Crystals: Low Cycle Fatigue Behavior of TC21 Titanium Alloy With Bi-Lamellar Basketweave Microstructure

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

crystals

Article
Low Cycle Fatigue Behavior of TC21 Titanium Alloy with
Bi-Lamellar Basketweave Microstructure
Baohua Nie 1,2 , Yu Song 1,2 , Xianyi Huang 1,2 , Haiying Qi 1,2 , Zihua Zhao 3, * and Dongchu Chen 1,2, *

1 School of Materials Science and Hydrogen Energy, Foshan University, Foshan 528000, China;
niebaohu@fosu.edu.cn (B.N.); 2112056006@stu.fosu.edu.cn (Y.S.); 20200580429@stu.fosu.edu.cn (X.H.);
haiyingqi@fosu.edu.cn (H.Q.)
2 Guangdong Key Laboratory for Hydrogen Energy Technologies, Foshan 528000, China
3 School of Materials Science and Engineering, Beihang University, Beijing 100191, China
* Correspondence: zhzh@buaa.edu.cn (Z.Z.); chendc@fosu.edu.cn (D.C.); Tel.: +86-010-8233-3264 (Z.Z.);
+86-075-7827-00525 (D.C.)

Abstract: Low cycle fatigue (LCF) behaviors of TC21 alloy with a bi-lamellar basketweave microstruc-
ture were investigated in this paper. The strain fatigue tests were carried out at total strain amplitudes
of 1.4% to 2.0%. The cyclic stress response showed the cyclic softening behavior. In addition, the
shape of the hysteresis rings exhibited a non-Masing model behavior. The cyclic stress–strain as
well as the strain-life equations were obtained. The fatigue life decreased significantly with an
increasing total strain from 1.4% to 2.0%. The cyclic softening behavior was interpreted by cyclic
back stress and friction stress. Low cycle fatigue cracks were predominantly initiated on the surface
of the samples. The relationship between the fatigue sub-critical crack and microstructure was also
discussed. The cyclic deformation behavior and crack initiation mechanism were revealed on the
basis of the deformation microstructure under different strain amplitudes.
Citation: Nie, B.; Song, Y.; Huang, X.;
Qi, H.; Zhao, Z.; Chen, D. Low Cycle
Keywords: low cycle fatigue; titanium alloy; basketweave microstructure; crack initiation
Fatigue Behavior of TC21 Titanium
Alloy with Bi-Lamellar Basketweave
Microstructure. Crystals 2022, 12, 796.
https://doi.org/10.3390/ 1. Introduction
cryst12060796 Titanium alloys are generally used in landing gear and other key load-bearing compo-
Academic Editors: Yuanfei Han, nents in aircrafts [1,2]. As critical load-bearing components, titanium alloy components
Zhonggang Sun, Hong Li, usually suffer a high cyclic loading during service, which may result in a low cycle fatigue
Lechun Xie and Cyril Cayron (LCF) fracture. Thus, LCF behavior and properties are essential for the safety and reliability
of titanium alloy components.
Received: 24 April 2022
The microstructure of titanium alloys significantly influences the LCF behavior and
Accepted: 21 May 2022
properties [3–8]. Lei [9] investigated the LCF properties of TA15 titanium alloy with
Published: 2 June 2022
a tri-modal microstructure, Widmanstatten microstructure and bimodal microstructure,
Publisher’s Note: MDPI stays neutral and found that the fatigue properties of the tri-modal microstructure were higher than
with regard to jurisdictional claims in that of the Widmanstatten microstructure and were equivalent to that of the bimodal
published maps and institutional affil- microstructures. The fatigue crack propagation was more tortuous and had a higher fatigue
iations. crack propagation resistance than that of bimodal microstructures. Xu [10] suggested that
titanium alloys with lamellar and bimodal microstructures had a different LCF behavior
due to their response to the cyclic strain. As for titanium alloys with a tri-modal [11]
and bimodal [12,13] microstructure, LCF cracks were initiated from the coarsened slip
Copyright: © 2022 by the authors.
band in the soft αp phase. However, the shear deformation and spheroidization of β
Licensee MDPI, Basel, Switzerland.
lamellar contributed to the LCF crack initiation for titanium alloys with a Widmanstatten
This article is an open access article
microstructure [14]. Furthermore, the dislocation evolution and the deformation twins
distributed under the terms and
conditions of the Creative Commons
were considered as important factors for the cyclic softening of titanium alloys with a
Attribution (CC BY) license (https://
lamellar microstructure [15,16].
creativecommons.org/licenses/by/ TC21 titanium alloy was used in aircraft key load-bearing components owing to its
4.0/). high strength and damage tolerance. TC21 titanium alloy was manufactured by a quasi-β

Crystals 2022, 12, 796. https://doi.org/10.3390/cryst12060796 https://www.mdpi.com/journal/crystals


Crystals 2022,
Crystals 12,12,
2022, 796x FOR PEER REVIEW 2 of 11 2 of 11

TC21
forging titanium
process toalloy
obtainwastheused in aircraft key
basketweave load‐bearing components
microstructure [1]. The LCF owing to its of TC21
behavior
high strength and damage tolerance. TC21 titanium alloy was manufactured
titanium alloy has attracted much attention. Du [17] reported that the residual compressiveby a quasi‐β
forging process to obtain the basketweave microstructure [1]. The LCF behavior of TC21
stress of both surface shot peening and high velocity oxygen fuel (HVOF) improved the
titanium alloy has attracted much attention. Du [17] reported that the residual compres‐
LCF properties of TC21 alloy. Yu [18] found that TC21 alloy with a bimodal microstructure
sive stress of both surface shot peening and high velocity oxygen fuel (HVOF) improved
illustrated the characteristic
the LCF properties of TC21 alloy. of Yu
cyclic
[18] softening.
found that TC21The research
alloy withofa Tan [19] micro‐
bimodal suggested that
the low cycle fatigue property of a bimodal structure was higher than
structure illustrated the characteristic of cyclic softening. The research of Tan [19] sug‐ that of a lamellar
structure,
gested that and thatcycle
the low the fatigue
low cycle cracks
property ofinitiated
a bimodalfrom the slip
structure was band
higherinthan
thethat
αp ofphase for a
bimodal
a lamellarmicrostructure andthe
structure, and that thelow phase
α/βcycle interface
cracks for from
initiated a lamellar microstructure.
the slip band in the αp However,
few
phaseinvestigations on cyclic deformation
for a bimodal microstructure and the α/β and fatigue
phase crack
interface fornucleation have been reported
a lamellar microstruc‐
ture.titanium
for However, few investigations
alloys on cyclicmicrostructure.
with a basketweave deformation andThe fatigue
LCFcrack nucleation
behavior of TC21 alloy
have been
with reported formicrostructure
a basketweave titanium alloys is with a basketweave
discussed in thismicrostructure.
paper. Cyclic The LCF be‐ and the
deformation
haviorinitiation
crack of TC21 alloy with a basketweave
mechanism are further microstructure
investigatedistodiscussed
supportinthe this paper. Cyclic
application of titanium
deformation and the crack initiation mechanism are further investigated to support the
alloy components.
application of titanium alloy components.
2. Materials and Methods
2. Materials and Methods
2.1. Materials
2.1. Materials
The nominal composition of TC21 titanium alloy used in this study is: Ti-6Al-2Zr-2Sn-
The nominal composition of TC21 titanium alloy used in this study is: Ti‐6Al‐2Zr‐
2Mo-2Nb-1.5Cr (wt%). TC21 alloy was manufactured by
2Sn‐2Mo‐2Nb‐1.5Cr (wt%). TC21 alloy was manufactured
a quasi-β forging process [1], and
by a quasi‐β forging process
the alloy was heat-treated as follows: 900 ◦ C/2 h + 600 ◦ C/4 h. A lamellar α (α ) and β
[1], and the alloy was heat‐treated as follows: 900 °C/2 h + 600 °C/4 h. A lamellar α (αL) L
transformed
and β transformedmatrix (β) (β)
matrix microstructure
microstructure are
areobserved (Figure1a),
observed(Figure 1a),and
andthethe mean
mean width of αL
width
is
of 3.4
αL isµm. Furthermore,
3.4 μm. Furthermore,thethe
fine lamellar
fine lamellarααphase,
phase, which is also
which is alsocalled
calledsecondary
secondary
αs αs phase,
is detected
phase, in theinβ the
is detected phase withwith
β phase a mean width
a mean of 500
width nmnm
of 500 (Figure
(Figure1b).1b).The
The samples
samples exhibit a
exhibit
high a highstrength
tensile tensile strength
of up toof 1070
up to MPa
1070 MPa
and and a high
a high yield
yield ratio
ratio (Figure2).
(Figure 2).

(a) (b)
Crystals 2022, 12, x FOR PEER REVIEW 3 of 11
Figure 1. Microstructures of TC21 titanium alloy: (a) optical micrograph (OM) and (b) backscattered
Figure 1. Microstructures of TC21 titanium alloy: (a) optical micrograph (OM) and (b) backscattered
electron micrograph (BEM).
electron micrograph (BEM).

Figure 2.
Figure 2. Tensile
Tensilestress–strain
stress–straincurve of TC21
curve titanium
of TC21 alloy.alloy.
titanium

2.2. Low Cycle Fatigue


LCF tests were carried out using a servo‐hydraulic testing machine (Instron 8801,
Norwood, MA, USA) under the condition of strain‐controlled conditions (Rε = −1), sinus‐
oidal waveforms, and a constant strain rate (dε/dt) of 4 × 10−3 s−1. A round bar specimen
Figure 2. Tensile stress–strain curve of TC21 titanium alloy.

Crystals 2022, 12, 796 3 of 11


2.2. Low Cycle Fatigue
LCF tests were carried out using a servo‐hydraulic testing machine (Instron 8801,
Norwood, MA, USA) under the condition of strain‐controlled conditions (Rε = −1), sinus‐
2.2. Low Cycle Fatigue
oidal waveforms, and a constant strain rate (dε/dt) of 4 × 10−3 s−1. A round bar specimen
was used for LCF
thetests
lowwere
cyclecarried
fatigueout using
test, a servo-hydraulic
as shown in Figure 3.testing machine
The total strains(Instron 8801, Nor-
were loaded
wood, MA, USA) under the condition of strain-controlled
at 1.4%, 1.6%, 1.8% and 2.0%. Each fatigue test was carried−out conditions (R ε = − 1), sinusoidal
until the cyclic stress de‐
waveforms, and a constant strain rate (dε/dt) of 4 × 10 3 s−1 . A round bar specimen was
creased to 90% of the fatigue stable stress, which was considered as the LCF life.
used for the low cycle fatigue test, as shown in Figure 3. The total strains were loaded at
1.4%, 1.6%, 1.8% and 2.0%. Each fatigue test was carried out until the cyclic stress decreased
to 90% of the fatigue stable stress, which was considered as the LCF life.

Figure 3. Geometric dimensions of specimens for the strain-controlled low cycle fatigue test.
Figure 3. Geometric dimensions of specimens for the strain‐controlled low cycle fatigue test.
The fracture surfaces and side crack morphology were observed by scanning electron
The fracture surfaces
microscopy (Ultim®and Max side
65,crack
Oxford,morphology
UK). A thin were observed
sample wasby cutscanning
close to electron
the fatigue
microscopy (Ultim ® Max 65, Oxford, UK). A thin sample was cut close to the fatigue frac‐
fracture and ground into 50–70 µm, then stamped into a small round sheet with a diameter
ture andof 3ground into 50–70
mm. Double μm, then thinning
jet electrolytic stampedwas intoadopted
a small in
round sheet withofa59
the electrolyte diameter of
pct methanol,
3 mm.35 Double jet electrolytic
pct n-butanol, thinning was
6 pct perchloric acidadopted in the
at a voltage of electrolyte
15 V and aof 59 pctof
current methanol,
15 mA. The
35 pct fatigue
n‐butanol,
dislocation structure was observed by Transmission electron microscopeThe
6 pct perchloric acid at a voltage of 15 V and a current of 15 mA. (TEM)
fatigue(JEM-2100,
dislocation structure
Tokyo, Japan).was observed by Transmission electron microscope (TEM)
(JEM‐2100, Tokyo, Japan).
3. Results and Discussion
3.1. Cyclic
3. Results Stress–Strain Behavior
and Discussion
The cyclic stress
3.1. Cyclic Stress–Strain curves with respect to different total strain of TC21 titanium alloy
Behavior
are shown in Figure 4. In the range of the total strain from 1.4% to 2%, the cyclic stress
The cyclic stress curves with respect to different total strain of TC21 titanium alloy4 of 11
Crystals 2022, 12, x FOR PEER REVIEW decreased rapidly at the initial stage of fatigue, illustrating the typical characteristic of
are shown
cyclicinsoftening.
Figure 4. ItInisthe range of
observed thethe
that total strain
cyclic from
stress 1.4% tofaster
decreased 2%, the
andcyclic stress
that the cyclic
decreased rapidly at the initial stage of fatigue, illustrating the typical
softening behavior of TC21 titanium alloy was more obvious with the increasing of characteristic of the
cyclic total
softening. It is observed that the cyclic stress decreased faster and
strain from 1.4% to 2%. Similar results have been reported by Xu [15] and Sen [16], that the cyclic
for whom
softening behaviortitanium
of TC21alloys with both
titanium alloya bimodal
was more microstructure
obvious with and
the lamellar microstructure
increasing of the
for whom titanium alloys with both a bimodal microstructure and lamellar microstructure
presented
total strain from cyclic
1.4% softening
to 2%. characteristic
Similar results at
have a high
been strain
reportedamplitude
by Xu due
[15] to
and the
Sen dislocation
[16],
presented cyclic softening characteristic at a high strain amplitude due to the dislocation
annihilation, twins in the α phase [15] and the spheroidization
annihilation, twins in the α phase [15] and the spheroidization of β lamellar [14]. of β lamellar [14].

Figure
Figure 4. The cyclic stress
stress curves
curvescorresponding
correspondingtotodifferent
different total
total strain.
strain.

The cyclic stress–strain relationship under LCF is usually described as [20]:

 
=K ' ( p )n ' (1)
2 2
Crystals 2022, 12, 796 4 of 11

Figure 4. The cyclic stress curves corresponding to different total strain.


The cyclic stress–strain relationship under LCF is usually described as [20]:
The cyclic stress–strain relationship under LCF is usually described as [20]:
0 n
∆σ = K' 0( ∆ε p
p n)
= K ( ) '
(1) (1)
22 22
where
where Δε∆εpp isisthe
thetotal
totalplastic
plasticstrain
strain range,
range, Δσ∆σ is the
is the total total stress
stress range,
range, and and K0 n′
K′ and and 0 are the
arenthe
cyclic strengthcoefficient
cyclic strength coefficientand andcyclic
cyclic strain
strain hardening
hardening index,
index, respectively.
respectively.
∆εppand
Δε andΔσ ∆σcan
canbe beobtained
obtainedbasedbasedononthe thecyclic
cyclic hysteresis
hysteresis loop
loop forfor a half
a half life.
life. According
Accord‐
to
ingEquation
to Equation (1),(1),
thethe
data areare
data linearly
linearly regressed
regressedusing usingdouble
doublelogarithmic coordinates. The
logarithmic coordinates.
obtained
The obtainedK0 value
K′ valueis is
1067.3
1067.3MPa
MPaand and n′ n0 value
valueisis0.07326;
0.07326; then,
then, we we obtain
obtain the expression
the expression
for the
the cycle
cycle stress–strain
stress–strainofofTC21 TC21titanium
titanium alloy. TheThe
alloy. fitted cyclic
fitted stress–strain
cyclic curves
stress–strain curves
were compared
comparedtotothe theexperimental
experimental results
results from
from uniaxial
uniaxial tension, as shown
tension, in Figure
as shown 5. It 5. It
in Figure
revealed that
revealed thatthe thecyclic
cyclicstress was
stress was less than
less thanthethe
unidirectional
unidirectional tensile stress
tensile under
stress the same
under the same
plastic strain.
plastic strain.Thus,Thus,thethecyclic softening
cyclic softening characteristic
characteristic of TC21of TC21alloy is demonstrated.
alloy is demonstrated.

Figure Comparisonofofuniaxial
5. Comparison
Figure 5. uniaxial tension
tension and
and cyclic
cyclic stress–strain
stress–strain curves.
curves.
Crystals 2022, 12, x FOR PEER REVIEW 5 of 11
3.2. HystersisLoop
3.2. Hystersis LoopAnalysis
Analysis
Figure66displays
Figure displaysthethe area
area of the
of the hysteresis
hysteresis loop loop asaswell
as well as the consumed
the consumed energy
energy in‐
increase with the increasing total strain. It can be inferred that the fatigue damage of
crease with the increasing total strain. It can be inferred that the fatigue damage of plasticplastic
deformation increased with the total strain. It is obvious that the hysteresis loops with
deformation increased with the total strain. It is obvious that the hysteresis loops with total
total strains of 1.4% and 1.6% are relatively small in comparison with those of 1.6% and
strains of 1.4% and 1.6% are relatively small in comparison with those of 1.6% and 2.0%,
2.0%, suggesting that fatigue damage caused by plastic deformation is much less at total
suggesting that fatigue damage caused by plastic deformation is much less at total strains
strains of 1.4% and 1.6%.
of 1.4% and 1.6%.

Thesuperimposed
Figure6.6.The
Figure superimposedhysteresis
hysteresisloops
loopsofofthe
thespecimens
specimensatatdifferent
differentstrain
strainlevels
levelsfor
forTC21
TC21 alloy.
alloy.

As for the half‐life hysteresis loops for the different total strain, the bottoms of the
hysteresis loops were moved to the same lowest points to analyze the Masing character‐
istics, which presented the same proportional limit. The J‐integral can be used to deal with
Figure 6. The superimposed hysteresis loops of the specimens at different strain levels for TC21
Crystals 2022, 12, 796 alloy. 5 of 11

As for the half‐life hysteresis loops for the different total strain, the bottoms of the
hysteresis
As forloops were moved
the half-life to the
hysteresis same
loops for lowest pointstotal
the different to analyze thebottoms
strain, the Masingofcharacter‐
the
hysteresis
istics, loops
which were moved
presented to the proportional
the same same lowest points
limit.toThe
analyze the Masing
J‐integral can be characteristics,
used to deal with
which
the LCFpresented
behaviorthewhensame proportional
metal displayslimit.
MasingThecharacteristics
J-integral can be used
[21]. to deal
Figure with the that
7 illustrates
the upper half of the hysteresis loops are non‐overlapping under the differentthat
LCF behavior when metal displays Masing characteristics [21]. Figure 7 illustrates thestrain;
total
upper half of the hysteresis loops are non-overlapping under the different total strain; thus,
thus, the alloy exhibits a non‐Masing behavior.
the alloy exhibits a non-Masing behavior.

7. The
Figure 7.
Figure Thesuperimposed
superimposed hysteresis loops along
hysteresis loopsthe linearthe
along portion to match
linear upper
portion to branch
matchfor TC21 alloy.
upper branch for
TC21 alloy.
3.3. Strain–Life Relationship
Based on Basquin and Coffin–Manson formulations, the relationship between fatigue
and the total strain can be expressed as [22]:
0
c
σf
0
∆ε t /2 = ε f (2N f ) + (2N f )b (2)
E
where σf 0 is the fatigue strength coefficient, b is the fatigue strength exponent, εf 0 is the
fatigue ductility coefficient, and c is the fatigue ductility exponent. The strain fatigue
parameters, evaluated by linear regression using the least square method, are listed in
Table 1.

Table 1. Strain fatigue parameters of TC21 titanium alloy.


0 0
σf b εf c

1634.7 −0.0821 12.428 −1.108

Figure 8 shows that the Coffin–Manson linear curves are well-fitted; however, Gao [11]
reported that the LCF of titanium alloy illustrated a bilinear characteristic due to the differ-
ent fatigue damages between plastic deformation and elatic deformation. It is indicated
that the fatigue damage of TC21 alloy is mainly controlled by elastic strain. Figure 8
also reveals that the transition life from elastic deformation to plastic deformation is only
350 reversals, which is in agreement with that of Ti1023 titanium alloy with 140 cycles [23].
The low transition life can be attributed to the low elastic modulus and high yield ratio of
titanium alloy. The fatigue life dominated by plastic deformation was only at 102 orders
of magnitude; thus, it was inferred that LCF of TC21 alloy was actually controlled by an
elastic deformation in a range over 103 cycles.
cated that the fatigue damage of TC21 alloy is mainly controlled by elastic strain. Figure
8 also reveals that the transition life from elastic deformation to plastic deformation is only
350 reversals, which is in agreement with that of Ti1023 titanium alloy with 140 cycles
[23]. The low transition life can be attributed to the low elastic modulus and high yield
Crystals 2022, 12, 796 ratio of titanium alloy. The fatigue life dominated by plastic deformation was only at 1026 of 11
orders of magnitude; thus, it was inferred that LCF of TC21 alloy was actually controlled
by an elastic deformation in a range over 103 cycles.

Figure 8. Number
Figure Numberof
ofhalf‐cycle
half-cyclereversals vs.vs.
reversals strain amplitude
strain in log–log
amplitude coordinates.
in log–log coordinates.

3.4.
3.4. TEM Observation
ObservationofofFatigue
FatigueMicrostructure
Microstructure
As for
for TC21
TC21 titanium
titaniumalloyalloywith
witha abi‐lamellar
bi-lamellar basketweave
basketweave microstructure,
microstructure, the βthe β
phase
phase containing
containing thethesecondary
secondaryαα phase
s phase
s obtained
obtained a a
highhigh hardness,
hardness, and and
the the
soft soft
α L αL phase
phase
preferentially deformed
preferentially deformedininthe thefatigue
fatigue process.
process.Thus,
Thus,thethe
fatigue microstructure
fatigue microstructurein theinαthe α
phase was investigated.
phase investigated.Figure
Figure9 9shows
shows the deformation
the deformation microstructure
microstructure in the
in α L phase
the αL phase
after low
after low cycle fatigue. There are a few free dislocations at the initial state
fatigue. There are a few free dislocations at the initial state (Figure (Figure 9a).9a).
At At a
a total strain of 1.4% and 1.6%, the fatigue dislocations in the α phase exhibit
total strain of 1.4% and 1.6%, the fatigue dislocations in the αL phase exhibit a plane slip
L a plane slip
(Figure 9b,c)
(Figure 9b,c)and
andpile-up
pile‐upatatthe
theαLα/β
L/β interface
interface(Figure
(Figure9d).
9d).The
Thedislocation
dislocationdensity
densityincreases
in‐
with the total strain (Figure 9e), but the alloy has a relatively low dislocation density in the
α phase. It is worth noting that the β phase can be sheared at a high total strain (Figure 9f),
and accordingly the alloy exhibited the cyclic softening characteristic. A similar result was
also reported for Ti-6242S titanium alloy with a Widmanstatten microstructure [14].

3.5. Cyclic Back Stress-Friction Stress Analysis


The cyclic softening and non-Masing behavior can be interpreted by the evolution of
back stress and friction stress, which can be calculated from each hysteresis loop according
to the reference [11]. As shown in Figure 10a, back stress can be divided into two types
for the different total strain. At a low total strain (εt < 2.0%), the back stress increased
slowly at a total strain of 1.8%, while the back stress increased rapidly at total strains of
1.4% and 1.6%. Furthermore, the increase of back stress in the total strain can be attributed
to the increase in dislocation density. However, the back stress firstly decreased and then
maintained a constant value at a total strain of 2.0%, which can be attributed to the shear of
the β phase (Figure 9f).
Friction stress had a decreasing trend with the number of cycles for each total strain.
Friction stress can also be divided into two types, as shown in Figure 10b. Friction stress
decreased continuously at total strains of 1.8% and 2.0%, while friction stress decreased
at two different rates under total strains of 1.4% and 1.6%. The decrease in friction was
related to the deformation dislocation. Multiple slip systems of dislocations were activated
at a high total strain and promoted the mobility of dislocations [11], which can reduce the
frictional internal stress of dislocations.
Cyclic softening behavior can be explained with the evolution of back stress and
friction stress. The cyclic softening for each strain amplitude (Figure 4) was attributed to
the competition effects between the increase in back stress and the decrease in friction stress.
As the decrease in friction stress was larger than the increase in back stress for total strains
of 1.4%, 1.6% and 1.8%, respectively, the alloy displayed the cyclic softening characteristic.
The cyclic softening at a total strain of 2.0% resulted from the superposition of the decreases
in both the back stress and friction stress.
Crystals 2022, 12, 796 7 of 11

Crystals 2022, 12, x FOR PEER REVIEW The Masing behavior was the result of the cyclic deformation microstructure 7 of 11 stabil-
ity [21]. As for the total strains of 1.4% and 1.6%, a low density of dislocation was observed
in the αL phase, while the β phase was sheared at a total strain of 2.0% (Figure 9f). The
alloy exhibited a different deformation microstructure at a different total strain. Meanwhile,
creases with the total strain (Figure 9e), but the alloy has a relatively low dislocation den‐
there
sity wasα no
in the direct
phase. It isrelationship
worth notingbetween
that the βthe cyclic
phase can back stress,atfriction
be sheared stress
a high total and cyclic
strain
(Figure 9f), and accordingly the alloy exhibited the cyclic softening characteristic. A simi‐ cyclic
strain (Figure 10), suggesting the instability of the fatigue microstructure under
deformation.
lar Therefore,
result was also reportedTC21 titanium
for Ti‐6242S alloy with
titanium alloyawith
basketweave microstructure
a Widmanstatten microstruc‐ displayed
a non-Masing
ture [14]. behavior.

(a) (b)

(c) (d)

(e) (f)
Figure 9. The microstructure of the α phase after low cycle fatigue deformation under different total
Figure 9. The microstructure of the α phase after low cycle fatigue deformation under different total
strain amplitudes: (a) initial state; (b) Δεt = 1.4%, Nf = 7988 cycles; (c) Δεt = 1.6%, Nf = 3200 cycles; (d)
strain
the amplitudes:
dislocation (a)atinitial
pile‐up state;N(b)
Δεt = 1.6%, ∆εt =cycles;
f = 3200
1.4%, N(e)f = 7988 cycles;
multiple (c) ∆εtat= Δε
dislocations 1.6%, Nf =N3200
t = 1.8%, f =
cycles;
(d) the dislocation pile-up at ∆ε
1731 cycles; (f) the shear of β phase at = 1.6%, N = 3200 cycles;
t Δεt = 2.0%,f Nf = 1302 cycles. (e) multiple dislocations at ∆ε t = 1.8%,
Nf = 1731 cycles; (f) the shear of β phase at ∆εt = 2.0%, Nf = 1302 cycles.
back stress and friction stress, which can be calculated from each hysteresis loop according
to the reference [11]. As shown in Figure 10a, back stress can be divided into two types for
the different total strain. At a low total strain (εt < 2.0%), the back stress increased slowly
at a total strain of 1.8%, while the back stress increased rapidly at total strains of 1.4% and
1.6%. Furthermore, the increase of back stress in the total strain can be attributed to the
Crystals 2022, 12, 796 increase in dislocation density. However, the back stress firstly decreased and then main‐ 8 of 11
tained a constant value at a total strain of 2.0%, which can be attributed to the shear of the
β phase (Figure 9f).

(a) (b)
Figure 10. The evolutions of back stress and friction stress with n/N at different strain amplitudes:
Figure 10. The evolutions of back stress and friction stress with n/N at different strain amplitudes:
(a) back stress; (b) friction stress.
(a) back stress; (b) friction stress.

3.6. Friction
Fatigue stress had
Fracture and aCrack
decreasing
Analysistrend with the number of cycles for each total strain.
Crystals 2022, 12, x FOR PEER REVIEW 9 of 11
Friction stress can also be divided into two types, as shown in Figure 10b. Friction stress
The typical fracture morphologies for TC21 titanium alloy are
decreased continuously at total strains of 1.8% and 2.0%, while friction stress decreased at
shown in Figure 11. In
the strain range from 1.4% to 2.0%, the fatigue cracks originated
two different rates under total strains of 1.4% and 1.6%. The decrease in friction was re‐ from the sample surface.
The
The
latedfatigue
fatigue crack
to the crack wasinitiated
was
deformation initiated from
from
dislocation. thelinear
the
Multiplelinear source
source
slip systemson on
the the sample
sample
of dislocations surface
surface
were a at
atactivated
high a high
strain,
strain, where
at a highwhere there were many
total strain and manypromotededges
edgesthewith
with a large
a large
mobility height
ofheight difference
difference
dislocations (Figure
[11], (Figure
which can 11a,b).
11a,b).
reduce It
It can can be
the
observed
be observed
frictional in Figure
in Figure
internal 11d that
11dofthat
stress many radial edges presented on the fracture
many radial edges presented on the fracture surface, which
dislocations. surface, which
was
was thethe typical
typical
Cyclic fatigue
fatigue
softening fracture
fracture
behavior morphology.
morphology.
can be explained TwoTwo
with similar
similar
the fractures
fractures
evolution morphologies
ofmorphologies
back stress andwere fric‐were
observed
observed
tion stress. in The
Figure
Figure 11c.
11c.
cyclic Thus,the
Thus,
softening the fatigue
fatigue
for crack
each crack
strain waswas initiated
initiated
amplitude from
(Figurefrom a multi-point
a4)multi‐point
was source
attributed source
to onthe on
the
the specimen effects
specimen
competition surface atataalow
surfacebetween strain,
lowthestrain,where inthe
where
increase fatigue
the
back stresscrack
fatigue initiation
crack
and the initiation
decreasewasindetermined
was determined
friction by by
stress.
micro‐plastic
micro-plastic
As the decrease damage
damage under
under
in friction macroscopic
macroscopic
stress was largerelastic
than deformation.
elasticthedeformation.
increase in back stress for total strains
of 1.4%, 1.6% and 1.8%, respectively, the alloy displayed the cyclic softening characteristic.
The cyclic softening at a total strain of 2.0% resulted from the superposition of the de‐
creases in both the back stress and friction stress.
The Masing behavior was the result of the cyclic deformation microstructure stability
[21]. As for the total strains of 1.4% and 1.6%, a low density of dislocation was observed
in the αL phase, while the β phase was sheared at a total strain of 2.0% (Figure 9f). The
alloy exhibited a different deformation microstructure at a different total strain. Mean‐
while, there was no direct relationship between the cyclic back stress, friction stress and
cyclic strain (Figure 10), suggesting the instability of the fatigue microstructure under cy‐
clic deformation. Therefore, TC21 titanium alloy with a basketweave microstructure dis‐
played a non‐Masing behavior.

3.6. Fatigue Fracture and Crack Analysis


The typical fracture morphologies for TC21 titanium alloy are shown in Figure 11. In
the strain range from 1.4% to 2.0%, the fatigue cracks originated from the sample surface.

Figure 11. Typical morphology of low cycle fatigue fracture: (a,b) ∆εt = 2.0%, Nf = 1302 cycles;
Figure 11. Typical morphology of low cycle fatigue fracture: (a,b) Δεt = 2.0%, Nf = 1302 cycles; (c,d)
Δεt =∆ε
(c,d) t = 1.4%,
1.4%, Nf =cycles.
Nf = 7998 7998 cycles.

The characteristics of a side crack were observed for LCF at a strain amplitude of
2.0%. Z‐shaped steps were presented on the side of the fatigue crack source, and some Z‐
shaped small cracks were also observed near the crack initiation site (Figure 12). The rela‐
Crystals 2022, 12, 796 9 of 11

The characteristics of a side crack were observed for LCF at a strain amplitude of 2.0%.
Z-shaped steps were presented on the side of the fatigue crack source, and some Z-shaped
small cracks were also observed near the crack initiation site (Figure 12). The relationship
between the fatigue sub-crack and microstructure is shown in Figure 13. Slip bands at
45◦ to the load direction can be observed on the αL phase and promoted the initiation of
fatigue sub-cracks. However, the αL /β interface cannot prevent the β phase from being
shear at the total strain of 2.0% (Figure 9f). The connection of these sub-cracks led to the
characteristics of Z-shaped steps. Actually, the propagation of sub-cracks was controlled by
mechanics rather than the microstructure, and cracks can pass through the three sequential
αL phases as a straight line. At a low total strain, slip bands were not obviously 10
Crystals 2022, 12, x FOR PEER REVIEW observed
of 11
Crystals 2022, 12, x FOR PEER REVIEWon the αL phase, and the dislocations were piled up at the αL /β interface 10(Figure
of 11 9d),

resulting in the initiation of a crack at the αL /β interface.

(a) (b)
(a) (b)
Figure 12. The morphology of the side crack at the crack initiation site at Δεt = 2.0%, Nf = 1302 cy‐
Figure
cles.12.
Figure 12.The
Themorphology
morphologyofof
the side
the crack
side at the
crack crack
at the initiation
crack sitesite
initiation att ∆ε
at Δε = 2.0%, Nf = 1302 cy‐
t = 2.0%, Nf = 1302 cycles.
cles.

(a) (b)
(a) (b)
Figure 13. The morphology of the fatigue sub‐crack and microstructure (Backscatter electron mi‐
Figure13.
Figure Themorphology
13.The morphology of the fatigue sub-crack and microstructure (Backscatter electron mi-
croscopy): (a) Δεt = 2.0%, of
Nf the fatigue
= 1302 sub‐crack
cycles; (b) Δεt =and
1.4%,microstructure (Backscatter
Nf = 7988 cycles. electron mi‐
croscopy): (a)Δε∆ε
croscopy):(a) t =
t 2.0%, Nf N
= 2.0%, = 1302
= f1302 cycles;
cycles; (b)t =∆ε
(b) Δε t = 1.4%,
1.4%, Nf =cycles.
Nf = 7988 7988 cycles.
4. Conclusions
4.Conclusions
Conclusions
4.
1. 1.TC21
1.
TC21titanium
TC21 titaniumalloy
titanium
alloydisplayed
alloy
displayed cyclic
displayed
cyclic softening
cyclic softening
softening and
and
and non‐Masing
non-Masing
non‐Masing
behavior
behavior
behavior
that
that
were
that
were were
interpreted on the basis of the cyclic back stress, friction stress and fatigue defor‐
interpretedon
interpreted onthethebasis
basisofof the
the cyclic
cyclic back
back stress,
stress, friction
friction stress
stress and
and fatigue
fatigue deformation
defor‐
mation microstructure.
microstructure.
2.mation
Lowmicrostructure.
cycle fatigue cracks were predominantly initiated from the slip bands on the
2. Low
2. Lowcyclecyclefatigue
fatigue cracks
cracks werewere predominantly
predominantly initiated
fromfrom the bands
slip bands on the
surface of the samples at a high total strain, initiated
and the crack the slip
initiation on the
occurred at the
surfaceofof
surface the
the samples
samples at athigh
a a high total
total strain,
strain, and and
the the crack
crack initiation
initiation occurred
occurred at the at the
αL/β interface at a relatively low total strain.
ααL/β
L /β interface
interface at aat a relatively
relatively low low
totaltotal strain.
strain.
Author Contributions: Z.Z. and D.C. conceived and designed the research; Y.S., X.H. and H.Q. per‐
Author Contributions:
formed Z.Z.B.N.
the experiment; and analyzed
D.C. conceived and designed
and wrote the research;
the manuscript. Y.S., X.H.
All authors haveand
readH.Q.
andper‐
agreed
formed the experiment; B.N. analyzed and wrote
to the published version of the manuscript. the manuscript. All authors have read and agreed
to the published version of the manuscript.
Funding: This research was funded by the R & D plan for key areas in Guangdong Province
Funding: This researchSpecial
(2020B010186001), was funded by the RProjects
Innovation & D plan of for key areas ininGuangdong
Universities Guangdong Province
Province
Crystals 2022, 12, 796 10 of 11

Author Contributions: Z.Z. and D.C. conceived and designed the research; Y.S., X.H. and H.Q.
performed the experiment; B.N. analyzed and wrote the manuscript. All authors have read and
agreed to the published version of the manuscript.
Funding: This research was funded by the R & D plan for key areas in Guangdong Province
(2020B010186001), Special Innovation Projects of Universities in Guangdong Province (2018KTSCX240),
Core technology research project of Foshan (1920001000412), Basic and applied basic research fund
project in Guangdong Province (2020b15120093), Science and technology project in Guangdong
(2020B121202002), and Innovation driven project of science and technology plan in Jiangxi Yichun.
Data Availability Statement: Not applicable.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Wen, X.; Wan, M.; Huang, C.; Lei, M. Strength and fracture toughness of TC21 alloy with multi-level lamellar microstructure.
Mater. Sci. Eng. A 2019, 740, 121–129. [CrossRef]
2. Ye, X.; Wan, M.; Huang, C.; Lei, M.; Jian, S.; Zhang, Y.; Xu, D.; Huang, F. Effect of aging temperature on mechanical properties of
TC21 alloy with multi-level lamellar microstructure. Mater. Sci. Eng. A 2022, 840, 142825. [CrossRef]
3. Singh, N.; Singh, V. Low cycle fatigue behavior of Ti alloy IMI 834 at room temperature. Mater. Sci. Eng. A 2002, 325, 324–332.
[CrossRef]
4. Singh, N.; Singh, V. Low cycle fatigue behavior of Ti alloy Timetal 834 at 873 K. Int. J. Fatigue 2007, 29, 843–851. [CrossRef]
5. Huang, J.; Wang, Z.; Xue, K.M. Cyclic deformation response and micromechanisms of Ti alloy Ti-5Al-5V-5Mo-3Cr-0.5Fe. Mater.
Sci. Eng. A 2011, 528, 8723–8732. [CrossRef]
6. Huang, J.; Wang, Z.R.; Zhou, J. Cyclic deformation response of b-Annealed Ti-5Al-5V-5M-3Cr alloy under compressive loading
conditions. Metall. Mater. Trans. A 2011, 42, 2868–2880. [CrossRef]
7. Wu, G.Q.; Shi, C.L.; Sha, W. Effect of microstructure on the fatigue properties of T-6Al-4V titanium alloys. Mater. Des. 2013, 46,
668–674. [CrossRef]
8. Li, S.; Xiong, B.; Hui, S. Comparison of the fatigue and fracture of Ti-6Al-2Zr-1Mo-1V with lamellar and bimodal microstructures.
Mater. Sci. Eng. A 2007, 460, 140–145. [CrossRef]
9. Lei, Z.; Gao, P.; Li, H.; Cai, Y.; Li, Y.; Zhan, M. Comparative analyses of the tensile and damage tolerance properties of tri-modal
microstructure to widmanstätten and bimodal microstructures of TA15 titanium alloy. J. Alloy. Compd. 2019, 788, 831–841.
[CrossRef]
10. Xu, Z.; Huang, C.; Wan, M.; Tan, C.; Zhao, Y.; Ji, S.; Zeng, W. Influence of microstructure on strain controlled low cycle fatigue
crack initiation and propagation of Ti-55531 alloy. Int. J. Fatigue 2022, 156, 106678. [CrossRef]
11. Gao, P.; Lei, Z.; Li, Y.; Zhan, M. Low-cycle fatigue behavior and property of TA15 titanium alloy with tri-modal microstructure.
Mater. Sci. Eng. A 2018, 736, 1–11. [CrossRef]
12. Helstroffer, A.; Hémery, S.; Andrieu, S.; Villechaise, P. Low cycle fatigue crack initiation in Ti-5Al-5Mo-5V-3Cr in relation to local
crystallographic orientations. Mater. Lett. 2020, 276, 128198. [CrossRef]
13. Joseph, S.; Bantounas, I.; Lindley, T.C.; Dye, D. Slip transfer and deformation structures resulting from the low cycle fatigue of
near-alpha titanium alloy Ti-6242Si. Int. J. Plast. 2017, 100, 90–103. [CrossRef]
14. Anoushe, A.S.; Zarei-Hanzaki, A.; Abedi, H.R.; Barabi, A.; Huang, C.; Berto, F. On the microstructure evolution during isothermal
low cycle fatigue of β-annealed Ti-6242S titanium alloy: Internal damage mechanism, substructure development and early
globularization. Int. J. Fatigue 2018, 116, 592–601. [CrossRef]
15. Xu, Z.; Huang, C.; Tan, C.; Wan, M.; Zhao, Y.; Ye, J.; Zeng, W. Influence of microstructure on cyclic deformation response and
micromechanics of Ti-55531 alloy. Mater. Sci. Eng. A 2020, 803, 140505. [CrossRef]
16. Sen, M.; Suman, S.; Mukherjee, S.; Banerjee, T.; Sivaprasad, S.; Tarafder, S.; Bhattacharjee, A.; Kar, S.K. Low cycle fatigue behavior
and deformation mechanism of different microstructures in Ti-5Al-5Mo-5V-3Cr alloy. Int. J. Fatigue 2021, 148, 106238. [CrossRef]
17. Du, D.; Liu, D.; Zhang, X.; Tang, J.; Meng, B. Effects of WC-17Co Coating Combined with Shot Peening Treatment on Fatigue
Behaviors of TC21 Titanium Alloy. Materials 2016, 9, 865. [CrossRef] [PubMed]
18. Yu, Z.L.; Zhao, Y.Q.; Zhou, L.; Sun, J. Strain and stress controlled low cycle fatigue behaviors of TC21 alloy. Rare. Metal. Mat. Eng.
2009, 38, 224–228.
19. Tan, C.; Li, X.; Sun, Q.; Xiao, L.; Zhao, Y.; Sun, J. Effect of α-phase morphology on low-cycle fatigue behavior of TC21 alloy. Int. J.
Fatigue 2015, 75, 1–9. [CrossRef]
20. Gao, C.; Ren, T.; Liu, M. Low-cycle fatigue characteristics of Cr18Mn18N0.6 austenitic steel under strain controlled condition at
100 C. Int. J. Fatigue 2019, 118, 35–43. [CrossRef]
21. Bai, F.; Zhou, H.; Liu, X.; Song, M.; Sun, Y.; Yi, H.; Huang, Z. Masing behavior and microstructural change ofquenched and
tempered high-strength steel under low cycle fatigue. Acta Metall. Sin. 2019, 32, 1346–1354. [CrossRef]
Crystals 2022, 12, 796 11 of 11

22. Szachogluchowicz, I.; Sniezek, L.; Hutsaylyuk, V. Low cycle fatigue properties of AA2519–Ti6Al4V laminate bonded by explosion
welding. Eng. Fail. Anal. 2016, 69, 77–87. [CrossRef]
23. Huang, L.; Huang, X. Low cycle fatigue behavior of Ti-1023 titanium alloy at room temperature. Rare. Metal. Mat. Eng. 2006, 35,
703–706.

You might also like