 December 2022

Publisher’s Note: MDPI stays neutral 1. Introduction


with regard to jurisdictional claims in Individual structural elements of machinery and units have been subjected to elasto-
published maps and institutional affil- plastic, static, or cyclic deformation caused by operation under high stress conditions related
iations. to the attempt to achieve maximum performance indicators (capacity, output, speed) while
maintaining minimum metal consumption [1,2]. For greater operational reliability and
safety of different structures and products, probabilistic methods are used increasingly
extensively to calculate their strength and durability. These methods are based primarily
Copyright: © 2022 by the authors.
on the use of statistical information on the mechanical properties and durability of the
Licensee MDPI, Basel, Switzerland.
material under cyclic loading [3–5].
This article is an open access article
distributed under the terms and
The probabilistic approach to the calculation of structures and the determination of
conditions of the Creative Commons
the design characteristics of materials has been developed for more than 80 years. Weibull
Attribution (CC BY) license (https:// made a considerable contribution to the development of probabilistic methods, probabilistic
creativecommons.org/licenses/by/ substantiation of the permissible stresses, and strength safety margins for the calculation of
4.0/). static and cyclic strength [6,7]. The ‘weakest link hypothesis’ developed by Weibull allowed

Materials 2022, 15, 8808. https://doi.org/10.3390/ma15248808 335 https://www.mdpi.com/journal/materials


Materials 2022, 15, 8808

the scholar to build the theory of probabilistic weak fracture of bodies under the action
of stress. This helped to solve the issues of fatigue failure theory. The strain-controlled
failure strength was investigated by Coffin [8], Manson [9], and Langer [10]. The study by
Iida and Inoue [11] provided findings on the probability of low-cycle fatigue failure. The
results of the life distribution under strain-controlled loading in low-cycle fatigue tests were
investigated using the normal, log-normal, and Weibull distributions. Low-cycle fatigue
durability distributions were found to be fairly consistent with the log-normal distribution
and Weibull distribution rather than with the normal distribution. Scattering of the crack
initiation life was found to generally exceed scattering of the durability to failure. The
concept of fatigue design curves was addressed in the investigation of the strength and life
reduction factors. A statistical evaluation of the fatigue characteristics of the sample tested
at the same level primarily requires addressing the issue of the distribution law. Different
distribution functions were proposed by Freudenthal and Gumbel [12,13] as well as other
researchers in relation to this issue. Serensen, Kogayev, Shneiderovich [14,15], Stepnov [16],
revised and upgraded probabilistic fatigue calculations under low-cycle fatigue strain-
controlled loading, performed the analysis of random deviations of the acting stress-strain
forces, and analyzed the distribution of the probability of fatigue failure according to the
durability characteristics.
Several researchers contributed considerably to the calculation of probabilistic meth-
ods for the mechanical and low-cycle properties. Makhutov et al. [17] presented the
results of experimental investigations and durability and life calculations of low alloy and
austenitic steels with different mechanical properties. Daunys et al. [18,19] investigated the
dependence of low-cycle fatigue durability on the mechanical properties of steel welds in
nuclear power plants. Timofeev et al. [20] and Raslavičius et al. [21] investigated the proba-
bilistic low-cycle fatigue life and the dependence of low-cycle durability on the mechanical
properties WWER-type reactor of the nuclear power plant made of steel 15Cr2MoVA and
structural steel C45. Another study [22] involved the analysis of the mechanical reliability
of magnesium alloys and systematic evaluation using the statistical Weibull analysis. The
results obtained are very important in terms of the safety and reliability evaluation of the
magnesium alloys for lightweight structures. Zhu et al. [23] have developed a probabilistic
methodology to predict low cycle fatigue life by using an energy-based damage parameter
built on the Bayes theorem. Strzelecki [24] presented the characteristics of the S-N curve
that employ the 2-parameter and 3-parameter Weibull distributions for fatigue limit and
limited life, respectively. The parameters of the proposed model were evaluated under the
maximum likelihood method. In addition, the article presented the solution to the problem
of estimating the initial values of the likelihood function. Fekete [25] proposed a new model
for low-cycle fatigue prediction based on strain energy to account for just the part of the
microstructure of the strain energy stored in the material that causes the fatigue damage.
The above review of the work has demonstrated that, until now, statistical evaluation
under low-cycle loading has only been applied to durability [26–28]. There are no studies
dedicated to the statistical evaluation of the strain diagrams or the parameters thereof.
The evaluation of the durability to the final rupture or crack initiation has generally been
performed under low-cycle fatigue conditions by applying the laws of normal, log-normal,
and Weibull distribution. The following could be concluded in relation to the findings of
the investigation of static characteristics and low-cycle fatigue:
• The mechanical characteristics and the characteristics of resistance to low-cycle de-
formation and failure have not been investigated to the extent that would be enable
comprehensive application of its results to the probabilistic calculations.
• In the case of low-cycle loading, there is lack of systematic data on the investigation
of the laws of distribution of durability for materials with different cyclic properties
(hardening, softening, stabilization), while static investigations of resistance to low-
cycle deformation under strain-controlled loading are non-existent.
• It is known that the curves of low-cycle fatigue under strain-controlled loading can
be built using the mechanical characteristics, for example, σys , σu , ψ etc. However,

336
Materials 2022, 15, 8808

there are no systematic investigations that would be dedicated to determination of


correlation relations between mechanical characteristics and durability.
Based on the above review of the scientific studies and the conclusions thereof, the
following objectives were pursued in the study presented here: (1) Definition of the laws of
distribution of the mechanical characteristics and low-cycle (deformation and durability)
characteristics for the materials that contrast to each other by their cyclic properties (hard-
ening, softening, stable) with the aim of subsequent reliable static assessment (by using
the confidence intervals) of the mechanical and resistance properties of low-cycle fatigue;
(2) Definition of the correlation relations between the key mechanical characteristics and
parameters of low-cycle fatigue that enable building of the low-cycle fatigue curves by us-
ing the mechanical characteristics; (3) Development of the methodology for determination
of the probabilistic curves of low-cycle failure on the basis of the investigation conducted.

2. Materials, Specimens and Methods


2.1. Materials
Due to the statistic nature of the fatigue failure as well as static failure, there is
scattering of the material properties, which may also depend on the cyclic properties of
materials and the character of failure under low-cycle loading (fatigue, quasistatic, transient
failure). Therefore, the static investigations conducted in the study used the specimens
produced of the materials that contrasted with each other by cyclic properties: cyclically
softening low alloy steel 15Cr2MoVA, cyclically stable medium carbon steel C45, and
cyclically hardening aluminium alloy D16T1. The same cast was used for production of
the specimens.
The 15Cr2MoVA specimens were cut from 120 mm thick rolled stock making sure that
the direction of the rolled stock corresponded to the specimen axis. Steel specimens C45
and aluminium alloy D16T1 were cut out of 50 mm diameter bar stock. The 15Cr2MoVA
forging and aluminium alloy were subject to thermal processing in the following modes:
hardening in oil by heating up to 1000 ◦ C followed by two subsequent quenching sequences
at 700 ◦ C for 14 h and in 670 ◦ C for the 70 h for steel 15Cr2MoVA and standard hardening
and quenching for the aluminium alloy D16T1. The chemical composition [29,30] and
mechanical characteristics of the investigated materials are presented in Tables 1 and 2.

Table 1. Chemical composition of the materials.

C Si Mn Cr Ni Mo V S P Mg Cu Al
Material
%
15Cr2MoVA (GOST 5632-2014) 0.18 0.27 0.43 2.7 0.17 0.67 0.30 0.019 0.013 - - -
C45 (GOST 1050-2013) 0.46 0.28 0.63 0.18 0.22 - - 0.038 0.035 - - -
D16T1 (GOST 4784-97) - - 0.70 - - - - - - 1.6 4.5 9.32

Table 2. Mechanical properties of the materials.

epr σpr σys σu σf ψ


Material
% MPa %
15Cr2MoVA (GOST 5632-2014) 0.200 280 400 580 1560 80
C45 (GOST 1050-2013) 0.260 340 340 800 1150 39
D16T1 (GOST 4784-97) 0.600 290 350 680 780 14

Table 2 provides the expected value of the mechanical characteristics of the investigated
materials, while Figure 1 provides the single deformation representing these characteristics,
that is, the curves corresponding to the 50% probability.

337
Materials 2022, 15, 8808

Figure 1. Single uniaxial tensile deformation curves: 1—aluminium alloy D16T, 2—structural steel
C45, 3—alloyed steel 15Cr2MoVA.

Based on Table 2, steels 15Cr2MoVA and C45, especially steel 15Cr2MoVA (ψ = 80%)
are plastic materials. The diagram of single uniaxial tensile deformation of steel C45
is characterized by the yield plateau (up to e = 0.6 − 0.65), while the diagrams of single
uniaxial tensile deformation of steel 15Cr2MoVA and aluminum alloy D16T1 do not contain
a yield plateau. The chosen materials represent key types of cyclic properties: cyclic
hardening, stabilization, and failure. Therefore, the experimental and theoretical findings
obtained can be applied to the assessment of a wide range of materials used in structures
subject to low-cycle loading.

2.2. Specimens
For the purpose of cyclic experiments, specimens with a cylindrical deformable part:
(length—23 mm, diameter—10 mm) were selected. The cutting modes of processing were
chosen to avoid traces of crushing and vibrations on the working surface. The specimens
were processed with a special 60mm diameter form turning tool for all the materials
investigated. A detailed drawing of the specimen is presented in Figure 2.

Figure 2. Specimens for low-cycle fatigue experiments (units in mm).

The specimens made of steel 15Cr2MoVA or C45 were machined with a 0.1 mm for
grinding on the main part and with allowance 0.15 mm on the surface of the deformable
part. The grinding of the deformable part and the base surfaces of the specimens was
performed on the external grinding machine with a rounding radius of 25 mm. The
minimum radial and face runout of the specimen heads in relation to the cylindrical
deformable part was achieved by grinding of the surfaces on the basis of the central holes.
The samples made of aluminium alloy D16T1 were processed by turning only. The final
passages were performed with cutting models that provided roughness of the working and
base surfaces in the detailed drawing (Figure 2).
After fatigue tests the fractured specimens were used as workpiece materials to pro-
duce monotonous tensile specimens with the aim of reaching the material properties nearly
identical to the properties of the material subjected to cyclic loading. The detailed drawing
of the specimen is presented in Figure 3. The specimens for uniaxial tension experiments
have a diameter of 5 mm and 25 mm length. They were taken from the parts of cyclic

338
Materials 2022, 15, 8808

test specimens (part of 17.5 mm diameter, Figure 2), that had not been subjected to plastic
deformation.

Figure 3. Specimens for single uniaxial tension experiments (units in mm).

To eliminate the elastoplastic bending of the specimen during production as a result


of the cutting forces, the cutting depth and feed were decreasing as the diameter to be
machined decreased. For the same purpose, the processing was performed using special
high-speed steel cutters the geometry whereof was chosen with the view towards minimum
cutting forces.

2.3. Methods
The experiments were carried out in the Laboratory of the Faculty of Mechanical
Engineering and Design of Kaunas Technology University. The experimental equipment
consisted of a 50 kN UMM-5T low-cycle tension-compression test machine (Kaunas Uni-
versity of Technology, Kaunas, Lithuania) and an electronic device designed to record
stress-strain diagrams, cycles, and load reversal. Mechanical characteristics were measured
with an error that did not exceed ±1% of the deformation scale or ±0.01% of the maximum
load. GOST 25502-79 standard (Strength analysis and testing in machine building. Meth-
ods of metals mechanical testing. Methods of fatigue testing) [31] was used to perform
low-cycle fatigue tests. Statistical characteristics were calculated according to the GOST
22015-76 standard (Quality of product, Regulation, and statistical quality evaluation of
metal materials and products on speed-torque characteristics) [32].
As mentioned above, to define the patterns of statistical distribution of mechanical
characteristics and parameters of low-cycle loading, three materials with contrasting cyclic
properties were investigated in the study: softening steel 15Cr2MoVA, mildly softening
(virtually stable) steel C45, and hardening aluminium alloy D16T1. The experiments were
conducted under low-cycle strain-controlled symmetric tension and compression. Low-
cycle fatigue curves under strain-controlled loading in relative coordinate system ‘total
strain amplitude e0 —durability to complete fracture N f  are presented in Figure 4.

Figure 4. Low-cycle fatigue durability curves under strain-controlled loading: 1—aluminum alloy
D16T, 2—alloyed steel 15Cr2MoVA, 3—structural steel C45.

339
Materials 2022, 15, 8808

The curves were built according to the guidelines established in the Rules and Norms
in Nuclear Power Engineering (PNAE) [33]. According to the guidelines, low-cycle fatigue
curves must be built after at least 15 tests on the specimens under different levels of
uniformly distributed load.
The levels of loading and the number of specimens under the static investigation of
each of them are presented in Table 3.

Table 3. Specification of low-cycle strain-controlled loading (e0 = constant).

Material Loading Level, e0 Number of Specimens, Pcs.


1.8 40
15Cr2MoVA 3.0 80
5.0 40
2.5 50
C45 4.0 120
6.0 50
1.0 20
D16T1 1.5 80
2.0 20

According to Table 3, static investigations under strain-controlled loading were con-


ducted at three levels of loading. The three levels of deformation under strain-controlled
loading were chosen to be uniform throughout the investigation range. Although only fa-
tigue failure was observed under strain-controlled loading, all three levels were considered
equal in relation to the type of failure. However, the number of specimens of the materials
investigated at the medium level was increased in order to identify the effect of the number
of specimens on the statistical characteristics of the deformation and durability diagrams.

3. Results
3.1. Statistical Investigation of the Relationship between the Mechanical Characteristics and
Parameters of Resistance to Cyclic Deformation and Failure
Under strain-controlled loading, only fatigue failure is attainable due to the limitation
on total deformation of the specimen due to the conditions of the experiment. Strain-
controlled loading tests are generally used for the definition of failure characteristics. The
most widely used Coffin—Manson equation has been applied to the strength calculation,
as it defines the relationship between the size of the plastic deformation and the number of
cycles to failure [8,9]:
ea N f m = Cψ (1)
where m and Cψ are the constants of material that, according to the Coffin data, have the
following values for the majority of materials: m = 0.5, Cψ = 12 ln 1−1 ψ .
Equation (1) expresses the linear relationship between plastic deformation and the
number of cycles to failure in coordinate system lge p –lgN f . Plastic deformation changes in
the process of strain-controlled loading and is constant only for cyclically stable materials.
Hence, in Equation (1), the authors recommend using value ea , that corresponds to 50% of
the loading cycles to failure, i.e., when the process of width stabilization of the elastoplastic
hysteresis loop starts. Manson, when testing Equation (1), found that, for 29 materials with
contrasting cyclic properties, constant m = 0.6. Manson expressed the relationship between
total elastoplastic deformation and the number of cycles to failure as a single dependency.
The amplitude of total deformation ea was calculated as a sum of amplitudes of plastic e p
and elastic strain ey , i.e.,

1 100 0.6 −0.6 σu


e a = e p + ey = (ln ) Nf + 1.75 N f −0.12 (2)
2 100 − ψ E

340
Materials 2022, 15, 8808

Based on the equation by Langer [10], to determine the failure amplitudes of deforma-
tion ea and conditional stresses σa∗ under strain-controlled symmetric loading, the following
dependencies were proposed:

1 1 100 σu
ea = ln + 0.4 (3)
4et N m
f 100 − ψ Ee t

1 E 100
σa∗ = ln + 0.4σu (4)
4 Nm
f 100 −ψ
where m—constant equal to 0.5 under σu ≤ 687 (MPa).
The above dependencies are often used by designers for the calculation of heavily
loaded parts and structures under low-cycle deformation conditions.
The prepared version of PNAE proposes calculating the elastoplastic deformation by
using the dependency:
100
0.5 ln
100 − ψ σu
ea = + (5)
(4N f )0.5 E(4N f )0.05
According to the investigations by Daunys [34], the following could be written down
for the majority of materials:
ea N f α1p = C1p (6)
In contrast to Coffin–Manson Equation (1), in this case α1p < m and C1p < Cψ .
Constants α1p and C1p can preliminarily be defined according to the mechanical properties
of the material:
σys 100
α1p = 0.17 + 0.55ψ , C1p = 0.75α1p ln (7)
σu 100 − ψ
Similarly, the same study [34] attempted to link the parameters of the generalized
curve of cyclic deformation A1 , A2 , ∝, c, S T to the mechanical properties of material. The
following was obtained:
σys
A1 = 0.3 + 0.6ψ (8)
σu
σu −7
A2 = 0.32( ) + A1 (9)
σys
σys
∝= 0.9 + 2.6ψ (10)
σu
σys
c = 0.13 − 0.21ψ (11)
σu
σys
S T = 2 − 0.83ψ (12)
σu
The analysis of the dependencies proposed by different authors for the calculation
of structures and elements under elastoplastic deformation conditions has shown that
durability Nc and N f and parameters of the generalized diagram of cyclic deformation
A1 , A2 , ∝, c, S T are often linked by the dependencies that are used for the calculation of
mechanical characteristics.
However, the scientific literature reviewed did not investigate the level of correlation
between these parameters. Therefore, the present study includes investigation of the corre-
σ
lation relations between Nc , N f and σys , σu , σ f , ψ, ψu , ψ σysu . The results were processed
according to the known methods of mathematical statistics [35].
The coefficient of correlation between two correlating values was determined as
follows:
m xy
r= (13)
σx σy

341
Materials 2022, 15, 8808

where the second central mixed moment was determined according to the following
dependency:
⎡ n n ⎤
⎢ n ∑ xi ∑ yi ⎥
1
m xy = ⎢ xi yi − 1 1 ⎥
n ⎣∑ ⎦ (14)
1
n

while root mean square deviations of the investigated correlating quantities were deter-
mined according to the following dependencies:
( (
1 n 2  1 n 2 
n∑ n∑
σx = xi − x2 , σy = y i − y2 (15)
1 1

The mean arithmetic values of the investigated correlating quantities were determined
according to the following dependencies:

1 n 1 n
n∑
x= xi , y = ∑ yi (16)
1
n 1

The values of coefficients of regression b between the correlating quantities were


determined on the basis of the following expressions:
σx σy
bx/y = r , by/x = r (17)
σy σx

In the course of calculation of the value of the coefficient of correlation r, it can approx-
imately be assumed that the estimate thereof has been distributed normally. Therefore, the
confidence interval of the valid values r is:

1 − r2
r ± tγ √ (18)
n

while in case of γ = 0.90 and tγ = 1.65 [36]:

1 − r2
r ± 1.65 √ (19)
n

The linear regression equation can be brought into the following form:

( x − x ) = bx/y (y − y) (20)

The analysis of the results of calculation of correlation coefficients has suggested that,
under strain-controlled loading, durability Nc and N f correlates very well (almost linearly)
σ
with mechanical characteristics σys , σu , σ f , ψ, ψu , ψ σysu for all the investigated materials
with contrasting cyclic properties. For steel 15Cr2MoVA, there is almost linear correlation
with durability Nc and N f yield strength σys and true fracture strength σ f (Figure 5).
According to Figure 5, there is a close correlation between the durability and me-
chanical characteristics of the 15Cr2MoVA steel. Similar results were obtained for the
regression coefficients of the resistance characteristics to cyclic deformation in relation to
the mechanical characteristics.
For steel C45, at the loading level e0 = 2.5 the correlation is better with durability Nc
and N f —reduction of area ψ, while at the level e0 = 4.0, the correlation is better with true
fracture strength σ f and yield strength σys . For the aluminium alloy D16T1, yield strength
σu correlates directly to durability. Interestingly, for steels 15Cr2MoVA and C45, multiplier
σ σ
ψ σysu insignificantly increases the coefficient of correlation between Nc ,N f and ψ σysu . For
σ
aluminium alloy D16T1, introduction of multiplier ψ σysu leads to certain reduction of the
coefficient of correlation for the levels of strain-controlled loading e0 = 1.0; 1.5, while for

342
Materials 2022, 15, 8808

level e0 = 2.0, the multiplier increases the coefficient of correlation, same as in the case of
steels 15Cr2MoVA and C45.

Figure 5. Regression lines under strain-controlled loading for steel 15Cr2MoVA (e0 = 1.8). Dots
represent the experimental values, straight lines—the theoretical calculated dependencies, while
dashed lines represent confidence intervals.

3.2. Statistical Assessment of the Low-Cycle Fatigue Curves under Strain-Controlled Loading
Until present, the low-cycle fatigue curves under strain-controlled loading have been
built using Equations (1)–(6) or similar equations by applying the mechanical characteristics
that correspond to a probability of 50%. Therefore, the curves calculated here correspond
to the same failure probability. There are no systematic investigations into the building of
calculated probabilistic low-cycle fatigue curves in the scientific literature.
In this study, probabilistic values of mechanical characteristics were determined. This
enabled the authors to build the calculated low-cycle fatigue curves corresponding to
different failure probabilities and to assess them from the probabilistic perspective.
Tables A1–A3 contain the values σys , σu , ψ, e pr , corresponding to 1%, 10%, 30%, 50%,
70%, 90%, 99% probabilities for all the materials investigated. The low-cycle fatigue curves
under strain-controlled loading corresponding to 1%, 10%, 30%, 50%, 70%, 90%, 99% failure
probabilities for the investigated steel 15Cr2MoVA were built using Equations (3), (5) and

343
Materials 2022, 15, 8808

(6) and the values of equal probability of mechanical characteristics (Figure 6a–c). The
calculated curves were built using the absolute coordinates lge0 − lgNc .

(a) (b)

(c)

Figure 6. Calculated probabilistic low-cycle fatigue curves under strain-controlled loading for steel
15Cr2MoVA, built using the absolute coordinates, according to: (a)—Equation (3), (b)—Equation (5),
(c)—Equation (6).

Based on Figure 6 the results of calculations using Equations (3) and (5), are little
different from each other in terms of both the slope and occupied scatter band of the
probabilistic low-cycle fatigue curves. The ratio between durability values for 99% and
1% curves is little dependent on the deformation level. The average ratio for the curves
calculated using Equation (3) was about 3.2, for curves calculated using Equation (5)—
about 3.3. Calculation of the probabilistic low-cycle fatigue curves using Equation (6) for
steel 15Cr2MoVA generates a non-satisfactory result as the curves cross each other in case
of N f = 200 − 400 cycle durability or are positioned in reverse order in case of durabilities
N f > 400, i.e., the durability is the lowest in case of 99% failure probability and the highest
in case of 1% failure probability. This is associated with dependency of α1p and C1p on
the mechanical characteristics of the material. As follows from Equations (7) and (8), the
probabilistic value of α1p largely depends on ψ, as the probabilistic values of ratio σys /σu
differ very little (Table A1). Constant C1p that depends on both α1p and ψ changes within
the range from 0.42 for probability curve 1% to 0.96 for probability curve 99%. Constant α1p
ranges from 0.41 (failure probability curve 1%) to 0.56 (failure probability curve 99%). The
specified changes α1p and C1p lead to the positioning of the probabilistic curves as depicted
in Figure 6c.
The dependencies built in the relative coordinates are used for calculation of the
low-cycle fatigue of parts and structural elements. Hence, the probabilistic curves of low-
cycle fatigue under strain-controlled loading for steel 15Cr2MoVA were also built in the
relative coordinates using Equations (3), (5) and (6). The amplitude strains of curve 1%

344
Materials 2022, 15, 8808

were divided by probabilities e pr —1%, strains of curve 10%—by probabilities e pr —10%, etc.
Hence, the designed calculated probabilistic curves of low-cycle fatigue in the coordinates
lge0 − lgNc are presented in Figure 7a–c.

(a) (b)

(c)

Figure 7. Calculated probabilistic low-cycle fatigue curves under strain-controlled loading for steel
15Cr2MoVA, built in the relative coordinates, according to: (a)—Equation (3), (b)—Equation (5),
(c)—Equation (6).

Figure 7 suggests that the application of the relative coordinates for the design of prob-
abilistic curves of low-cycle fatigue generates an implausible picture for steel 15Cr2MoVA.
The implausibility of the mutual position of the calculated probabilistic curves lies in that
curve 99% is characterized by the lowest durability, while curve 1%—the highest durability.
The ‘reverse’ layout of the probabilistic curves of low-cycle durability is related to very
vast scatter e pr compared to the scatter of other mechanical characteristics, for example, ψ,
that largely determine the durability. As suggested by Table A1, for steel 15Cr2MoVA, the
ratio of values e pr for probability 99% to 1% probability is 7.6, while the ratio of values ψ
for probability 99% to 1% probability—1.4.
The sharp contrast in the values of the scatter between e pr on one side and σys , σu , ψ on
the other side is likely to be due to higher sensitivity e pr to thermal processing, hardening
during mechanical processing, accuracy of the experiment, and other factors, in comparison
to other mechanical characteristics. The conducted analysis of the calculated probabilistic
curves for steel 15Cr2MoVA suggests that the probabilistic values of strain e pr , cannot be
used for the design of probabilistic curves of low-cycle fatigue in the relative coordinates as
they distort the true layout of the curves.
To define more truthful layout of the calculated probabilistic curves of low-cycle
fatigue under strain-controlled loading in the relative coordinates for steel 15Cr2MoVA,
the percentage curve strain was divided by the mean arithmetic value of e pr (Figure 8a–c).

345
Materials 2022, 15, 8808

(a) (b)

(c)

Figure 8. Comparison of the calculated probabilistic low-cycle fatigue curves under strain-controlled
loading with the experimental ones for steel 115Cr2MoVA; straight lines—calculated, dashed lines—
experiment; according to: (a)—Equation (3), (b)—Equation (5), (c)—Equation (6).

Here, as suggested by Figure 8a–c there is little difference in the layout of the prob-
abilistic curves in case of the absolute coordinates (Figure 6). For the calculated curves
built according to Equations (3) and (5), the ratio of durability for curve 99% and curve
1% was little dependent on the strain. The average ratio for the curves built according to
Equation (3) was about 2.9, and for the curves built according to Equation (5)—3.2.
To validate the calculation for steel 15Cr2MoVA, the experimental curves of equal
probability were compared to the calculated curves. Figure 8a,b suggests that the slope
angle and the occupied scatter band of the experimental curves of equal probability is
little different from those of the calculated curves designed according to Equations (3)
and (5). Nonetheless, the experimental curves were located lower than the calculated
curves. For example, at low durability, experimental curve 99% corresponded to calculated
curve 30% (Equation (3)), while experimental curve 50% corresponded to calculated curve
1% (Figure 8a). Correspondence between the experimental curves and calculated curves
according to Equation (5) was less accurate at high durability. In this case, experimental
curve 99% was calculated above than calculated area 1% for Nc > 3000 cycle durability. In
case of durabilities Nc < 3000, experimental curve 99% corresponded to calculated curve 1%
(Figure 8b). Figure 8c suggests that the calculated probabilistic curves designed according
to Equation (6) completely fall within the zone of the experimental curves. Nonetheless, the
comparison renders the ‘reverse’ layout of the calculated probabilistic curves impossible in
the area of Nc > 200 − 400 cycle durabilities. The reasons have already been covered above.
Similar analysis was conducted with the calculated and experimental probabilistic low-
cycle fatigue curves under strain-controlled loading for steel C45. Equations (3), (5) and (6)
were used to calculate the curves of equal probability by applying the probabilistic values
of mechanical characteristics (Table A2). The obtained results were built using the absolute
coordinates lge0 − lgNc . For steel C45, same as for steel 15Cr2MoVA, the probabilistic

346
Materials 2022, 15, 8808

calculated curves obtained according to Equations (3) and (5) were positioned in a similar
way in terms of both the slope angle and the occupied scatter band. For probabilistic curves
calculated according to Equation (6), increase accompanied by higher failure probability
was observed, same as in the case of steel 15Cr2MoVA. For steel C45, however, the larger
range of variation of the probabilistic value and smaller range of variation of probabilistic
value C1p , than for steel 15Cr2MoVA lead to regular layout of the probabilistic curves, i.e.,
curve 1% provides the lowest durability, while curve 99%—the highest durability.
The investigation of the durability scatter band for steel C45 has demonstrated that
the ratio of durability for probabilistic curves 99% and 1% depends on the strain level. As
suggested by the analysis performed, for curves designed according to Equations (3), (5)
and (6), the ratio of durability of curves 99% and 1% at strain amplitude e0 = 0.9% was
7.3; 8.2; 10.6 respectively, while at e0 = 0.4% — 10; 18.7; 3.7. For steel C45, same as for
steel 15Cr2MoVA, the use of probabilistic value e pr for design of probabilistic calculated
curves in the relative coordinates lge0 − lgNc distorts their true layout in terms of all the
dependencies applied Equations (3), (5) and (6), i.e., the curves are positioned in the reverse
order. At strain amplitude e0 = 4, the ratio of durability of probabilistic curves 1% and 99%
was respective 2.8; 3.5; 8.3, and at e0 = 2 — 4.0; 6.7; 20.7. To obtain a valid layout of the
calculated probabilistic curves in the relative coordinates for steel 15Cr2MoVA and steel
C45, strains e0 were divided by mean arithmetic value e pr . Hence, the obtained probabilistic
curves are depicted in Figure 9a–c.

(a) (b)

(c)

Figure 9. Comparison of the calculated probabilistic low-cycle fatigue curves under strain-controlled
loading with the experimental ones for steel C45; straight lines—calculated, dashed lines—experiment;
according to: (a)—Equation (3), (b)—Equation (5), (c)—Equation (6).

The calculated probabilistic curves are close to the curves in the absolute coordinates by
the layout character and slope angle. For these curves, the durability ratio of the curves 99%
and 1% depend on the strain level, the same as for the curves in the absolute coordinates, i.e.,
at strain amplitude e0 = 4, the ratio of durability according to Equations (3), (5) and (6) is 7.1;
7.6; 10.3, and at e0 = 2 — 8.4; 11.7; 5.3.

347
Materials 2022, 15, 8808

At the same time, the same figures also include the experimental low-cycle fatigue
curves under strain-controlled loading for steel C45 (Figure 9a–c). Comparison of the
experimental probabilistic curves with the calculated one has shown that the calculated
probabilistic curves for steel C45 (Figure 9a–c) are located below the 1% experimental
probabilistic curve. Of all the dependencies applied to the calculation of probabilistic
low-cycle fatigue curves under strain-controlled loading, the calculated (Equation (6)) for
steel C45 was the closest to reality as demonstrated by the investigation.
The calculations of the probabilistic low-cycle fatigue curves under strain-controlled
loading for aluminum alloy D16T1 (Table A3) were performed according to Equation (6), as
Equations (3) and (5) were designed for low-alloy steels used for energy purpose. The same
methodology was used for the design of the calculated curves as for the steels 15Cr2MoVA
and C45, that is, the curves were designed in the absolute coordinates and in the relative
coordinates by using probabilistic and mean arithmetic values e pr .
Investigating the durability scatter band for the D16T1 aluminium alloy has demon-
strated that the ratio for probabilistic curves 99% and 1% depends on the strain level
(Figure 10).

Figure 10. Probabilistic low-cycle fatigue curves under strain-controlled loading for aluminium alloy
D16T1 built using: the absolute coordinates (a); the relative coordinates (b); straight lines—calculated,
dashed lines—experiment.

The conducted analysis has shown that the ratio of durability at strain amplitude
e0 = 0.3 was 37, and at e0 = 0.18% — 24. At the same time, in the case of the calculated
probabilistic curves, an increase in the slope angle with the increase in the failure probability
has been observed. This is directly affected by the scatter of reduction of area ψ, same as
for steels 15Cr2MoVA and C45. Application of relative strain e0 , determined according to
the probabilistic values of strain e pr to the calculations leads to narrowing of the durability
scatter band. In this case, the ratio of durability of the probabilistic curves 99% and 1% at
strain amplitude e0 = 4 was 3.3, and at e0 = 3 − 2.7. The slope angles of the calculated
probabilistic curves in the relative coordinates were reducing in comparison to the same
curves in the absolute coordinates. However, the curve angle of the slope increased with
increasing failure probability.
Same as for steels 15Cr2MoVA and C45, to calculate the strains e0 , of aluminium
alloy D16T1, mean arithmetic value e pr was used. In this case, the ratio of durability of
probabilistic curves 99% and 1% was close to the results of the ratio of durabilities of the
probabilistic curves designed in the absolute coordinates, i.e., at e0 = 4, the ratio was 3.3,
and at e0 = 3 − 2.7 (Figure 10b). Figure 10b also depicts the comparison of the calculated
probabilistic curves with the experimental ones for aluminium alloy D16T1. As suggested
by Figure 10b the correspondence of the experiment results with the calculated results is
non-satisfactory, as the calculated curves are fully reflected in the elastic area.

348
Materials 2022, 15, 8808

Figures 11 and 12 for steels 15Cr2MoVA and C45 compare the experimental low-cycle
fatigue curves under strain-controlled loading of failure probability 1%, 50%, 99% with the
calculated curves designed according to Equations (3), (5) and (6) by using the normalized
mechanical characteristics determined according to Equation (21) and mechanical charac-
teristics taken from reference documents [37], and the low-cycle fatigue curves defined
using the safety factor n N = 10 for cycles and ne = 2 for strain as used in the field of
mechanical engineering.
xmax
K= (21)
xmin

Figure 11. Comparison of the probabilistic low-cycle fatigue curves calculated under strain-controlled
loading with the experimental ones for steel 15Cr2MoVA; straight lines—calculated, dashed lines–
experiment; normalized characteristic according to: 1—Equation (3), 2—Equation (5), 3—Equation (6);
reference characteristics, according to: 1 —Equation (3), 2 —Equation (5), 3 —Equation (6); calculated,
according to: 4—Equation (3) ne = 2, 5—Equation (3) n N = 10.

Figure 12. Comparison of the probabilistic low-cycle fatigue curves calculated under strain-controlled
loading with the experimental ones for steel C45; straight lines—calculated, dashed lines– exper-
iment; normalized characteristic according to: 1—Equation (3), 2—Equation (5), 3—Equation (6);
reference characteristics, according to: 1 —Equation (3), 2 —Equation (5), 3 —Equation (6); calculated,
according to: 4—Equation (3) if ne = 2, 5—Equation (3) if nN = 10.

Table A4 provides the values of ratios K of the highest and lowest mechanical charac-
teristics σpr , σys , σu , σ f , ψ, ψu .
The design of the last curves employed the low-cycle fatigue curves designed accord-
ing to the reference mechanical characteristics and the dependencies providing the most
accurate description of experimental durability, i.e., Equation (3) for steel 15Cr2MoVA and
Equation (6) for steel C45.
As suggested in Figure 11, for steel 15Cr2MoVA, the low-cycle fatigue curves deter-
mined according to Equations (3), (5) and (6) using normalized mechanical characteristics

349
Materials 2022, 15, 8808

are above the experimental curve of probability of failure 99%. The same curves designed
by using the reference mechanical characteristics are located in the durability band between
curves 1% and 50%. This was predictable, as the normalized mechanical characteristics
of the steel 15Cr2MoVA are close to the experimental mechanical characteristics with a
probability of 12% to 25%, while the reference mechanical characteristics are located in
the probability band of the experimental characteristics of 0.0003 to 74%. Due to the ‘high’
layout of the calculated low-cycle fatigue curves compared to the experimental ones, the
curves designed using safety factors ne = 2 and n N = 10 are also positioned fairly high. The
low-cycle fatigue curve designed using ne = 2 virtually corresponds to the experimental
curve of failure probability 1%, while the curve designed using n N = 10 is located below.
Another situation is presented in Figure 12 that shows the listed low-cycle fatigue
curves for steel C45. In this case, the calculated curves designed using both the normalized
and the reference mechanical characteristics are located considerably lower than the experi-
mental ones. However, this is the consequence of poor correspondence of the calculated
curves designed according to Equation (6) with the experimental curves for the steel C45
(Figure 9c).

3.3. Case Study Objective


To Determine the Probabilistic Values of Cumulative Durability Damage (at the Crack
Initiation Phase) for the Zone of Sleeve Connection to the Vessel Body under Hydraulic
Forging at Temperature 20◦ C. The Nominal Strain Range for the Outer Surface of the Sleeve
Connection under Hydraulic Forging: ε 1n = 0, ε 2n = 0.54%, ε 3n = −0.97%. Concentration
Factor (Theoretical) of Elastic Stress ασ = 3, Strain-Controlled Loading Mode, Vessel
Material—Grade 15Cr2MoVA Steel.
Mechanical and cyclic characteristics determined during the course of the study
(Table A1 and Table A5) were used in the calculations.
All calculations were carried out for the failure probability of 1%, 10%, 30%, 50%,
70%, 90%, 99%. The case study presents the calculation for probability 1%. For other
probabilities, the calculated values are presented in A6. The calculation was carried out
as follows.
Hardening rate m0 (Table A5) was used to design the tensile stress-strain diagrams
σ − e. Their linear approximation resulted in following relative linear hardening modules:

σ−1
GT = , (22)
e−1
2.06 − 1
GT = = 0.0272
40 − 1
Parameter χ1 characterizing the sensitivity to cycle asymmetry was determined ac-
cording to the following dependency:
σys
χ1 = 2.7ψ − 1, (23)
σu
300
χ1 = 2.7 · 0.74 − 1 = 0.1988
500
χ2 = 0.23( σσysu − 1) + χ1 ,
(24)
χ2 = 0.23( 300
500
− 1) + 0.1988 = 0.3521
The coefficient of the cycle asymmetry intensity of the nominal stresses was assumed
to be rσn = −1.05, and the range of intensity of nominal stresses for the first semi-cycle of
loading was determined according to the following dependency:
!
2GT εinK S T − B · A1;2 · F (k)
SinK = , (25)
S T [2GT + p1;2 · A1;2 · F (k )]

350
Materials 2022, 15, 8808

2 · 0.0272[9.37 · 1.16 − 37.19 · 0.23 · 1]


SinK =
1.15[2 · 0.0272 + 1.0048 · 0.23 · 1]
where
1 + rσ n
p1;2 = 1 + χ1;2 , (26)
1 − rσ n
1 + 1.05
p1 = 1 + 0.1988 = 1.0048,
1 − 1.05
1 S
B = 1− − T, (27)
GT 2
1 1.15
B = 1− − = −37.19
0.0272 2
Hence, to calculate the Poisson’s ratio in normal section μn for the first semi-cycle of
loading, it was necessary to use the value of the relative linear hardening module gn1 , and
in the first approximation after replacement of the sign of main strains with the opposite
sign, the following was assumed:
ε in1
εin1 = ε1n1 , ε1n1 = , (28)
e pr S T

0.97
εin1 = = 9.37
0.09 · 1.15
Linear hardening module:
Sin1 − 1
gn1 = , (29)
ε in1 − 1
3.0231 − 1
gn1 = = 0.2417
9.37 − 1
and, accordingly, the Poisson’s ratio:

1 − gn1 (εin1 − 1)
μn1 = 0.5 − 0.2 , (30)
εin1

1 − 0.2417(9.37 − 1)
μn1 = 0.5 − 0.2 = 0.4355
9.37
Intensity of nominal strains of the first approximation were determined according to
dependency [34]:

2
ε in1 = ( e1 − e2 ) 2 + ( e2 − e3 ) 2 + ( e3 − e1 ) 2 , (31)
2(1 + μn1 )

2
ε in1 = (0.97 − 0.54)2 + (0.54)2 + (−0.97)2 = 0.5764
2(1 + 0.4355)
In relative units:
0.5864
ε in1 = = 5.6657
0.09 · 1.15
The obtained value and dependencies were used to repeat the calculation and deter-
mine final values Sin1, gn1 , μn1 and εin1 (Table A5). If symmetrical cycle of strain intensity
in the nominal section was accepted for the task considered, then:

εin1 S T
ein = (32)
1 − ren

351
Materials 2022, 15, 8808

The coefficient of asymmetry of the intensity cycle of nominal strain was assumed as
ren = −1, then:
4.8279 · 1.15
ein = = 2.7761
1+1
The range of main stress in the first semi-cycle of loading in the nominal section was
determined according to the following dependencies:

ε 1n1 + μn1 ε 2n1  ε + μn1 ε 1n1 


S1n1 = E 1 , S2n1 = 2n1 E1 (33)
1 − μ2n1 1 − μ2n1

Sin1 2.3744 · 205


E 1 = = = 112, 023MPa
ε in1 4.8279 · 0.0009
0.0097 + 0.4151 · 0.0054
S1n1 = · 112, 023 = 1616MPa,
1 − 0.41512
0.054 + 0.4151 · 0.0097
S2n1 = · 112, 023 = 1276MPa
1 − 0.41512
In accordance with the ranges of main stress available, the strain-controlled stiffness
coefficients were determined according to the dependency:

(S1n1 − S2n1 )2 + (S2n1 − S3n1 )2 + (S3n1 − S1n1 )2
De = √ , (34)
2(σ1 + σ2 + σ3 )

(1616 − 1276)2 + (1276)2 + (−1616)2
De = √ = 0.5103
2(1616 + 1276)
By using ein , α B , ren , χ1 , χ2 , GT (Table A5) and A1 , A2 , ST , β as well as the Matlab
programme of calculations designed according to the dependencies in papers [32], the
data characterizing the stress-strain state in the zone of maximum concentration were
determined (Table A6). In the durability calculation, the same as in relation to the durability
calculation norms [33], the safety factor for maximum deformations ne = 2. was accepted.
Then, the cumulative quasi-static damage [34]:

2( e i − σ i )
dk = , (35)
e B De

Based on study [38], e B = m0 was assumed:

m0 0.15
eB = = = 166.6
e pr 0.0009

then
2 · (15.9 − 1.3405)
dk = = 0.3424, and dy = 1 − dk = 0.6576
166.6 · 0.5103
The number of semi-cycles k0 , within which fatigue damage dy = 0.6576 would be
cumulated was determined according to the following dependencies:

k0 α3
δik δik
∑ De ( De + Sik S T )
1
dy = α (36)
C2 C3 3

Durability k0 calculations were conducted by using the following two techniques:

352
Materials 2022, 15, 8808

(a) in the calculations, probabilistic values of characteristics A1 , A2 , χ1 , χ2 , S T , GT , β, De , C2 , C3


and α3 were used:

k0 0.98
2δik 2δik
∑ ( + Sik · 1.28)
1 0.5103 0.5103
0.7259 = , k = 12.5 or N = 6
500 · 2860.98

(b) in the calculations, the values of probability 50% of characteristics A1 , A2 , χ1 , χ2 , S T , GT ,


β, De and probabilistic values of parameters C2 , C3 and α3 were used:
0.98
k0 δik 2δik
∑ ( + Sik · 1.15)
1 0.5103 0.5103
0.6576 = ,
500 · 2860.98
The results of probabilistic calculations are presented in Table 4.

Table 4. Results of probabilistic calculation of life in the concentration zone.

Probability, %
Parameter
1 10 30 50 70 90 99
k0 12.5 21.5 52.0 86.0 201.0 287.0 389.0
k0 59.5 66.5 78.5 86.0 98.5 123.0 171.0

The total number of start-stop operations of the system for the vessel considered
expected during the lifetime is equal to 25 [33]. Hence, fatigue cracks may appear in the
sleeve-to-body connection during the lifetime of the vessel considered at failure probability
30% according to calculation (a) and at failure probability lower than 1% according to
calculation (b).

4. Conclusions
The methodology of conduction of an integrated experiment in a probabilistic setting
to investigate the parameters of resistance to cyclic deformation and failure for the materials
representing major types of cyclic properties (hardening, softening, stabilization).
The investigations presented in the paper point to the correlation relationship between
mechanical characteristics and durability as well as the cyclic deformation that is close to the
linear correlation. This supports the correctness and physical rationale of the mathematical
dependencies proposed by different authors for describing the low-cycle deformation
and failure process under strain-controlled loading. The regression coefficients allow
for the calculation of preliminary cyclic characteristics and durability using the available
mechanical characteristics.
The analysis of mutual layout of the calculated and experimental probabilistic low-
cycle fatigue curves conducted in the study once again has demonstrated that the design of
low-cycle fatigue curves according to the both analytical dependencies for specific materials
may lead to considerable errors. Hence, in calculations of critical structures, it is necessary
to have at least an experimental curve of failure probability 50% to conduct a reliable
assessment of the strength and durability of the structure considered.
Investigations have pointed at the presence of a stable correlation relationship between
mechanical characteristics and durability as well as the parameters of cyclic deformation
parameters. This supports the physical rationale of the phenomenological dependencies
proposed by different authors for the description of the low-cycle deformation and failure
process under strain-controlled loading by using mechanical characteristics.
The above investigation has suggested that probabilistic curves should not be built
in the relative coordinates using the probabilistic e pr , if scatter e pr is considerably larger
than scatter σys , σu and ψ. Equations (3), (5) and (6) are not universal, since, for steel

353
Materials 2022, 15, 8808

15Cr2MoVA, the best correspondence is provided by Equations (3) and (5), while for steel
C45 Equation (6).
The statistical investigations conducted in the paper have shown that that phenomeno-
logical dependencies used at present for the description of low-cycle fatigue curves on the
basis of mechanical characteristics are not universal for the materials with contrasting cyclic
properties, and a reliable assessment of the durability of structure requires the probabilistic
experimental curve 50%. For example, the calculated low-cycle fatigue curves under strain-
controlled loading defined using the safety factor [31] (ne = 2, Nn = 10) and the calculated
curves that provide more accurate description of the experimental data provide the follow-
ing strength safety margins in comparison to the experimental curves: ne = 0.45, Nn = 4.10
(for steel 15Cr2MoVA) and ne = 4.85, Nn = 100 (for steel C45).

Author Contributions: Conceptualization, methodology, validation, writing—original draft prepara-


tion, investigation, writing—review and editing Ž.B. and V.L.; software, visualization V.L.; formal
analysis, data curation, resources project administration Ž.B. All authors have read and agreed to the
published version of the manuscript.
Funding: This research received no external funding.
Acknowledgments: All co-authors have contributed equally. The authors declare that they have no
known competing financial interests or personal relationships that could have appeared to influence
the work reported in this paper.
Conflicts of Interest: The authors declare no conflict of interest.

Nomenclature
Latin symbols
constants of the low–cycle fatigue curve under strain-controlled loading describing
A1 , A2
the first and second semi-cycle form respectively;
ST
B parameter ( B = 1 − G1T − 2 );
b, b yx , b y regression coefficients;
x
C2 , C3 , ∝, c constants of the low–cycle fatigue curve under strain-controlled loading;
C1p , Cψ constants of the Coffin–Manson equation;
De strain-controlled stiffness coefficient;
dk cumulated quasi-static damage;
dy cumulated fatigue damage;
E modulus of elasticity (MPa);
E1 secant modulus for the first semi-cycle diagram (MPa);
linear hardening modules for single and cyclic (k-th semi-cycle) loading respectively
ET , ETk
(GPa);
e strain intensity under linear strain-controlled state (%);
e1 , e2 , e3 principal strain (%);
ei strain intensity (%);
ein nominal strain intensity (%);  
e, ei , ein respective deformations in relative units ei = ei /e pr , ein = ein /e pr , (%);
intensity of maximum strain at the initial loading under the linear strain-controlled
e0
state (%);
intensity of maximum strain at the initial loading normalized to proportional limit
e0
strain;
ea total strain amplitude (%);
ey amplitude of elastic strain (%);
ep amplitude of plastic strain (%);
e pr proportional limit strain (%);
material plasticity indicator determined by assessment of the variation in the cross
et
-section area of the standard cylindrical specimen subjected to tension;
i = 1...n specimen ranks in the rank order;
F (k) function of k semi-cycles [36];

354
Materials 2022, 15, 8808

relative linear hardening modules for single and cyclic loading diagrams respectively
GT , gni
( GT = ET /E, gk = ETk /E), (GPa);
k number of loading semi-cycle;
k0 number of semi-cycles of the loading cycles to fatigue crack initiation;
K coefficient K values of relative measures of the key mechanical properties;
m constant of the Coffin–Manson equation;
m0 hardening parameter for the tensile diagram under its stepped approximation;
m xy second central mixed moment;
N0 number of loading cycles to fatigue crack initiation;
Nc durability (number of load cycles) until crack initiation;
Nf durability (number of load cycles) till the cracks propagated to complete rupture;
n number of of measurements;
ne safety factor of strength by maximum deformations;
nN safety factor of strength by the number of cycles;
p1;2 parameter ( p1;2 = 1 + χ1;2 11+ rσn
−rσn );
r coefficient of correlation;
ren coefficient of asymmetry of the intensity cycle of nominal;
rσn coefficient of asymmetry of the intensity cycle of nominal stress;
cyclic stress intensity under linear strain-controlled state, calculated from the start of
S
unloading (MPa);
Si cyclic stress intensity (MPa);
Sin nominal cyclic stress intensity (MPa);
Sk, Sik, Sink respective cyclic strain for the k-th semi-cycle
 of loading (MPa); 
respective cyclic strains in relative units Sk = Sk /ST , Sik = Sik /ST , Sik = Sink /ST ,
Sk, Sik, Sink
(MPa);
ST cyclic limit of proportionality calculated from the start of unloading
 (MPa);
ST cyclic stress normalized respectively to proportional limit stress S T = ST /σpr ) , (MPa);
tγ Student’s t-distribution;
x, y statistical mean;
xi , yi independent variables;
maximum values of the material mechanical properties in the bins (statistical
xmax
intervals);
minimum values of the material mechanical properties in the bins (statistical
xmin
intervals);
Greek symbols
constants of the low–cycle fatigue curve under strain-controlled loading characterizing
α, β
materials hardening or softening;
α1p constant of the Coffin–Manson equation;
ασ theoretical coefficient of concentration of elastic stresses;
γ reliability of normal distribution;
intensity of cyclic plastic deformations under linear strain-controlled states or
δ
width of the elastoplastic hysteresis loop;
ε1n , ε2n , ε3n range of main nominal strains (%);
cyclic elastoplastic strain intensity under linear strain-controlled state, calculated
ε
from the start of unloading (%);
εi cyclic elastoplastic strain intensity (%);
ε in nominal cyclic elastoplastic strain intensity (%);
ε k, ε ik, ε ink respective cyclic strains for the k-th semi-cycle of loading (%);
εk, εik, εink respective cyclic strain in relative units (εk = ε k /ε T , εik = ε ik /ε T , εink = ε ink /ε T );
εT strain corresponding to ST (%);
εT strain corresponding to S T (%);
μni Poisson’s ratio for elastoplastic deformation;
σx , σy root mean square deviation;
σa∗ conditional elastic stresses (σa∗ = e0 Eet ), (MPa);
σpr proportional limit stress (MPa);
σys elastic limit or yield strength, the stress at which 0.2% plastic strain occurs (MPa);
σu ultimate tensile stress (MPa);
σf fracture strength (MPa);

355
Materials 2022, 15, 8808

σ normalized to proportional limit stress (MPa), σ = σ/σpr ;


χ1 , χ2 parameter characterizing the sensitivity to cycle asymmetry;
ψ cross-sectional narrowing (%);
ψu continuous cross-sectional narrowing (%).

Appendix A

Table A1. Probabilistic values of mechanical characteristics for steel 15Cr2MoVA.

Probability, %
Characteristic
1 10 30 50 70 90 99
σys , MPa 300 340 370 400 430 475 535
σu , MPa 500 530 560 580 600 640 680
ψ, % 74 76 79 80 82 85 90
e pr , % 0.090 0.130 0.170 0.200 0.245 0.320 0.475

Table A2. Probabilistic values of mechanical characteristics for steel C45.

Probability, %
Characteristic
1 10 30 50 70 90 99
σys , MPa 220 265 300 340 360 420 500
σu , MPa 620 700 750 800 850 900 1020
ψ, % 28 32 37 39 42 47 54
e pr , % 0.140 0.180 0.225 0.260 0.300 0.360 0.480

Table A3. Probabilistic values of mechanical characteristics for aluminium alloy D16T1.

Probability, %
Characteristic
1 10 30 50 70 90 99
σys , MPa 260 300 320 350 370 405 460
σu , MPa 580 620 650 680 700 750 800
ψ, % 9.5 11.3 12.8 14.0 15.5 17.5 21.0
e pr , % 0.46 0.52 0.56 0.60 0.64 0.70 0.78

Table A4. Coefficient K values of relative measures of the key mechanical properties.

σ pr σ ys σu σf ψ ψu
Material
MPA %
15Cr2MoVa 2.06 1.88 1.52 1.78 1.24 4.19
C45 2.41 2.41 1.57 1.62 1.71 3.09
D16T1 1.72 1.66 1.45 1.43 2.24 1.38

356
Materials 2022, 15, 8808

Table A5. Probabilistic values of parameters of cyclic deformation resistance for steel 15Cr2MoVA.

Probability, %
Parameter
1 10 30 50 70 90 99
A1 0.23 0.87 1.32 1.60 1.94 2.40 3.04
A1 0.30 0.94 1.34 1.64 2.00 2.43 3.06
ST 1.15 1.20 1.25 1.28 1.30 1.35 1.40
m0 0.15 0.17 0.19 0.21 0.23 0.26 0.30

Table A6. Calculation data in the concentration zone for failure probabilities 1%, 10%, 30%,50%, 70%,
90%, 99%.

Probability, %
Parameter
1 10 30 50 70 90 99
GT 0.0272 0.0328 0.0400 0.0451 0.0497 0.0605 0.0754
χ1 0.1988 0.3164 0.4093 0.4896 0.5867 0.7033 0.9118
χ2 0.3521 0.4449 0.5274 0.5931 0.6776 0.7832 0.9741
p1 1.0048 1.0077 1.0099 1.0119 1.0143 1.0171 1.0222
B −37.19 −30.08 −24.62 −21.81 −20.06 −16.20 −12.96
ε1n1 9.37 6.16 4.56 3.79 3.04 2.26 1.46
S1n1 3.0231 1.9699 1.7305 1.6435 1.5871 1.4779 1.3771
gn1 0.2417 0.1879 0.2049 0.2300 2.2870 0.3786 0.8223
ε in1 0.4355 0.4360 0.4242 0.4133 0.3958 0.3693 0.3111
ein 2.7761 0.23073 1.9606 1.5412 1.2836 1.0122 0.7181
β 0.00004 0.00016 0.00042 0.00085 0.0018 0.0048 0.0190
E1 112,023 92,705 87,001 90,992 94,974 98,169 105,772
S1n1 1616 1315 1199 1226 1236 1220 1227
S2n1 1276 1031 929 939 931 895 858
De 0.5103 0.5109 0.5119 0.5130 0.5146 0.5174 0.5230
ασ 3 3 3 3 3 3 3
ren −1 −1 −1 −1 −1 −1 −1
dk 0.3424 0.3443 0.2976 0.2741 0.2551 0.1232 0.2197
dy 0.6576 0.6557 0.7024 0.7259 0.7449 0.7802 0.8768

References
1. Makhutov, N.A.; Gadenin, M.M. Integrated Assessment of the Durability, Resources, Survivability, and Safety of Machinery
Loaded under Complex Conditions. J. Mach. Manuf. Reliab. 2020, 49, 292–300. [CrossRef]
2. Makhutov, N.A.; Panov, A.N.; Yudina, O.N. The development of models of risk assessment complex transport systems. In IOP
Conference Series: Materials Science and Engineering, Proceedings of the V International Scientific Conference, Survivability and Structural
Material Science (SSMS 2020), Moscow, Russia, 27–29 October 2020; Institute of Physics Publishing (IOP): Moscow, Russia, 2020;
Volume 1023. Available online: https://scholar.google.lt/scholar?hl=lt&as_sdt=0%2C5&q=The+development+of+models+of+
risk+assessment+complex+transport+systems&btnG= (accessed on 16 July 2022).
3. Sekhar, A.P.; Nandy, S.; Bakkar, M.A.; Ray, K.K.; Das, D. Low cycle fatigue response of differently aged AA6063 alloy: Statistical
analysis and microstructural evolution. Materialia 2021, 20, 101219. [CrossRef]
4. Jiang, Z.; Han, Z.; Li, M. A probabilistic model for low-cycle fatigue crack initiation under variable load cycles. Int. J. Fatigue 2022,
155, 106528. [CrossRef]
5. Li, X.-K.; Chen, S.; Zhu, S.-P.; Ding Liao, D.; Gao, J.-W. Probabilistic fatigue life prediction of notched components using strain
energy density approach. Eng. Fail. Anal. 2021, 124, 105375. [CrossRef]
6. Weibull, W. Fatigue Testing and Analysis of Results; Pergamon Press: New York, NY, USA, 1961. Available online:
https://books.google.lt/books?hl=lt&lr=&id=YM4gBQAAQBAJ&oi=fnd&pg=PP1&dq=Weibull,+W.+Fatigue+Testing+
and+Analysis+of+Results&ots=VIVGA6VzJY&sig=UsrDkGvsP8go6hPS6IuYK3q9-H0&redir_esc=y#v=onepage&q=Weibull%
2C%20W.%20Fatigue%20Testing%20and%20Analysis%20of%20Results&f=false (accessed on 20 July 2022).

357
Materials 2022, 15, 8808

7. Weibull, W.; Rockey, K.C. Fatigue Testing and Analysis of Results. J. Appl. Mech. 1962, 29, 607. [CrossRef]
8. Coffin, L.F., Jr. A study of the effects of cyclic thermal stresses on a ductile metal. Trans. Metall. Soc. ASME 1954, 76, 931–950.
[CrossRef]
9. Manson, S.S. Fatigue: A complex subject—Some simple approximations. Exp. Mech. 1965, 7, 193–225. [CrossRef]
10. Langer, B.F. Design of Pressure Vessels for Low-Cycle Fatigue. ASME J. Basic Eng. 1962, 84, 389–399. [CrossRef]
11. Iida, K.; Inoue, H. Evaluation of low cycle fatigue design curve based on life distribution shape. J. Soc. Nav. Archit. Jpn. 1973, 133,
235–247. [CrossRef]
12. Freudenthal, A.M.; Gumbel, E.J. On the statistical interpretation of fatigue tests. Proc. R. Soc. Lond. Ser. A 1953, 216, 309–332.
[CrossRef]
13. Freudenthal, A.M.; Gumbel, E.J. Physical and Statistical Aspects of Fatigue. Adv. Appl. Mech. 1956, 4, 117–158. [CrossRef]
14. Serensen, S.V.; Shneiderovich, R.M. Deformations and rupture criteria under low-cycles fatigue. Exp. Mech. 1966, 6, 587–592.
[CrossRef]
15. Serensen, S.V.; Kogayev, V.P.; Shneiderovich, R.M. Load Carrying Ability and Strength Evaluation of Machine Components, 3rd ed.;
Mashinostroeniya: Moscow, Russia, 1975; pp. 255–311. (In Russian)
16. Stepnov, M.N. Statistical Methods for Computation of the Results of Mechanical Experiments, Mechanical Engineering; Mashinostroeniya:
Moscow, Russia, 2005. (In Russian)
17. Makhutov, N.A.; Zatsarinny, V.V.; Reznikov, D.O. Fatigue prediction on the basis of analysis of probabilistic mechanical properties.
AIP Conf. Proc. 2020, 2315, 040025. [CrossRef]
18. Daunys, M.; Šniuolis, R. Statistical evaluation of low cycle loading curves parameters for structural materials by mechanical
characteristics. Nucl. Eng. Des. 2006, 236, 13. [CrossRef]
19. Daunys, M.; Bazaras, Z.; Timofeev, B.T. Low cycle fatigue of materials in nuclear industry. Mechanics 2008, 73, 12–17. Available
online: https://scholar.google.lt/scholar?hl=lt&as_sdt=0%2C5&q=28.%09Daunys%2C+M.%3B+Bazaras%2C+Z.%3B+Timofeev%
2C+B.+T.+Low+cycle+fatigue+of+materials+in+nuclear+industry&btnG= (accessed on 22 December 2021).
20. Timofeev, B.T.; Bazaras, Z.L. Cyclic strength of the equipment of nuclear power plants made of 22K steel. Mater. Sci. 2005, 41,
680–685. [CrossRef]
21. Raslavičius, L.; Bazaras, Ž.; Lukoševičius, V.; Vilkauskas, A.; Česnavičius, R. Statistical investigation of the weld joint efficiencies
in the repaired WWER pressure vessel. Int. J. Press. Vessel. Pip. 2021, 189, 1–9. [CrossRef]
22. Guo, S.; Liu, R.; Jiang, X.; Zhang, H.; Zhang, D.; Wang, J.; Pan, F. Statistical Analysis on the Mechanical Properties of Magnesium
Alloys. Materials 2017, 10, 1271. [CrossRef]
23. Zhu, S.-P.; Huang, H.-Z.; Smith, R.; Ontiveros, V.; He, L.-P.; Modarres, M. Bayesian framework for probabilistic low cycle fatigue
life prediction and uncertainty modeling of aircraft turbine disk alloys. Probabilistic Eng. Mech. 2013, 34, 114–122. [CrossRef]
24. Strzelecki, P. Determination of fatigue life for low probability of failure for different stress levels using 3-parameter Weibull
distribution. Int. J. Fatigue 2021, 145, 106080. [CrossRef]
25. Fekete, B. New energy-based low cycle fatigue model for reactor steels. Mater. Des. 2015, 79, 42–52. [CrossRef]
26. Bazaras, Ž.; Lukoševičius, V. Statistical Assessment of Low-Cycle Fatigue Durability. Symmetry 2022, 14, 1205. [CrossRef]
27. Chen, L.; Wang, D.S.; Shi, F.; Sun, Z.G. Low-Cycle Fatigue Properties of Austenitic Stainless Steel S30408 under Large Plastic
Strain Amplitude. Adv. Steel Constr. 2022, 18, 517–527. [CrossRef]
28. Pelegatti, M.; Lanzutti, A.; Salvati, E.; Srnec Novak, J.; De Bona, F.; Benasciutti, D. Cyclic Plasticity and Low Cycle Fatigue of
an AISI 316L Stainless Steel: Experimental Evaluation of Material Parameters for Durability Design. Materials 2021, 14, 3588.
[CrossRef] [PubMed]
29. Lee, S.J.; Theerthagiri, J.; Nithyadharseni, P.; Arunachalam, P.; Balaji, D.; Kumar, A.M.; Madhavan, J.; Mittal, V.; Choi, M.Y.
Heteroatom-doped graphene-based materials for sustainable energy applications: A review. Renew. Sustain. Energy Rev. 2021,
143, 110849. [CrossRef]
30. Lee, Y.; Yu, Y.; Das, H.T.; Theerthagiri, J.; Lee, S.J.; Min, A.; Kim, G.-A.; Choi, H.C.; Choi, M.Y. Pulsed laser-driven green synthesis
of trimetallic AuPtCu nanoalloys for formic acid electro-oxidation in acidic environment. Fuel 2023, 332 Pt 2, 126164. [CrossRef]
31. GOST 25502-79 Standard; Strength Analysis and Testing in Machine Building. Methods of Metals Mechanical Testing. Methods of
Fatigue Testing. Standardinform: Moscow, Russia, 1993.
32. GOST 22015–76 Standard; Quality of Product. Regulation and Statistical Quality Evaluation of Metal Materials and Products on
Speed-torque Characteristics. Standardinform: Moscow, Russia, 2010.
33. Regularities and Norms in Nuclear Power Engineering (PNAE) No. G-7-002-89. In Rules of Equipment and Pipelines Strength
Calculation of Nuclear Power Plant; Energoatomizdat: Moscow, Russia, 1989.
34. Daunys, M. Cycle Strength and Durability of Structures; Technologija: Kaunas, Lithuania, 2005. (In Lithuanian)
35. Montgomery, D.C.; Runger, G.C. Applied Statistics and Probability for Engineers, 7th ed.; Wiley: Hoboken, NJ, USA, 2018.
36. Čekanavičius, V.; Murauskas, G. Statistics and Its Applications: I Part; TEV: Vilnius, Lithuania, 2002; pp. 129–133. (In Lithuanian)
37. Pavaras, A.; Žvinys, J. Steels; Technologija: Kaunas, Lithuania, 1995. (In Lithuanian)
38. Makhutov, N.A. Strength analysis under long-term static and cyclic loading. In Structural Durability, Resource and ASafety; Nauka:
Novosibirsk, Russia, 2005; Volume 1, pp. 223–256. (In Russian)

358
MDPI
St. Alban-Anlage 66
4052 Basel
Switzerland
Tel. +41 61 683 77 34
Fax +41 61 302 89 18
www.mdpi.com

Materials Editorial Office


E-mail: [email protected]
www.mdpi.com/journal/materials
Academic Open
Access Publishing

www.mdpi.com ISBN 978-3-0365-7601-5

You might also like