You are on page 1of 25

Engineering Fracture Mechanics 235 (2020) 107126

Contents lists available at ScienceDirect

Engineering Fracture Mechanics


journal homepage: www.elsevier.com/locate/engfracmech

Review

A review on mixed mode fracture of metals


T
Yanlin Wang, Weigang Wang, Bohua Zhang, Chun-Qing Li

School of Engineering, RMIT University, Melbourne 3001, Australia

ARTICLE INFO ABSTRACT

Keywords: Mode I fracture mechanics has been relatively mature but mixed mode fracture mechanics
Mixed mode fracture continues to present challenges, in particular for ductile materials under multiaxial stresses. Since
Fracture parameters most engineering materials, e.g., metals, exhibit elasto-plastic or ductile behaviour and are
Fracture criteria subjected to inclined cracks and/or multiaxial loading, there is a well justified need to establish a
Fracture tests
solid understanding of their fracture behaviour under mixed modes. The intention of this paper is
Elasto-plastic metals
to critically review mixed mode fracture criteria for predicting the crack propagation in metals.
Experimental methods for determining the fracture toughness under different mode mixities are
reviewed and discussed. More importantly, research gaps in current studies are identified and
further research needs are proposed. Through this review, it is found that the mixed mode
fracture criteria need to be based on elasto-plastic fracture mechanics to account for large plastic
deformation, and that for ductile metals, the addition of Mode II or Mode III loading can lead to a
remarkable decrease of total fracture toughness of the metal, whilst the total fracture toughness
of less ductile metals is insensitive to the addition of Mode II or Mode III loading. This review can
contribute to the development of advanced mixed mode fracture criteria and testing methods for
ductile metals.

1. Introduction

Cracks have been well recognised as a major cause for failures of engineering structures made of metals [1]. For example, the
Liberty ship accident was induced by the presence of welding cracks [2] and the Comet I aeroplane accident was caused by fatigue
cracks [3,4]. Also, many failures of underground pipelines were found to be initiated by corrosion-induced pipe wall cracks [5,6].
Cracks in these important engineering structures can not only impose adverse impacts on structural integrity but also cause severe
economic loss and subsequent environmental and social hazards, such as a burst of underground pipes. Therefore, it is imperative to
thoroughly understand the failure mechanisms of cracked metals and develop advanced fracture mechanics-based failure criteria for
structural integrity assessment.
Depending on the type of applied loading, there are three different modes of fracture, namely Mode I (opening mode), Mode II (in-
plane shear mode) and Mode III (out-of-plane shear mode) [2]. The fracture toughness of Mode I has been employed to quantify the
material capacity to resist crack extension [7]. A simple fracture criterion states that the fracture initiates when the fracture driving
force, e.g., stress intensity factor K, energy release rate G, J-integral, reaches the Mode I fracture toughness [8]. This simple criterion
has been widely used in current engineering failure assessment for cracked bodies, e.g., structural materials and components [9].
However, in practical situations, structural materials and components are prone to mixed mode fracture due to the inclined cracks or
multiaxial loading [5]. Unlike Mode I fracture, cracks under mixed mode loading may extend in neither pure opening mode nor shear


Corresponding author.
E-mail address: chunqing.li@rmit.edu.au (C.-Q. Li).

https://doi.org/10.1016/j.engfracmech.2020.107126
Received 10 April 2020; Received in revised form 24 May 2020; Accepted 25 May 2020
Available online 03 June 2020
0013-7944/ © 2020 Elsevier Ltd. All rights reserved.
Y. Wang, et al. Engineering Fracture Mechanics 235 (2020) 107126

mode [10]. The stress field at the crack tip is much more complicated due to the interaction of loading [11]. Furthermore, numerous
experimental observations (e.g., [12–14]) have indicated that the fracture toughness of some metals (e.g., steel, aluminium alloy)
decreases significantly with an additional Mode II or Mode III loading. Therefore, the single mode fracture criteria using Mode I
fracture toughness are hardly appropriate for mixed mode fracture assessment for structures or bodies subjected to inclined cracks or
multiaxial loadings.
Considerable research has been undertaken to develop fracture criteria for predicting crack initiation and propagation under
mixed mode loading. For instance, Erdogan and Sih [15] theoretically predicated that a crack evolves perpendicularly to the max-
imum tangential stress at the crack tip in brittle materials under mixed mode loading. Sih [16,17] proposed the minimum strain
energy density (MSED) criterion for predicting the direction of crack growth under mixed mode loading. Richard [18] developed an
empirical fracture criterion, which built a relationship between all three modes of fracture and Mode I fracture toughness to simplify
the prediction of crack growth under multiaxial loading. This criterion was later modified by Richard et al. [19] to fit the testing
results of different brittle materials. Chang et al. [20] expended the maximum energy release rate (MERR) criterion, which was firstly
proposed by Hussain et al. [21], to a general brittle fracture criterion for cracked bodies under three-dimensional loading. They also
conducted experiments to verify the general brittle fracture criterion.
However, the above-mentioned works were based on linear elastic fracture mechanics (LEFM), the corresponding fracture
parameters of which are no longer able to characterise the ductile fracture behaviour accurately due to the large yielding at the crack
tip [22]. As a result, the LEFM-based criteria become inapplicable for ductile materials. With the need for a nonlinear fracture
mechanics model, Rice [23] firstly proposed the J-integral as a nonlinear elastic fracture parameter. Hutchinson [24] and Rice and
Rosengren [25] independently showed that the J-integral could characterise the stress and strain field at the crack tip of nonlinear
elastic materials. Wells [26] proposed the critical crack tip opening displacement (CTOD) as a measurement of fracture toughness.
O'dowd and Shih [27,28] performed a series of finite element analysis (FEA) on different body geometries based on the theory of
plasticity. They proposed a J-Q theory, where Q characterises the level of crack tip constraint effect, to describe the elasto-plastic
crack-tip field. All these works together contributed to the development of the elasto-plastic fracture mechanics (EPFM).
With the EPFM developed, mixed mode fracture criteria for elasto-plastic or ductile materials have gradually been established. For
example, Ma et al. [29] proposed a CTOD-based criterion to predict the crack extension and fracture failure of both brittle and ductile
or elasto-plastic materials under mixed mode loading. Li et al. [30] established a J - M P -based criterion, where M P represents the
plastic mode mixity, for determining the transition regime between the opening crack manner and the shear crack manner of elasto-
plastic materials under mixed mode I-II loading. Furthermore, an increasing number of experimental studies on mixed mode fracture
of elasto-plastic materials have been carried out. Manoharan et al. [31] experimentally studied the mixed mode I-III fracture be-
haviour of steel. Their study found that the addition of Mode III had a minor effect on the ability of fracture resistance of spherodized
1090 steel. Keiichiro and Hitoshi [32] developed a simple testing procedure and calculated the total J-integral of an aluminium alloy
under mixed mode I-II loading. They found that the crack initiation under mixed mode I-II loading could be divided into tensile type
and shear type as shown in Fig. 1.
Densley and Hirth [33] carried out a mixed mode I-III fracture test on HSLA-100 steel and observed that the total fracture
toughness of HSLA-100 steel initially decreased with the imposition of mode III shear loading but eventually increased again. The
increase of fracture toughness at later testing stage was attributed to the friction between the fracture surfaces under Mode III
dominance (high ratio of Mode III to Mode I loading). Roy et al. [34] experimentally compared the fracture toughness of an alu-
minium alloy with two different crack length to specimen width ratios under mixed mode I-II loading. Their results suggested that the
crack length shows a marginal effect on the fracture toughness of 2014-O aluminium alloy under Mode II dominance (high ratio of
Mode II to Mode I loading). Qian and Yang [35] experimentally investigated the effect of mode mixities, i.e., the ratios of Mode I to
Mode II loading, on the total fracture toughness of the 5083-H112 aluminium alloy. According to their experimental results, the

Fig. 1. Crack initiation types under mixed mode I-II loading: (a) tensile type (Mode I dominance), (b) shear type (Mode II dominance) (modified
from Keiichiro and Hitoshi [32]).

2
Y. Wang, et al. Engineering Fracture Mechanics 235 (2020) 107126

specimens under Mode I dominance (high ratio of Mode I to Mode II loading) exhibit a decrease in total fracture toughness when the
mode mixity increases. On the contrary, the total fracture toughness of specimens under Mode II dominance increases with the
increase of mode mixity. These studies, together with some others (e.g., [36–38]), contributed to the body of knowledge of mixed
mode fracture behaviour of elasto-plastic or ductile materials.
A thorough review of published literatures identifies some important issues to be further investigated which are: (1) accurate
understanding of the fracture behaviour under mixed mode I-II-III loading for elasto-plastic or ductile materials; (2) experimental
validation of fracture criteria based on elasto-plastic fracture mechanics (EPFM) for metals under plane strain condition; and (3)
effective testing procedures to determine the mixed mode fracture toughness. Although some review papers have been published to
summarise the current state of understanding on mixed mode fracture behaviour and proposed the future work, most of them focus
on brittle materials with fatigue crack growth. Rozumek and Macha [39] and Qian and Fatemi [40] respectively reviewed the fracture
mechanics parameters, mixed mode fracture criteria and some experimental findings for brittle materials. Pook [41], Rozumek and
Macha [39] and Qian and Fatemi [40] also respectively reviewed the fracture behaviour of brittle materials under mixed mode static
or monotonic loading. For brittle materials containing cracks, fracture occurs rapidly with no or small plastic yielding at the crack tip.
For fatigue fracture in ductile materials, cracks usually grow slowly and steadily at first, and then a rapid fracture occurs when a crack
reaches a critical length. This essential difference necessitates a review on mixed mode fracture of elasto-plastic or ductile materials.
Although in the reviews of Pook [41], Rozumek and Macha [39] and Qian and Faremi [40], the failure criteria and characteristics
for mixed mode fatigue crack growth in ductile materials under mixed mode cyclic loading were discussed, current research on mixed
mode fracture behaviour of ductile materials, where the cracks grow stably with considerable plastic deformation, has not been
critically reviewed so far. Since most engineering materials, e.g., metals, exhibit elasto-plastic or ductile behaviour to certain extent,
it is well justified to have a solid understanding of their mixed mode fracture behaviour when they are subjected to inclined cracks
and/or multiaxial loading with large plastic deformation at the crack tips.
This paper intends to present a state-of-the-art review on mixed mode fracture criteria and experiments for metals. The objectives
of this review are: (1) to investigate the fracture characteristics of elasto-plastic or ductile metals under mixed mode loading; (2) to
evaluate the engineering application of mixed mode fracture criteria to metals and its scope; (3) to provide a reference guide for
mixed mode fracture tests on metals and determination of total fracture toughness; and (4) to explore the critical issues for future
research on mixed mode fracture behaviour of metals. The review presented in this paper enables a better understanding of mixed
mode fracture mechanics of elasto-plastic materials and can be beneficial to both researchers and engineers with a view to the
development of a standardised testing method to determine the mixed mode fracture toughness of elasto-plastic materials.

2. Basic parameters for mixed mode fracture

Stress intensity factor K , elastic strain energy release rate G , J-integral and crack tip opening displacement (CTOD) are the basic
parameters used in fracture mechanics. The crack-tip stress field is the basis for establishing fracture criteria for the cracked bodies
[39]. Fig. 2 shows a crack front in the cylindrical coordinate system with the origin at point O. The stress field near the crack tip in a
linear elastic material can be described by the near-field solutions in the cylindrical coordinate system as follows [42]:

KI 5 1 3 KII 5 3 3
r = cos cos sin sin
2 r 4 2 4 2 2 r 4 2 4 2 (1)

KI 3 1 3 KII 3 3 3
= cos + cos sin + sin
2 r 4 2 4 2 2 r 4 2 4 2 (2)

KI 1 1 3 KII 1 3 3
r = sin + sin + cos + cos
2 r 4 2 4 2 2 r 4 2 4 2 (3)

KIII
rz = sin
2 r 2 (4)

Fig. 2. Cylindrical coordinate system ad stress components at a crack front (modified from Richard et al. [42]).

3
Y. Wang, et al. Engineering Fracture Mechanics 235 (2020) 107126

KIII
z = cos
2 r 2 (5)

z = ( r + )= 2KI cos 2KII sin for plane strain


2 r 2 2 (6)

z = 0 for plane stress (7)

where r is the distance from the calculated point to the crack tip, is the angle defined in Fig. 2, is the Poisson’s ratio, the
parameters KI ,KII and KIII are the stress intensity factors (SIFs) for Mode I, Mode II and Mode III fracture, respectively. Equations (1)
to (7) are the first term, namely the singular K field, derived from Williams series expansion [43]. The higher order terms are often
ignored as the SIFs can generally represent the magnitude of the stress field surrounding the crack tip in the linear elastic materials. In
general, the SIFs can be expressed in the form as follows [2]:
KI , II , III = a (8)

where is the characteristic stress, a is the characteristic crack dimension, is a dimensionless equation that depends on the
geometry and mode of loading. With the SIFs, an elastic mode mixity parameter Me has been developed to quantify the mixity ratio.
This parameter is defined by Shih [44] as follows:

2 KI
Me = arctan
KS (9)

where Ks = KII for mixed mode I-II and Ks = KIII for mixed mode I-III.
For elasto-plastic materials, Rice [23] proposed a nonlinear-elastic fracture parameter called J-integral for characterising the
crack-tip stress and strain field. Physically, J-integral represents the rate of change of the total potential energy in elasto-plastic
materials per unit fracture surface area. The J-integral is a path independent integral around the crack tip and defined as follows:
ui
J= wdy Ti ds
x (10)

where w is the strain energy density, Ti is the components of traction vector, ui is the components of the displacement vector, ds, is the
length of increment of the contour, is an arbitrary curve around the crack, x and y are the Cartesian coordinates. The coordinate
system and integration path are shown in Fig. 3.
For materials with the stress and plastic strain being described by a power-law relationship, the crack-tip stress and strain field can
be characterised as follows [24,25]:
1 (n + 1)
EJ
ij = 0 2 ij (n , )
0 In r (11)

n (n + 1)
0 EJ
ij = 2 ij (n , )
E 0 In r (12)

where ij and ij are the stress and strain tensors, respectively, 0 is the reference stress usually equal to the yield stress, is a
dimensionless constant, E is Young’s modulus, n is the strain hardening exponent, In is integration constant depending on n , and r
are polar coordinates as shown in Fig. 3, ij and ij are the dimensionless functions of n and . Equations (11) and (12) together are
called HRR singularity named after Hutchinson, Rice, and Rosengren, and the J-integral represents the amplitude of HRR singularity
[45]. In linear elastic fracture mechanics (LEFM), J-integral can be interpreted as the elastic strain energy release rate G and be
calculated directly from the SIFs as follows [2]:

KI2 K2 K2
J=G= + II + III
E E 2µ (13)

Fig. 3. An arbitrary integration path Γ of J-integral.

4
Y. Wang, et al. Engineering Fracture Mechanics 235 (2020) 107126

Fig. 4. Mixed mode CTOD.

where E = E for plane stress and E = E (1 2 ) for plane strain, E is Young’s modulus, µ is the shear modulus.

The crack-tip-opening displacement (CTOD) is another parameter used to characterise the fracture behaviour of elasto-plastic
materials. As shown in Fig. 4, the mixed mode CTOD is defined as the relative displacement of two opposite points on the upper and
lower crack surfaces at a fixed distance behind the crack tip [46]. The mixed mode CTOD is the crack tip displacement vectors in
essence.
Pirondi and Donne [11] carried out a series of experiments on ferritic steel to study the CTOD characteristics under mixed mode I-
II loading. In their study, a J-CTOD relationship in HRR singularity was established as follows:

1 n J
CTOD = ( 0) DN
0 (14)
where 0 = 0 E , DN is a function of mode mixity. For materials with large values of strain hardening exponent, it can be seen that
( 0)1 n 1 and the dependence of CTOD on ( 0)1 n is negligible.
Of four most significant parameters, i.e., K , G , J-integral and CTOD, the LEFM based parameters K and G are established under
the assumption of isotropic linear elastic materials. They are only applicable to some metals (e.g., grey cast iron, high-carbon steel) of
which the yielding zone near the crack tip is very small compared to the crack size. For ductile metals, such as mild steel, the cracks
often grow in the form of ductile tearing with large yielding around the crack tip [47]. For such materials, the EPFM based parameters
J-integral and CTOD are more suitable for characterising the fracture behaviour since they take nonlinear deformation (e.g., plastic
deformation) into account. However, in case of ductile metals that turn to brittle, LEFM analysis can still be performed. For instance,
some steels become brittle below a specific range of temperature, i.e., ductile to brittle transition temperature, or some metals (e.g.,
steel, copper) lose ductility when exposed to hydrogen. Although there are no universal rules for considering the plasticity effect or
choosing whether LEFM or EPFM analysis [2], J-integral is recommended for both cases as it considers plastic yielding during crack
propagation and the SIFs in LEFM as shown in Equation (13).

3. Mixed mode fracture criteria

Cracks in structures or a body are usually not orthogonal to longitudinal or transverse direction, i.e., they are inclined. Also,
structural bodies are usually subjected to multiaxial stresses. A structure or body subjected to inclined cracks and/or under multiaxial
stresses would suffer mixed mode fracture. For a cracked body under a single mode loading, the fracture is considered to occur when
the corresponding fracture mechanics parameter reaches its critical value, i.e., fracture toughness. However, this fracture criterion
may be non-conservative for the mixed mode fracture conditions. The fracture can occur under lower fracture driving forces with the
superposition of Mode I, Mode II and Mode III [42]. For single mode loading conditions, a crack extends in the same plane under
Mode I loading, kinks in the same plane under Mode II loading, and twists out of plane under Mode III loading [20,48]. The
combinations of mixed mode loading can lead to an alteration of the crack orientation and extension path [42]. In order to predict the
angle of crack extension and the critical conditions that fracture occurs under mixed mode loading, several fracture criteria have been
theoretically established for 2D and 3D mixed mode loading conditions.

3.1. Mixed mode fracture criteria based on LEFM

The available fracture criteria based on linear elastic fracture mechanics (LEFM) can be mainly classified into energy-based and
stress-based criteria. Table 1 summarises the LEFM-based criteria for mixed mode fracture from the published literature. In the group
of energy-based criteria, the maximum energy release rate (MERR) criterion [21] predicts that the crack will initiate along the
direction where the energy release rate reaches the maximum. However, as it is difficult to obtain an analytical solution to the energy
release rate in an arbitrary direction for a cracked body, the application of MERR criterion is very limited. Chang et al. [20] proposed
an approximate formula for the energy release rate based on the analytical analysis of 3D crack-tip stress fields, which can be
expressed as follows:

G=
1
E
cos2
2 {1 2 [(1 + cos ) KI2 4 sin KI KII + (5 3 cos ) KII2 ] + KIII
2
} (15)
where is the crack kinking angle. With this approximation, Chang et al. [20] established a general fracture criterion based on the
MERR criterion. The crack propagation direction and fracture failure can be determined as follows:

5
Y. Wang, et al. Engineering Fracture Mechanics 235 (2020) 107126

Table 1
Mixed mode fracture criteria based on LEFM.
Reference Fracture mode Description

Hussain et al. [21], Mixed I-II-III Crack grows along the maximum energy release rate G at fracture angle f; crack extension initiates when the
Chang et al. maximum G attains a critical value G ,c .
[20]
Sih [16,17,52] Mixed I-II-III Crack grows along the minimum strain energy density S at fracture angle f ;crack extension initiates when the
minimum Smin attains a critical value Sc .
Koo and Choy Mixed I-II Crack grows along the maximum tangential strain energy density (TSED).
[113],
Ayatollahi and
Saboori [79]
Chang [114] Mixed I-II Crack grows along the maximum tangential strain .
Erdogan and Sih Mixed I-II Crack grows perpendicular to the maximum tangential stress .
[15]
Williams and Ewing Mixed I-II Crack grows along the maximum tangential stress with T-stress.
[115], Smith
et al. [110]
Papadopoulos [116] Mixed I-II Crack grows along the maximum third stress invariant Det ( ij ).
Kong et al. [61] Mixed I-II-III Crack grows along the maximum stress triaxiality M at fracture angle f ;crack extension initiates when the maximum
M attains a critical value Mc .
Matvienko [117] Mixed I-II Crack grows along the maximum average tangential stress (ATS).
Sajjadi et al. [118] Mixed I-II Crack grows along the maximum effective stress (ES).
Schöllmann et al. Mixed I-II-III Crack grows along maximum principal stress σ'1 at fracture angle f.
[48]
Pook [59,60] Mixed I- Crack grows along the kinking angle f: KI sin f = KII (3 cos f 1), ( 70. 5° f 70. 5°) Crack grows along the
IIandMixed I- 2KIII
twisting angle f: tan 2 = , ( 45° 45°)
III f KI (1 2 ) f

Richard et al.[19] Mixed I-II-III Crack initiates when:


KI 1
+ 4( 2 KIII )2 = KIC Crack kinking angle
2
+
2
KI2 + 4( 1 KII )2 f and twisting angle f are

( ) ( )
2 2
|KII | |KII | |KIII | |KIII |
predicted by: f = A +B , f = C +D where
KI + |KII | + |KIII | KI + |KII | + |KIII | KI + |KII | + |KIII | KI + |KII | + |KIII |

A , B , C , and D are the fitting angles.

G 2G
= 0, <0
2 (16)

+1 2
G = Gc = KIc
8µ (17)
where KIc is the Mode I fracture toughness (critical KI value), = 3 4 for plane strain and = (3 4 ) (1 + ) for plane stress.
Equation (16) represents the maximum energy release rate, which can be used for the prediction of the direction of crack propa-
gation. Equation (17) proposes the critical value that can be used for predicting the crack extension.
The minimum strain energy density (MSED) criterion [16,17] assumes that the crack will initiate along with a minimum strain
energy density factor (S) at a fracture angle f . The unstable crack will occur when this factor attains a critical value Sc . The strain
energy density S is defined as follows:
S = a11 KI2 + a12 KI KII + a22 KII2 (18)
where a11, a12 , and a22 are the dimensionless factors which are given as follows:
1
a11 = ( cos )(1 + cos )
16µ (19a)

1
a12 = (2 cos + 1) sin
16µ (19b)

1
a22 = [( + 1)(1 cos ) + (3 cos 1)(1 + cos )]
16µ (19c)
where = 3 4 for plane strain and = (3 ) (1 + ) for plane stress. The conditions for crack initiation can be expressed as
follows:
S 2S
= 0, 2
>0
= f = f (20)

S = Sc (21)
This criterion has been experimentally verified by a number of researchers (e.g., Hua et al. [49], Sih and Barthelemy [50],

6
Y. Wang, et al. Engineering Fracture Mechanics 235 (2020) 107126

Badaliance [51]). Sih [52] has extended the 2D MSED criterion to 3D fracture, from which strain energy density factor can be
expressed as follows:
1
S= 16 µ
{(1 + cos )( cos ) KI2 + 2 sin (2 cos + 1) KI KII
+ [( + 1)(1 cos ) + (1 + cos )(3 cos 1)] KII2 + 4KIII
2
} (22)
This new strain energy density factor has a simple mathematical expression. It is also able to handle various combined loadings.
However, this 3D MSED criterion was found to be inconsistent with some experimental results in Mageed and Pandey [53,54]. Also,
Wang et al. [55] pointed out that the expressions of 3D MSED and the general fracture criterion are independent on the twisting angle
(see Fig. 2), suggesting that these two criteria are not appropriate for the prediction of the crack propagation direction under Mode
III dominance.
In stress-based criteria, the maximum tangential stress (MTS) criterion [15] has become one of the most commonly used criteria
due to its simplicity and good agreement with the micromechanical models [56]. The MTS criterion assumes that the crack will
propagate in the direction perpendicular to the maximum tangential stress 0, max . The condition for the maximum tangential stress
can be mathematically expressed as follows:
2
0 0
=0, <0
0 0
2 (23)
where the tangential stress 0 near the crack tip. In 2D polar coordinates, it can be expressed as follows:
1 3
= cos 0 KI cos2 0 KII sin 0
0
2 r 2 2 2 (24)
where 0 is the polar angle in a 2D polar coordinate system. Fracture tests carried out by Swankie and Smith [57] and Maccagno and
Knott [58] showed that MTS criterion agreed with the test results of steel under mixed mode I-II loading at low temperatures
(−196 °C and − 120 °C). The fracture behaviour of the steel, under such low temperatures, was found to be of brittle type due to the
transgranular cleavage [58].
Pook [59,60] developed a criterion by solving the traction on a kinked crack. In this criterion, the crack is assumed to grow along
the kinking angle f under mixed mode I-II and twisting angle f under mixed mode I-III, which can be approximately calculated as
follows:
KI sin f = KII (3 cos f 1), ( 70. 5° f 70. 5°) (25)
and
2KIII
tan 2 = , ( 45° 45°)
(26)
f f
KI (1 2 )
Kong et al. [61] developed a model of distribution of triaxial stress M around the crack tip. By analytically analysing the 3D stress
fields near the crack tip in elastic bodies, M can be obtained as follows:
F1
M=
F2 (27)
with

F1 = 2 2 (1 + ) KI cos { 2
KII sin
2 } (28a)

F2 = 3 {( K 3
2
2
I
9 2
K
2 II ) sin2 + [2KI2 (1 2 )2 + 6KII2 ]cos2 2

+ 8KII2 (1 + 2)sin2
2
+ KI KII [3 sin 2 2(1 2 1
2 )2sin ] + 6KIII } 2
(28b)
where F1 represents the normal stress, F2 is an equivalent stress defined by Kong et al. (1995). With the analytical solution to M, the
crack growth angle is assumed as follows:
M 2M
= 0, <0
=
2
= (29)
This criterion has been experimentally validated by Kong et al. [61]. In their study, mixed mode fracture tests were conducted on
FeE 550 steel at a low temperature (-140 °C) with high-speed loading (50 mm/sec), where the specimens turned to be brittle. The
theoretical prediction has shown a good consistency with the experimental results under mixed mode I-II loading condition. How-
ever, as it can be seen from Equation (27) and (28), M is independent of the twisting angle . This criterion may not be applicable to
Mode III dominant conditions.
Schöllmann et al. [48] proposed a criterion based on maximum principal stress 1 (Fig. 2) for predicting the crack growth under
multiaxial loading with all three modes considered. This criterion assumes that the crack will grow in the direction perpendicular to
the maximum principal stress 1 . The unstable crack occurs when the maximum principal stress reaches a critical value 1c .

7
Y. Wang, et al. Engineering Fracture Mechanics 235 (2020) 107126

Richard [18] proposed an empirical criterion for predicting fracture failure and crack growth direction under 3D mixed mode
loading. In this criterion, an equivalent stress intensity factor K eq is introduced. The unstable crack is predicted to occur when the
fracture driving forces are out of the range of the failure envelope described as follows:
KI 1
K eq = + KI2 + 4( 1 KII )2 + 4( 2 KIII )2 = KIC
2 2 (30)
where 1 is the ratio of Mode I fracture toughness to Mode II fracture toughness, 2 is the ratio of Mode I fracture toughness to Mode
III fracture toughness. The crack kinking angle f and twisting angle f can be estimated as follows [19,42]:
2
|KII | |KII |
f = A +B
KI + |KII| + |KIII | KI + |KII| + |KIII | (31)
2
|KIII | |KIII |
f = C +D
KI + |KII| + |KIII | KI + |KII| + |KIII | (32)
where A, B, C, D are the fitting angles, f < 0° for KII > 0 and f > 0° for KII < 0 and KI 0 , f < 0° for KIII > 0 and f > 0° for
KIII < 0 and KI 0 . With 1 = 1.155, 2 = 1, A = 140°, B = − 70°, C = 78°, and D = − 33°, Equations (30), (31), and (32) are in
good agreement with the criterion developed by Schöllmann et al. [19,42,48]. An experimental investigation by Richard et al. [62]
found that the criterion by Richard [18] and by Schöllmann et al. [48] are consistent well with the experimental measurement of
kinking angles and twisting angles.
As the aforementioned criteria are established based on LEFM, they are only valid for brittle materials or for the situations, in
which some ductile metals turn to brittle (e.g., below the ductile to brittle temperature). In these particular cases, the metals are
under small scale yielding (SSY) and the plastic regions are sufficiently small compared to the size of the crack. For ductile metals,
however, crack propagation is often accompanied by considerable non-linear plastic deformation, which is out of the scope of LFEM.
Experimental works by Bhattacharjee and Knott [22] and Maccagno and Knott [63] also proved that none of the available mixed
mode fracture criteria based on LEFM could accurately predict the crack propagation for steel under mixed mode loading at room
temperature. Therefore, for ductile metals, new fracture criteria for mixed mode loading conditions need to be developed to account
for large plastic deformation.

3.2. Mixed mode fracture criteria based on EPFM

Published literature suggests that mixed mode fracture criteria based on elasto-plastic fracture mechanics (EPFM) are relatively
scarce. According to Pirondi and Donne [11], Keiichiro and Hitoshi [32], and Maccagno and Knott [63], the crack extension of ductile
metals under mixed mode I-II loading can be either tensile type (Mode I dominance) or shear type (Mode II dominance) (see Fig. 1). Li
et al. [30] developed a tension-shear (TS) criterion to assess the crack extension under mixed mode I-II loading based on the gen-
eralised HRR singularity. The generalised HRR singularity was first developed by Shih [44] to characterise the crack-tip fields of
elasto-plastic materials under plane strain condition. The crack-tip stress field of an elasto-plastic body under mixed mode loading can
be expressed as follows:
1 (n + 1)
EJ
ij = 0 2 ij ( , n, M p)
0 In r (33)
where Mp is the near-field mixity, which is defined as follows:

2 ( = 0)
M p = lim tan 1
r 0 r ( = 0) (34)
where is the normal traction near the crack tip, r is the shear traction near the crack tip. ranges from 0 to 1 with
Mp = 0 for Mp
pure Mode II and M p = 1 for pure Mode I. The coordinate system of Equation (34) is shown in Fig. 3.
The type of crack extension is predicted by comparing the near-field mixity M p with its critical value Mcp . The value of critical
mixity Mcp is obtained from the M P curve, as shown in Fig. 5, when the critical c ( c = JIc JIIc ) value is given. The parameter is
defined by Li et al. [30] to represent the ratio of maximum circumferential stress to the maximum shear stress, which can be
expressed as follows:
(n + 1)
max r (M p = 0, = 0) In(I )
=
max ( rr )2 4+ 2
r (M p = 1, = 0) In(II ) (35)
where max denotes taking the maximum with respect to the angular coordinate as shown in Fig. 3, is the integration constant
In
dependent on n and M p , which can be calculated by a numerical associated J-integral method developed by Li [64]. From Fig. 5, it
can be seen that the crack will be a tensile type when M p > Mcp and a shear type with M P < Mcp .
To apply the tension-shear criterion, it is necessary to experimentally determine the ratio of Mode I fracture toughness ( JIc ) to
Mode II fracture toughness (JIIc ). A review of the literature shows that very limited research has been carried out with the exception of
Recho et al. [65]. To use tension-shear criterion for structural steel, Recho et al. [65] carried out a series of mixed mode fracture tests

8
Y. Wang, et al. Engineering Fracture Mechanics 235 (2020) 107126

Fig. 5. Tension-shear transition criterion (modified from Li et al. [30]).

to obtain JIc JIIc by use of 4 mm thick and 90 mm wide compact-tension-shear (CTS) specimens. It is known that the near-field mixity
M p derived from the generalised HRR singularity by Shih [44] is only available for plane strain condition. However, the specimens of
Recho et al. [65] are typical specimens leading to plane stress rather than strict plane strain condition, according to ASTM E1820
(2018). Therefore, this criterion needs to be further verified with experimental results of fracture toughness of Mode I and Mode II
under plane strain condition.
The crack-tip-opening displacement (CTOD) is another parameter used to predict the crack initiation and growth direction in both
LEFM and EPFM [29]. The CTOD criterion states that crack extends when the CTOD reaches a critical value. Under LEFM condition,
the crack growth is along the direction perpendicular to the maximum CTOD. Under EPFM condition, the crack type under mixed
mode I-II loading can be predicted by the competition between the opening component ( I ) and shearing component ( II ) of CTOD. To
be specific, the crack growth will be tensile type when I > II . Inversely, the crack growth will be of the shear manner. To predict the
crack growth direction ( c ), Sutton et al. [46] proposed an empirical expression as follows:

a1 arctan(b1 ) for | | < c


c =
a2 cos(b2 ) | for | |
| c (36)

where = arctan ( ),
II
I
c is regarded as the critical mixity for the transition of tensile type to shear type, and a1, a2 , b1 and b2 are the
fitting parameters.
Ma et al. [29] analytically studied the rationale of CTOD criterion and found that CTOD criterion is equivalent to the maximum
circumferential stress (MCS) criterion and the maximum energy release rate (MERR) criterion for brittle materials. For ductile
materials, the prediction of the crack type by CTOD criterion has shown good agreement with experimental observations from
Amstutz et al. [66,67]. Also, a study by Sutton et al. [46] shown that the CTOD criterion performed well in predicting the direction of
crack growth.

3.3. Application of mixed mode fracture criteria

A number of procedures for structural integrity assessment based on fracture mechanics, e.g., BS7910 [68], R6 revision 4 [69],
SINTAP (1999) [70], have been developed. In these procedures, a two-criterion failure diagram, which is known as the failure
assessment diagram (FAD) with the interaction between fracture failure and plastic collapse, is commonly used to evaluate the
integrity of structures as shown in Fig. 6. When applying the concept of this two-criterion failure diagram to the assessment of mixed
mode fracture, BS7910 [68] and R6 revision 4 [69] defined the FAD curve as follows:
Kr = f (Lr ) forLr < Lr max (37a)

f (Lr ) = 0for Lr Lr max (37b)

with
Kr = K eff Kmat and Lr = P PL (37c)
where Kr is the ratio of applied fracture driving force to Mode I fracture toughness, Lr is the load ratio, P is the applied load, PL is the
corresponding plastic limit load, Lmax
r is defined as the plastic collapse limit, Kmat is the Mode I fracture toughness K eff is the effective

9
Y. Wang, et al. Engineering Fracture Mechanics 235 (2020) 107126

Fig. 6. Failure assessment diagram (modified from Haldey and Lei [71]).

stress intensity factor which can be expressed as follows [69]:

K eff = KI2 + KII2 + KIII


2
(1 ) for Kmat y 0.2 m (38a)

K eff = 2
K12 + 2
KIII (1 ) for Kmat y < 0.2 m (38b)with
3
2KI + 6 KI2 + 8KII2 KI2 + 12KII2 + KI KI2 + 8KII2 2
K12 = and |KI KII | 0.466
8 2KI2 + 18KII2
(38c)

KII
K12 = for KI KII < 0.466
0.7 (38d)
where is the factor representing the contribution of Mode III loading, and usually = 1, m denotes meter that is used for the unit
correction. Once the two parameters Kr and Lr are calculated, a failure curve (as shown in Fig. 6) can be established by different
options. These options depend on the stress–strain curve characteristics of materials, the availability of detailed material properties,
the complexity of cracks and the requirement of accuracy in the assessment. A detailed guideline of choosing which option is
provided in R6 revision 4 [69] . For instance, the option 1 diagram in the R6 revision 4 [69] is recommended for the mixed mode
fracture assessment as follows:

Kr = f (L r ) = 1 + 0.5Lr2 [0.3 + 0.7 exp( 0.65) Lr6] (39a)


with a cut off at:
Lr = Lrmax (39b)
The failure curve established by f (Lr ) is the boundary between safety and failure.
Due to the difficulties in determining the mixed mode fracture toughness of materials, the values of Kmat used in Equations (37)
and (38) are simply assumed as the Mode I fracture toughness in the procedures of BS7910 [68] and R6 revision 4 [69]. However,
there is a lack of theoretical justifications for applying the curves (e.g., Option 1 in R6 revision 4 [69]) to the assessment of mixed
mode fracture failures. For ductile metals, K eff and Kmat can be converted from J-integral from Equation (13). However, the as-
sumption of Mode I fracture toughness as Kmat may lead to an overestimated capacity to resist crack extension for ductile metals, e.g.,
steel, under mixed mode loading [14,72]. Experimental observations [11] and finite element analysis [73] showed that the critical
CTOD of ductile metals is independent on mode mixity. Thus, Pirondi [74] applied mixed mode CTOD criterion to establishing the
FAD curve for mixed mode fracture, which proposed Kr as follows:

e
Kr =
mat (40)
where e is the elastic component of mixed mode CTOD which is derived from Equations (13) and (14), mat is the critical mixed mode
CTOD. For strain hardening materials, the FAD curve is established as follows [74]:
1
1 2 2
Kr = f 1 + Lr for Lr 1
2 (41a)
1

Kr = f Lr
( N1 1) +
1 (3
Lr
1
N ) 2
for Lr > 1
2 (41b)
where Lr is from Equation (37c), N is the exponent of the piecewise power-law model for the stress–strain curve. The use of CTOD

10
Y. Wang, et al. Engineering Fracture Mechanics 235 (2020) 107126

Table 2
Mixed mode I-II specimens for metals.
Specimen Illustration Reference

Asymmetric three-point bending specimen Keiichiro and Hitoshi [32]

Asymmetric four-point bending specimen Keiichiro and Hitoshi [32], Bhattacharjee and Knott [22]

Inclined cracked tension specimen Xu and Li [119], Miao et al. [120]

Modified Arcan specimen and fixtures Greer et al. [121]

Compact tension shear specimen and fixtures Kamat and Hirth [7], Yuan et al. [122]

T-specimen and fixture Demir et al. [123]

criterion in FAD overcomes the concern that Kmat is non-conservative under mixed mode loading. However, the critical value of
mixed mode CTOD is difficult to measure due to the non-uniform deformation of the crack tip in ductile metals [74]. Besides, this
CTOD-based approach is only for the fracture initiation analysis, not for the tearing analysis due to the complexity related to the
growth of crack under mixed mode loading.

4. Experimental work on mixed mode fracture

Unlike well-developed testing guidelines (e.g., ASTM E1820 [75], ISO 12,135 [76]) to measure Mode I fracture toughness, there
are some difficulties in determining mixed mode fracture toughness ( Jtot , c ). The challenges include measuring the local deformation
near crack [35], monitoring the length of crack extension [77] and lack of a universal agreed method to determine the fracture
initiation. Difficulties mainly exist in the experimental set-ups as well as to the analysis of test results. According to a large number of
experimental investigations on fracture behaviour of metals under mixed mode loading, there are several approaches that have
shown great potential to overcome these difficulties.

4.1. Fixtures and specimens for mixed mode fracture testing

In order to determine the mixed mode fracture toughness Jtot , c , numerous experimental set-ups and specimens have been de-
veloped for different combinations of mixed mode loading. For the fracture tests of mixed mode I-II, asymmetric bending and biaxial
loading are the most commonly used method to produce the tension and in-plane shear in the cracked specimens. Specimens and
fixtures used in these two methods for metals are summarised in Table 2.
In general, all specimens and fixtures shown in Table 2 can be employed to study mixed mode I-II fracture behaviour. However,
the asymmetric three-point bending specimens and the inclined cracked tension specimens cannot produce pure mode II condition.

11
Y. Wang, et al. Engineering Fracture Mechanics 235 (2020) 107126

Fig. 7. Mixed mode I-III specimens: (a) modified compact tension specimen with the slanted notch, (b) mixed mode I-III loading fixture and tilted-
tension specimen with straight pre-crack (modified from Ayatollahi and Saboori [79]).

Although the asymmetric four-point bending specimens can offer a wide range of Mode I/Mode II ratios by changing the distance
between the notch and loading point, Aoki et al. [78] indicated that friction from the supports of the four-point bending type
specimens could result in a considerable error in estimating the applied load. The lower half of Table 2 summarises the specimens
using biaxial loading, which can cover a full range of Mode I/Mode II ratios by changing the loading holes (points) in the fixtures. It
should be noted that thin-walled specimens may be pinned to the fixtures aslant, which can lead to torsion during the test. The
specimens must be compactly pinned to the fixtures to avoid the undesired out-of-plane bending (i.e., Mode III loading).
For mixed mode I-III, the combination of tensile stress and out-of-plane shear stress is usually achieved in a modified compact
tension (CT) specimen with a slanted notch under tension as illustrated in Fig. 7 (a). This modified CT specimen can produce different
Mode I/Mode III ratios by increasing the slanted angle without inducing the unwanted mode II loading. The main limitations of this
modified CT specimen include the difficulty of loading the fatigue pre-crack along the slanted notch direction and the unavailability
of the pure mode III loading since this kind of specimen is designed to be loaded on the in-plane tension testing machine.
Bending specimens with an inclined notch are also used to study mixed mode I-III fracture behaviour. However, there is un-
avoidable in-plane shear stress (Mode II component) along the crack tip caused by the reaction forces of supports. Ayatollahi and
Saboori [79] proposed a compact tilted-tension specimen with a straight pre-crack to study the mixed mode I-III fracture behaviour.
As shown in Fig. 7 (b), this testing apparatus can cover the full range of loading from pure mode I to pure mode III by tilting the
tensile forces against the straight crack. Also, the straight notch of the specimen can be loaded to a fatigue pre-crack. However, high
cyclic loading rate may be not applicable due to the heavy fixtures.
In comparison, there is less attention to the study of the fracture behaviour under mixed mode II-III loading, as Mode I component
almost always exists for a cracked body in practice. Saboori and Ayatollahi [80] recently developed a fracture testing apparatus for
this mixed mode fracture condition. The anti-symmetrical loading system can produce a pure shear stress field with no bending
moment at the crack plane. The rotated specimen around the beam axis makes the shear force separated into in-plane and out-of-
plane components to achieve the mixed mode II-III fracture condition. As shown in Fig. 8, the mode mixity can range from pure mode
II to pure mode III by changing the rotation angle α of the bending specimen. For this testing procedure, only some computational
results are available so far. There is a lack of experimental data for comparison and validations.
For more complex mixed mode I-II-III (i.e., 3D mixed mode) fracture tests, there are three fracture testing procedures available
from the published literature, namely compact tension-shear-rotation (CTSR) test, all-fracture-mode (AFM) test and compact tension-
shear-torsion (CTST) test. Chang et al. [20] proposed the compact tension-shear-torsion (CTST) testing procedure to validate the
general mixed mode fracture criterion proposed in their study. As shown in Fig. 9 (a), the tensile and torsional forces (P and T in Fig. 9
(a)) are applied simultaneously through the loading fixtures on the cylindrical specimen with a circumferential notch (Fig. 9 (b)).
Various mode mixities can be achieved by changing the loading angle and the loading combination of tension P and torsion T. The
CTST fracture test was conducted on aluminium alloy under different combinations of loading. The test results of Chang et al. [20]
showed good agreement with the proposed general mixed mode fracture criterion. However, such a cylindrical specimen with
circumferential notch generally cannot show the angle of crack growth, especially the twisting angle, under different mode mixities.
Also, as this testing procedure is designed for validating the mixed mode fracture criteria for brittle materials, no test on ductile

12
Y. Wang, et al. Engineering Fracture Mechanics 235 (2020) 107126

Fig. 8. Rotated singe-edge-cracked bending specimen for mixed II-III loading condition: (a) pure Mode III loading condition, (b) mixed mode II-III
condition (modified from Saboori and Ayatollahi [80]).

Fig. 9. Compact tension-shear-torsion (CTST) specimen for mixed mode I-II-III loading condition: (a) loading fixture, (b) specimen (modified from
Chang et al. [20]).

13
Y. Wang, et al. Engineering Fracture Mechanics 235 (2020) 107126

Fig. 10. All-fracture-mode (AFM) testing fixture and specimen for mixed mode I-II-III loading condition: (a) loading fixture, (b) specimen (modified
from Richard and Kuna [81]).

metals was carried out.


The all-fracture-mode (AFM) testing procedure was proposed by Richard and Kuna [81] to produce a superimposed crack-tip
stress field of three modes by changing the loading direction. As shown in Fig. 10, two force introducing fixtures can apply the tensile
force P to the centre of the specimen. The crack tip of the specimen can be subjected to pure tensile, in-plane or out-of-plane shear
stress by changing the direction of the force P.
When the fixture is loaded by force P with angles and (See Fig. 10 (a)), the normal force N, in-plane shear force Qz and out-of-
plane shear force Q y acting at the crack-tip can be determined as follows [81]:

N = P cos cos , Qz = P sin , Qy = P cos sin (42)

For pure Mode I, = = 0° . For pure Mode II, = 90° and = 0° . For pure Mode III, = 0° and = 90° .
However, due to the large weight (inertia) of the fixture and deformation of the specimen, the all-fracture-mode (AFM) testing
procedure is not suitable for thin-walled alloy and fatigue tests. Therefore, Richard et al. [62] proposed another testing procedure
named compact tension-shear-rotation (CTSR) test. The loading fixture and test specimen are shown in Fig. 11. The tensile force P is
applied through two force introducing fixtures with various loading angle α (Fig. 11 (a)). Like the AFM test, the load P can always
pass through the centre of the specimen (crack-tip). The Mode III is induced by rotating the specimen inside the loading fixture with a
variable angle (Fig. 11 (b)). Although the fracture tests using CTSR specimens have been conducted on Polymethylmethacrylate
(PMMA) and thin-wall aluminium specimens to validate Richard’s criterion, only the fracture toughness of PMMA was measured.
There is a lack of experimental studies on the determination of fracture toughness of ductile metals with CTSR specimens. Also, the
current CTSR testing apparatus designed by Richard et al. [62] has relatively low loading capacity, which limits its application to
some metals (e.g., steel, titanium).

4.2. Measurement of crack extension

For ductile metals where the plastic deformation dominates at the crack tip, the crack often has a continuous process of stable
growth. A plot of J-integral versus crack extension ( a ), known as crack resistance curve (J-R curve), can be employed to characterise
the ductile fracture resistance. In order to construct a crack resistance curve, multiple specimens are often used to measure the crack
extension [14]. Specifically, multiple specimens are tested under various loading levels so that different crack extensions can be
attained. After unloading, each specimen will be fatigued to mark the area of crack extension. Although the measurement is
straightforward, this method has the disadvantage that the experimental process is tedious and multiple specimens are required for J-
R curve.
The electrical potential drop method can also be used to monitor crack extension. In this method, a constant direct current (DC) or
alternating current (AC) can be applied to the specimens. As the crack extends, the electrical resistance of the specimens increases,

14
Y. Wang, et al. Engineering Fracture Mechanics 235 (2020) 107126

Fig. 11. Compact-tension-shear-rotation (CTSR) testing fixture and specimen for mixed mode I-II-III loading condition: (a) loading angle α for
producing Mode II loading, (b) loading angle γ for producing Mode III loading (modified from Richard et al. [62]).

which can be continuously measured by on-line computer data logging systems [82]. A function of voltage difference can estimate the
length of crack extension from a calibration curve in the form of Va V0 verse a W , in which Va is the voltage across the crack length a
and V0 is the voltage across the notch tip [83]. This method has the advantages of simple testing set-ups and the feasibility with
various temperature ranges [84]. However, the potential drop signal can be affected by the shearing of the crack front [77] and the
extensive plastic deformation [45,84].

4.3. Determination of mixed mode fracture toughness

So far in the published literature, the threshold of crack extension under mixed mode loading is defined arbitrarily and there is no
standardised method to determine the mixed mode fracture toughness. To determine a single critical J-integral value as the mixed
mode fracture toughness from J-R curve, Manoharan et al. [85] proposed a blunting construction line with an empirical slope. The
critical value was determined by the intersection of a 0.2 mm offset line and a power-law regression line fitted from test data.
However, the test results on a ductile metal (Armco iron) from Kamat et al. [86] showed that the blunting line recommended by
Manoharan et al. [85] could overestimate the crack extension in low strength and high strain hardening materials due to the crack
blunting. Also, as shown in Fig. 12, the J-R curve did not intersect the 0.2 mm offset blunting line proposed by Manoharan et al. [85]
when Mode I loading dominated.
Fig. 12 suggests that the blunting line proposed by Manoharan et al. [85] is not suitable for determining the mixed mode fracture
toughness of ductile metals. Kamat et al. [86] later proposed a method for determining the mixed mode fracture toughness by the
intersection of the critical stretch zone width (SZWc) line and the power-law curve. This method has good physical meaning. As
shown in Fig. 13, a stretch zone is formed in ductile metals due to the crack tip blunting and the crack extension occurs when the
stretch zone width reaches a critical value. As these two methods can lead to a significant difference in mixed mode fracture

15
Y. Wang, et al. Engineering Fracture Mechanics 235 (2020) 107126

Fig. 12. J-SZWc line (modified from Kamat et al. [86]).

Fig. 13. A schematic of the stretch zone.

toughness (e.g., Fig. 12), a universally agreed method to determine the mixed mode fracture toughness still needs to be confirmed.

4.4. General experimental observations

Fracture behaviour under mixed mode loading can be explicitly observed from experimental results. Table 3 summaries the
experimental observations on the fracture behaviour of metals under the two-dimensional loading.
Most of the available experimental data on mixed mode fracture toughness from the published literature are obtained under
mixed mode I-II or I-III loading conditions. For ductile metals under mixed mode I-II loading, some experimental results (e.g.,
[7,11,22,34,63,72]) indicate that the total fracture toughness decreases with the increasing Mode II loading component. A detailed
review shows that this phenomenon is mainly observed on steels [11,72,87]. With the increase of Mode II loading proportion, the
plastic deformation at the crack tip is confined to a narrow band, dissipating less energy [11]. Therefore, more energy is available for
the crack extension and the total fracture toughness decreases with the increasing Mode II loading component. Likewise, some
experimental observations show that the imposition of mode III loading also results in a reduction in the total fracture toughness of
steels due to the narrowed plastic deformation zone near the crack tip [13,14,88–90].
Other experimental observations on metals with lower fracture toughness, e.g., aluminium alloys [36,91], show that the total
fracture toughness increases with an increasing Mode II loading component. According to Raghavachary et al. [14], the total fracture
toughness under mixed mode loading depends on the overall level of fracture toughness. For those materials with lower fracture
toughness and limited plastic zones, Mode II or Mode III deformation will add redundant work without remarkably changing the
damage process during crack propagation, leading to the increase of total fracture toughness.
The microstructure of metals can have a great effect on the total fracture toughness. For example, it is found in the mixed mode I-
II test results of Kamat and Hirth [7] that aluminium alloys with a high fraction of brittle acicular coarse inclusions increase the
number of void initiation sites. In the process of fracture, the voids linkage and extensive primary voids occur without significant void
growth. As a result, the shear localisation caused by Mode II loading is limited. Therefore, the imposition of Mode II loading does not
have a significant effect on the total fracture toughness. Probably for the same reason, the experimental observations from Mano-
haran et al. [31] showed that the fracture toughness of AISI spherodized 1090 steel, with 0.75% manganese (i.e., brittle acicular
coarse inclusions), remained nearly constant under mixed mode I-III loading.

16
Table 3
Fracture testing observations on metals under 2D mixed mode loading conditions.
Reference Fracture Material Specimen type Fracture characteristic
Y. Wang, et al.

mode

Schroth et al. [124] Mixed I-III HSLA steel Modified CT specimen with a slanted notch The addition of Mode III loading component decreases the total fracture
toughness.Pure Mode III fracture toughness is lower than Mode I
fracture toughness.
Raghavachary et al. [14] Mixed I-III Ni-Cr-Mo-V rotor steel Modified CT specimen with a slanted notch The minimum fracture toughness occurs under mixed mode loading
condition.The addtion of Mode III loading component enhances void
formation near the crack tip.
Manoharan et al. [31] Mixed I-III 1090 steel Modified CT specimen with a slanted notch The total fracture toughness remains nearly a constant with the addition
of Mode III loading component.
Keiichiro and Hitoshi Mixed I-II 6061-T651 aluminium alloy Asymmetric three-point bending specimenandAsymmetric four- The crack extension occurs as a tensile type under Mode I loading
[32] point bending specimen dominance and in a brittle manner.The crack extension occurs as a shear
type under Mode II loading dominance and in a ductile manner.The total
fracture toughness is higher under Mode II loading dominance than that
under Mode I loading dominance.
Feng et al. [38] Mixed I-III 2034 aluminium alloys with Modified CT specimen with a slanted notch The addition of Mode III loading component decreases the total fracture
different Manganese content toughness.The addtion of Mode III loading component enhances early
void formation near the crack tip.High manganese content forms large
particles which result in lower fracture toughness.
Bhattacharjee and Knott Mixed I-II HY 100 steel Asymmetric four-point bending specimen Mixed mode fracture criteria based on LEFM fail to predict the crack
[22] propagation accurately.Crack tip deforms in a manner as one side
blunted, and the other side sharpened with the addition of Mode II
loading component.

17
Kamat and Hirth [7] Mixed I-II 2034 aluminium alloy with Compact tension shear specimen The addition of Mode II loading component decreases the total fracture
different Manganese contents toughness of specimens with low Manganese content.The effect of Mode
II loading component is marginal on the total fracture toughness of
specimens with high Manganese content.Shear localisation near the
crack tip of 2034 aluminium alloy with low Manganese content occurs
under Mode II loading dominance.
Kannan and Hirth [37] Mixed I-III HSLA-80 steel Modified CT specimen with a slanted notch The addition of Mode III loading component decreases the total fracture
toughness.
Roy et al. [34] Mixed I-II 2014-O aluminium alloy Compact tension shear specimen The addition of Mode II loading component decreases the total fracture
toughness.The crack tip constraint has a marginal effect on total fracture
toughness under Mode II loading dominance.
Sha et al. [36] Mixed I-II Ly 12 aluminium alloy Compact tension shear specimen The addition of Mode II loading component increases J -integral and
CTOD.
Pirondi and Donne [11] Mixed I-II StE 550 steel Compact tension shear specimen The addition of Mode II loading component decreases the total fracture
toughness.The crack propagates along the maximum tangential stress
direction under Mode I loading dominance.The crack propagates at a
constant angle from the maximum shear strain direction under Mode II
loading dominance.
Qian and Yang [35] Mixed I-II 5083-H112 aluminium alloy Asymmetric four-point bending specimen The addition of Mode II loading component increases the total fracture
toughness under Mode I loading dominance.The addition of Mode II
loading component decreases the total fracture toughness under Mode II
loading dominance.
Miao et al. [120] Mixed I-II Commercially pure Titanium (CP- Compact tension shear specimen, Asymmetric four-point For compact tension shear specimen and asymmetric four-point bending
Ti) bending specimen and Inclined cracked tension specimen specimen, a transition from a shear type to a tensile type occurs when
the component of mode I plastic mixity parameter Mp increases.For
inclined cracked tension specimen, crack always propagates as tensile
type.
Engineering Fracture Mechanics 235 (2020) 107126
Y. Wang, et al. Engineering Fracture Mechanics 235 (2020) 107126

5. Determination of parameters for mixed mode fracture

5.1. Calculation of J-integral

Most specimens mentioned in the last section (e.g., Table 2) have been employed to conduct experimental studies on fracture
toughness of metals under mixed mode loading. As such, fracture mechanics parameters (e.g., K , J-integral) for these specimens have
been numerically or experimentally determined. As the plastic deformation at the crack tip in ductile metals is relatively large, most
experimental studies must resort to elasto-plastic fracture mechanics (EPFM) and the J-based fracture toughness. According to Zhu
and Joyce [45], the J-integral can be divided into an elastic part and a plastic part as follows:
J = Jel + Jpl (43)
where the elastic part Jel can be determined through Equation (13) and the plastic part Jpl can be estimated as follows [87]:
Upl
Jpl = pl
B (W a) (44)
where Upl is the plastic work done by the applied load, pl is the plastic geometry factor, B is the specimen thickness, W is the
specimen width and a is the pre-crack length. According to Equations (13) and (44), the total J-integral can be expressed as follows:
KI2 + KII2 K2 Upl
J= + III + pl
E 2µ B (W a) (45)
In Equation (45), the stress intensity factors (SIFs) are commonly calculated by numerical methods like the boundary collocation
method [92] and FEA [93]. For some common configurations, e.g., centre cracked plate, thick-walled cylinder with an external
circumferential crack, solutions to SIFs under various loading regimes can be found in handbooks [94–96]. For some other specimens
under different mode mixities, e.g., asymmetric four-point bending specimen, CTS specimen, some researchers [7,38,63,97,98] have
proposed the solutions to SIFs.
For the plastic part Jpl , the plastic work Upl can be measured by the plastic area under the load–displacement curve [99], which
can be directly recorded from the loading machine and clip gauge. For plastic geometry factor pl , FEA on different specimen
configurations and pre-crack length to width ratios can be firstly performed to obtain the J-integral using the domain integral method
[100]. The load line displacement (LLD) can then be estimated as the relative displacement along the loading direction of symmetry
nodes on the two fracture surfaces, and the plastic work Upl can be calculated by the area under the load-LLD curve. With the obtained
Upl
plastic work Upl and geometry information, pl factor can be determined by the slop of the quantities of Jpl (i.e., J Je ) and .
B (W a)
Pirondi and Donnie [99] developed an expression of pl for compact-tension-shear (CTS) specimen under mixed mode I-II through
FEA, which is expressed as follows:

pl = 0.915 + 1.141Me 5.566Me2 + 14.1Me3 8.191Me4 (46)


where Me is determined by Equation (9). Equation (46) is only valid for CTS specimen under plane stress condition [99].
According to the superposition principle of energy, the total J-integral can also be calculated by the sum of Mode I, Mode II and
Mode III components, which can be expressed as follows:
KI2 + KII2 K2 1
Jtot = JI + II + III = + III + ( pl, I Upl, I + pl, II Upl, II + pl, III Upl, III )
E 2µ B (W a) (47)
where pl, I = 2 for CTS specimen [7], pl, II = 1 for CTS specimen [7], pl, II = 0.9 for Arcan shape specimen [101], pl, III = 2 for CTS
specimen [38].

5.2. Measurement of CTOD

Compared to determining J-integral under mixed mode fracture conditions, measuring the CTOD of cracked specimens under
mixed mode loading is straightforward. For example, the CTOD of mixed mode I-II can be determined by measuring the open and
shear-displacement components as:

v = I
2 + II
2
(48)
where v is the CTOD under mixed mode loading condition, I and II are the opening and shearing displacements, respectively. As
shown in Fig. 14, these two parameters can be determined by measuring the relative displacement of the grids on the surface of a
specimen by an optical microscope.
Digital image correlation (DIC) technology can also be employed to measure the mixed mode CTOD. Amstutz et al. [67] are one of
the first researchers who employed DIC to measure the CTOD of metals under mixed mode I-II loading conditions. In Amstutz et al.
[67], one side of the specimen was spray painted as a black-white random pattern for the measurement of displacement. Three pairs
of subsets were chosen at a typical distance behind the crack tip. The CTOD was determined by the average relative displacements of
these subsets. Similarly, the CTOD of metals under mixed mode I-III can be measured by taking advantage of 3D DIC technology,
which can take the measurement of full-field deformation of a non-planar body. Fig. 15 shows a schematic of the measurement of

18
Y. Wang, et al. Engineering Fracture Mechanics 235 (2020) 107126

Fig. 14. Measuring CTOD under mixed mode I-II loading.

Fig. 15. A schematic of CTOD measurement using DIC under mixed mode loading condition: (a) image before crack extension, (b) image after crack
extension.

CTOD by DIC under mixed mode loading.


The mixed mode CTOD can be optically measured on the facture surface at a fixed distance behind the crack tip. According to
Sutton et al. [46], the chosen typical distance behind the crack tip is the key to measure the CTOD accurately. For a small distance,
the CTOD may not be measured due to the crack blunting and extension. On the other hand, the CTOD measured from a far distance
behind the crack tip may include the deformations in the far-field, which cannot represent the true deformation around the crack tip.
Also, it was found that an appropriate distance for measuring the mixed mode CTOD depends on the material properties of the
specimen [46]. At the moment, there is no reference guide for choosing the distance to measure the mixed mode CTOD. A measuring
standard for mixed mode CTOD of different materials needs to be established.

5.3. Two-parameter methods

A single parameter (K , G , J-integral, or CTOD) can characterise the crack tip field of materials under small-scale yielding (SSY)
conditions, but may not work when the fracture occurs with excessive plastic deformation, e.g., under the large scale yielding (LSY)
condition [102]. Also, crack size, specimen configuration, section thickness, and loading types, which are referred as constraint
effects, can have a significant effect on the fracture toughness of specimens under LSY [103]. To address the limitations of single-
parameter fracture mechanics and quantify the constraint effect, two-parameter fracture mechanics, e.g., J-T [104], J-Q [27,28], have

19
Y. Wang, et al. Engineering Fracture Mechanics 235 (2020) 107126

been developed. In these methods, the second parameters (e.g., T, Q) are introduced to quantify the constraint effect, which is largely
associated with loading, material properties and specimen geometry [102]. Several studies (e.g., [105,106]) have investigated the
constraint effect with T-stress, which is the second term of Williams series expansion, on the cracked metals under mixed mode
loading. However, the T-stress is a LEFM-based parameter which becomes inappropriate for large plastic near the crack tip [45].
The two-parameter J-Q theory, which is a numerical method in essence, has been established by O'dowd and Shih [27,28] for
analysing the crack tip stress field of elasto-plastic materials under LSY. In the J-Q theory, the stress field near the crack tip can be
obtained as follows [45,107]:
J
ij = ( ij ) HRR + Q 0 ij, for r > and
0 2 (49)
where ( ij ) HRR is the HRR stress field in Equation (11), ij is the Kronecker delta, Q is a measurement of stress triaxiality parameter (or
crack tip constraint) which is obtained from finite element analysis (FEA) results of crack opening stress ( ) FEA as follows:
( ) FEA ( ) HRR 2J
Q= , at r = and =0
0 0 (50)
The solutions to the parameter Q have been proposed by O'dowd and Shih [27,28,108] for some basic body geometries at
r 0 J = 2 . The distance r from calculated point to crack tip is proposed to be just outside the finite strain blunting zone since the Q
values obtained from a small strain analysis should not be considerably affected [27].
A review of the literature suggests that the J-Q theory has not been well applied to metals under mixed mode loading conditions.
Oskui et al. [107] are one of the few researchers who investigated the constraint effect under the mixed mode I-II loading. In their
study, J-Q theory was employed, and FEA was performed on a modified Arcan specimen, which is commonly used for producing
mode mixities by changing loading angles in the fixtures (see Table 1), using FE software ABAQUS. According to their analysis, the Q
value increased with the increasing loading angles and crack length ratio, leading to a decrease of J-integral. This phenomenon may
be attributed to the faster plastic strain localisation with the increase of Mode II loading and the accelerated microcrack growth due
to the increasing crack length ratio. However, the work by Oskui et al. [107] was only limited to the AM 60 magnesium alloy with the
modified Arcan specimen. More studies on commonly used specimens with different metals are recommended to further understand
the Q-stress effect.

6. Further discussion and future work

6.1. Mixed mode fracture criteria

From the above literature review, it can be seen that most criteria for mixed mode fracture have been established based on linear
elastic fracture mechanics (LEFM). Mixed mode fracture criteria based on LEFM can be applied to metals under small scale yielding
conditions where the size of plastic zone is considerably small, or low temperature conditions (temperature below ductile-to-brittle
transition). For 3D (I-II-III) mixed mode fracture criteria, the maximum energy release rate (MERR) criterion needs to calculate the
maximum energy release rate, the minimum strain energy density (MSED) criterion needs to calculate the minimum strain energy
density and Kong’s criterion needs to calculate the maximum triaxial stress. According to Kong et al. [61], only the triaxial stress has a
single maximum value near the crack tip, whereas the other two have multiple extrema. Thus, Kong’s criterion theoretically performs
better than MERR criterion and MSED criterion. However, all these three criteria are not available for predicting the out-of-plane
shear as their mathematical expressions are independent of the twisting angle ( ). The criterion by Schöllmann et al. can predict the
twisting angles with physical meanings, but the complexity of stress analysis is inevitable. Richard’s criterion has simple mathe-
matical expressions to predict the fracture failure, kinking angles and twisting angles, but its fitting parameters (e.g., 1, 2 ) are highly
dependent on material properties. To apply Richard’s criterion, the experimental determination of the fitting parameters for different
materials is necessary.
Some contradictions have been found between the prediction from LEFM based criteria and experimental investigations. For
instance, an experimental study conducted by Richard et al. [62] found that the criterion by Pook [59,60] for predicting the crack
angles showed great deviations from the measured crack angles when Mode III loading applied. Besides, experimental investigations
carried out by Demir et al. [109] indicated that some existing criteria (e.g., MTS criterion, the criterion by Richard) somehow showed
disagreements with the experimental observations when the Mode II loading was dominating. One of the possible reasons, as some
researchers, e.g., Smith et al. [110], Saghafi et al. [111], indicated that most criteria only include the first order term (i.e., K field) in
the series expansion of crack-tip stress field. However, the higher order terms can have a significant influence on the mixed mode
fracture [110], which needs to be further investigated.
For mixed mode fracture criteria based on EPFM, the tension-shear criterion can be used to predict the crack type (shear or tensile
type) of a crack under mixed mode I-II loading. The parameters used in this criterion are derived from the generalised HRR sin-
gularity, which is only available for plane strain condition. The review of literature in this paper shows that there is a lack of
experimental validations. Further experimental work is necessary to evaluate its accuracy using the specimens under plane strain
condition and to determine the values of required critical parameter, e.g., JIc , JIIc and Mcp . Also, the CTOD criterion was primarily
verified by experimental data on thin-walled aluminium alloys under plane stress and mixed mode I-II loading conditions. Its ap-
plicability to other mixed mode cases, e.g., mixed mode I-III, mixed mode I-II-III, and other metals, e.g., structural steel, are unclear.
Further experimental verifications of CTOD criterion are necessary on other metals under plane strain condition and other mixed

20
Y. Wang, et al. Engineering Fracture Mechanics 235 (2020) 107126

mode fracture cases.

6.2. Experimental work on mixed mode fracture

It is known that most of the metals are often subjected to mixed mode loading conditions. In order to apply conservative values for
fracture toughness to engineering practices, it is necessary to develop reliable fracture toughness testing apparatus for mixed mode
loading conditions. The friction effect from supports of bending specimens could impose an adverse impact on mixed mode fracture
growth. Thus, the pre-cracked specimens under multiaxial loading (e.g., CTS specimens, Arcan specimens) are recommended for
mixed mode fracture tests. From the review of 3D fracture testing procedures, it is found that the all-fracture-mode (AFM) testing
apparatus is not suitable for producing fatigue pre-crack due to the large weight (inertia) of fixtures [62]. The compact tension-shear-
torsion (CTST) specimens generally cannot demonstrate a clear crack growth direction. Therefore, the compact tension-shear-rotation
(CTSR) specimen and fixture are suggested to be used for the tests of 3D ductile fracture toughness due to the full range of mode
mixity and the similar geometrical features with the standard CT specimen. The specimen and fixture recently developed by Razavi
and Berto [112] with the same advantages are also recommended for the 3D fracture tests on ductile metals. As can be seen in Fig. 11,
the specimen has very limited space to allow an optical system to be applied. This can cause difficulties in measuring the local surface
deformation and CTOD. Therefore, a new design of specimen and fixture is required to eliminate this issue.
The effect of mode mixity on the total fracture toughness highly depends on material properties. For metals with low fracture
toughness, the total fracture toughness is relatively insensitive to the mode mixities or slightly increases with the imposition of Mode
II or Mode III loading. For metals with high fracture toughness, the total fracture toughness often shows a considerable decrease with
the addition of Mode II or Mode III component. Besides, the minimum total J-integral of metals with high total fracture toughness
most likely occurs under Mode II or Mode III dominance [11,14,33,35,77], from which shear type crack appears. In this case, Mode I
fracture toughness cannot be used as a conservative design value. As such, more work on mixed mode fracture toughness is re-
commended on metals when it is used as a material property for engineering design and structural integrity assessment.
To further understand the mechanisms of mode mixity effect on total fracture toughness, it is recommended to investigate the
microstructure of materials. It has been found that the decrease of total fracture toughness under low mode mixities (Mode II or Mode
III dominance) is caused by the shear localisation and early formed voids. Moreover, the material composition may have a great effect
on fracture behaviour under mixed mode loading. The review of literature suggests that investigations of microstructural effect on the
total fracture toughness of steel are relatively scarce. It is also necessary to investigate the effect of element composition, phases and
grain size on the fracture behaviour of steel under mixed mode loading.

6.3. Determination of parameters for mixed mode fracture

The use of the plastic geometry factor pl in Equations (45) and (47) can simplify the calculation of J-integral. The values of plastic
geometry factor are dependent on the specimen geometries, pre-crack length to width ratios and mode mixities [11]. The slop of
Upl
terms and Jpl from FEA results can be employed to determine pl factors for different specimen configurations. The critical
B (W a)
value of J-integral is highly dependent on the constraint effect, which can be quantified by Q stress for elasto-plastic materials.
According to Roy et al. [34], the negative Q stress can lead to a constraint loss, which physically results in a delay of void growth and
fracture initiation. Thus far, studies of constraint effect and two-parameter method for cracked metals under mixed mode loading are
very limited. More numerical and experimental studies on constraint effect, such as the effect of specimen thickness, the effect of
crack depth and the effect of specimen configurations, need to be conducted for cracked metals under mixed mode loading.
The mixed mode CTOD can be measured from a specified distance behind the crack tip by an optical microscope or DIC tech-
nology. The typical locations chosen can have a great effect on the accuracy of the measured CTOD under mixed mode loading, and
the distance chosen is most likely material dependent. A standardised method of choosing the locations for different materials is
recommended to be established for accurate measurement of CTOD under mixed mode loading.

7. Conclusion

This paper has mainly reviewed the fracture mechanics parameters for mixed mode loading conditions, mixed mode fracture
criteria and experimental work on fracture behaviour of metals under mixed mode loading. Whilst significant progress has been
achieved in the study of mixed mode fracture behaviour of metals, there is still space to advance knowledge of fracture mechanics, in
particular, the mechanisms of mixed mode fracture and elasto-plastic fracture mechanics (EPFM) with a view to extend wider
applications of fracture mechanics. From this review, key conclusions and future research on mixed mode fracture of metals can be
brought forward:

(1) Mixed mode fracture criteria are mainly established based on LEFM and the scopes of their applications are limited to predicting
the fracture of metals with small plastic deformation zone. For elasto-plastic or ductile metals with considerable plastic de-
formation near the crack tip, EPFM based parameters, i.e., J-integral, CTOD and M P are recommended in establishing the mixed
mode fracture criteria.
(2) In EPFM, the CTOD criterion provides a simple tool to predict the crack initiation and growth direction. But this criterion is
mainly verified for metals under mixed mode I-II loading condition. Other loading conditions, such as the mixed mode I-III

21
Y. Wang, et al. Engineering Fracture Mechanics 235 (2020) 107126

condition or more complex mixed mode I-II-III condition, need to be experimentally verified. With the development of optical
measuring devices, this is possible.
(3) Also in EPFM, the tension-shear criterion has been numerically verified for predicting the crack type (shear or tensile type) in
metals under mixed mode I-II loading and only experimentally verified under plane stress condition. This criterion needs to be
further verified from fracture tests on metal specimens under plane strain and mixed mode I-II loading conditions.
(4) For ductile metals under mixed mode loading, the addition of Mode II or Mode III loading can lead to a remarkable decrease of
total fracture toughness, while the total fracture toughness of less ductile metals is insensitive to the addition of Mode II or Mode
III loading. The effect of mode mixity on fracture behaviour of cracked metals can be better understood by investigating their
microstructure, such as element composition and grain size.
(5) Up to date, there is no standardised testing method for total fracture toughness of metals under mixed mode loading. Such a
testing procedure is highly recommended to be developed which should include designing specimens, establishing J-R curve and
determining the total fracture toughness Jtot , c . Specimens in Table 2 are recommended as a start with some possible improvements
in, e.g., support friction, undesirable loading and the capacity of loading fixtures.
(6) Equation (45) or (47) is recommended to calculate the J-integral for metals under mixed mode loading where the crack extension
can be measured straightforwardly by breaking the specimens after tests. Also, potential drop method can be employed to
monitor the crack extension, but its accuracy is dubious. The method proposed by Kamat et al. [86] is recommended to determine
the mixed mode fracture toughness, as shown in Fig. 12.

Declaration of Competing Interest

The authors declared that there is no conflict of interest.

Acknowledgements

Financial support from the Australian Research Council under DP140101547, LP150100413 and DP170102211, and the National
Natural Science Foundation of China with Grant No. 51820105014 is gratefully acknowledged.

References

[1] Li C-Q, Fu G, Yang W, Yang S. J-Integral solution for elastic fracture toughness for plates with inclined cracks under biaxial loading. J Engng Mech
2018;144(6):04018026. https://doi.org/10.1061/(ASCE)EM.1943-7889.0001444.
[2] Anderson TL. Fracture mechanics: Fundamentals and Applications, fourth edition. Boca Raton: CRC press; 2017.
[3] Erdogan F. Fracture mechanics. Int J Solids Struct 2000;37(1):171–83. https://doi.org/10.1016/S0020-7683(99)00086-4.
[4] Welss AA. The conditions for fast fracture in aluminium alloys with particular reference to the comet failures. British Welding Research Association Report
1955.
[5] Fu G, Yang W, Li C-Q. Stress intensity factors for mixed mode fracture induced by inclined cracks in pipes under axial tension and bending. Theor Appl Fract
Mech 2017;89:100–9. https://doi.org/10.1016/j.tafmec.2017.02.001.
[6] Makar JM, Desnoyers R, McDonald SE. Failure modes and mechanisms in gray cast iron pipe. In: Knight M, Thomson N, (Ed.). Underground Infrastructure
Research: Municipal Industrial and Environmental Applications: Proceedings of the International Conference on Underground Infrastructure Research;
Kitchener, Ontario; CRC Press; 2001; 1-10.
[7] Kamat SV, Hirth JP. Mixed mode I/II fracture toughness of 2034 aluminum alloys. Acta Mater 1996;44(1):201–8. https://doi.org/10.1016/1359-6454(95)
00169-8.
[8] Ast J, Ghidelli M, Durst K, Göken M, Sebastiani M, Korsunsky AM. A review of experimental approaches to fracture toughness evaluation at the micro-scale.
Mater Des 2019;173:107762https://doi.org/10.1016/j.matdes.2019.107762.
[9] Zerbst U, Klinger C, Clegg R. Fracture mechanics as a tool in failure analysis — Prospects and limitations. Engng Fail Anal 2015;55:376–410. https://doi.org/10.
1016/j.engfailanal.2015.07.001.
[10] Chao YJ, Liu S. On the failure of cracks under mixed-mode loads. Int J Fract 1997;87(3):201–23. https://doi.org/10.1023/A:1007499309587.
[11] Pirondi A, Dalle Donne C. Characterisation of ductile mixed-mode fracture with the crack-tip displacement vector. Engng Fract Mech 2001;68(12):1385–402.
https://doi.org/10.1016/S0013-7944(01)00023-6.
[12] Srinivas M, Kamat S, Rao PR. Mixed mode I/III fracture toughness of mild steel. In: ICF10; Honolulu; 2001.
[13] Schroth JG, Hirth JP, Hoagland RG, Rosenfield AR. Combined mode I-mode III fracture of a high strength low-alloy steel. Metall Trans A 1991;18(6):1061–72.
https://doi.org/10.1007/BF03325716.
[14] Raghavachary S, Rosenfield AR, Hirth JP. Mixed mode I/III fracture toughness of an experimental rotor steel. Metall Trans A 1990;21(9):2539–45. https://doi.
org/10.1007/BF02646999.
[15] Erdogan F, Sih GC. On the crack extension in plates under plane loading and transverse shear. J Basic Engng 1963;85(4):519–25. https://doi.org/10.1115/1.
3656897.
[16] Sih GC. Strain-energy-density factor applied to mixed mode crack problems. Int J Fract 1974;10(3):305–21. https://doi.org/10.1007/BF00035493.
[17] Sih GC. Methods of Analysis and Solution of Crack Problems: Recent Developments in Fracture Mechanics, Theory and Methods of Solving Crack Problems.
Springer; 1973.
[18] Richard HA. Experimental and numerical simulation of mixed-mode crack growth. In: Andrea C, Freitas Md, Spagnoli A, (Ed.). The Sixth International
Conference on Biaxial/Multiaxial Fatigue & Fracture; Lisbon; 2001.
[19] Richard HA, Eberlein A, Kullmer G. Concepts and experimental results for stable and unstable crack growth under 3D-mixed-mode-loadings. Engng Fract Mech
2017;174:10–20. https://doi.org/10.1016/j.engfracmech.2016.12.005.
[20] Chang J, Xu J-q, Mutoh Y. A general mixed-mode brittle fracture criterion for cracked materials. Engng Fract Mech 2006;73(9):1249–63. https://doi.org/10.
1016/j.engfracmech.2005.12.011.
[21] Hussain M, Pu SL, Underwood J. Strain Energy Release Rate for a Crack Under Combined Mode I and Mode II. In: Irwin G, (Ed.). Fracture Analysis: Proceedings
of the 1973 National Symposium on Fracture Mechanics, Part II; West Conshohocken, PA; ASTM International; 1974; 2-28. 10.1520/STP33130S.
[22] Bhattacharjee D, Knott JF. Ductile fracture in HY100 steel under mixed mode I/Mode II loading. Acta Metall Mater 1994;42(5):1747–54. https://doi.org/10.
1016/0956-7151(94)90385-9.
[23] Rice JR. A path independent integral and the approximate analysis of strain concentration by notches and cracks. J Appl Mech 1968;35(2):379–86. https://doi.

22
Y. Wang, et al. Engineering Fracture Mechanics 235 (2020) 107126

org/10.1115/1.3601206.
[24] Hutchinson JW. Singular behaviour at the end of a tensile crack in a hardening material. J Mech Phys Solids 1968;16(1):13–31. https://doi.org/10.1016/0022-
5096(68)90014-8.
[25] Rice JR, Rosengren GF. Plane strain deformation near a crack tip in a power-law hardening material. J Mech Phys Solids 1968;16(1):1–12. https://doi.org/10.
1016/0022-5096(68)90013-6.
[26] Wells AA. Crack opening displacements from elastic-plastic analyses of externally notched tension bars. Engng Fract Mech 1969;1(3):399–410. https://doi.org/
10.1016/0013-7944(69)90001-0.
[27] O'Dowd NP, Shih CF. Two-Parameter Fracture Mechanics: Theory and Applications. In: Landes J, McCabe D, Boulet J, (Ed.). Fracture Mechanics: Twenty-Fourth
Volume; West Conshohocken, PA; ASTM International; 1994; 21-47. 10.1520/STP13698S.
[28] O'Dowd NP, Shih CF. Family of crack-tip fields characterized by a triaxiality parameter—I. Structure of fields. J Mech Phys Solids 1991;39(8):989–1015.
https://doi.org/10.1016/0022-5096(91)90049-T.
[29] Ma F, Deng X, Sutton MA, Newman JC. A CTOD-Based Mixed-Mode Fracture Criterion. In: Miller K, McDowell D, (Ed.). Mixed-Mode Crack Behavior; West
Conshohocken, PA; ASTM International; 1999; 86-110. 10.1520/STP14245S.
[30] Li J, Zhang X-B, Recho N. J-Mp based criteria for bifurcation assessment of a crack in elastic–plastic materials under mixed mode I–II loading. Engng Fract Mech
2004;71(3):329–43. https://doi.org/10.1016/S0013-7944(03)00117-6.
[31] Manoharan M, Hirth JP, Rosenfield AR. Combined mode I-mode III fracture toughness of a spherodized 1090 steel. Acta Metall Mater 1991;39(6):1203–10.
https://doi.org/10.1016/0956-7151(91)90208-I.
[32] Keiichiro T, Hitoshi I. Elastic-plastic fracture toughness test under mixed mode I-II loading. Engng Fract Mech 1992;41(4):529–40. https://doi.org/10.1016/
0013-7944(92)90299-T.
[33] Densley JM, Hirth JP. Mixed mode fracture of an HSLA-100 steel. Scr Mater 1998;39(7):881–5. https://doi.org/10.1016/S1359-6462(98)00266-8.
[34] Arun Roy Y, Narasimhan R, Arora PR. An experimental investigation of constraint effects on mixed mode fracture initiation in a ductile aluminium alloy. Acta
Mater 1999;47(5):1587–96. https://doi.org/10.1016/S1359-6454(99)00015-4.
[35] Qian X, Yang W. Initiation of ductile fracture in mixed-mode I and II aluminum alloy specimens. Engng Fract Mech 2012;93:189–203. https://doi.org/10.1016/
j.engfracmech.2012.06.018.
[36] Sha J, Sun J, Zuh P, Deng Z, Zhou H. Study of the relationship between J-integral and COD parameters under mixed mode I + II loading in aluminum alloy Ly
12. Int J Fract 2000;104(4):409–23. https://doi.org/10.1023/A:1007647230722.
[37] Kannan K, Hirth JP. Mixed mode fracture toughness and low cycle fatigue behavior in an HSLA-80 steel. Scr Mater 1998;39(6):743–8. https://doi.org/10.1016/
S1359-6462(98)00185-7.
[38] Feng X, Kumar AM, Hirth JP. Mixed mode I/III fracture toughness of 2034 aluminum alloys. Acta Metall Mater 1993;41(9):2755–64. https://doi.org/10.1016/
0956-7151(93)90144-H.
[39] Rozumek D, Macha E. A survey of failure criteria and parameters in mixed-mode fatigue crack growth. Mater Sci 2009;45(2):190. https://doi.org/10.1007/
s11003-009-9179-2.
[40] Qian J, Fatemi A. Mixed mode fatigue crack growth: A literature survey. Engng Fract Mech 1996;55(6):969–90. https://doi.org/10.1016/S0013-7944(96)
00071-9.
[41] Pook LP. Mixed-mode fatigue crack growth thresholds: A personal historical review of work at the National Engineering Laboratory, 1975–1989. Engng Fract
Mech 2018;187:115–41. https://doi.org/10.1016/j.engfracmech.2017.10.028.
[42] Richard HA, Fulland M, Sander M. Theoretical crack path prediction. Fatigue Fract Engng Mater Struct 2005;28(1–2):3–12. https://doi.org/10.1111/j.1460-
2695.2004.00855.x.
[43] Williams M. On the stress distribution at the base of a stationary crack. J Appl Mech 1956;24(1):109–14.
[44] Shih CF. Small-Scale Yielding Analysis of Mixed Mode Plane-Strain Crack Problems. In: Irwin G, (Ed.). Fracture Analysis: Proceedings of the 1973 National
Symposium on Fracture Mechanics, Part II; West Conshohocken, PA; ASTM International; 1974. 10.1520/STP33141S.
[45] Zhu X-K, Joyce JA. Review of fracture toughness (G, K, J, CTOD, CTOA) testing and standardization. Engng Fract Mech 2012;85:1–46. https://doi.org/10.1016/
j.engfracmech.2012.02.001.
[46] Sutton MA, Deng X, Ma F, Newman Jr JC, James M. Development and application of a crack tip opening displacement-based mixed mode fracture criterion. Int
J Solids Struct 2000;37(26):3591–618. https://doi.org/10.1016/S0020-7683(99)00055-4.
[47] Ashby MF, Jones DRH. Engineering materials 1: an introduction to properties, applications and design. Elsevier; 2012.
[48] Schöllmann M, Richard HA, Kullmer G, Fulland M. A new criterion for the prediction of crack development in multiaxially loaded structures. Int J Fract
2002;117(2):129–41. https://doi.org/10.1023/A:1020980311611.
[49] Hua G, Brown MW, Miller KJ. Mixed-mode fatigue thresholds. Fatigue Fract Engng Mater Struct 1982;5(1):1–17. https://doi.org/10.1111/j.1460-2695.1982.
tb01220.x.
[50] Sih GC, Barthelemy BM. Mixed mode fatigue crack growth predictions. Engng Fract Mech 1980;13(3):439–51. https://doi.org/10.1016/0013-7944(80)
90076-4.
[51] Badaliance R. Application of strain energy density factor to fatigue crack growth analysis. Engng Fract Mech 1980;13(3):657–66. https://doi.org/10.1016/
0013-7944(80)90094-6.
[52] Sih G. A three-dimensional strain energy density factor theory of crack propagation. Mechanics of Fracture Initiation and Propagation: Springer; 1991. p. 23-56.
[53] Mageed AMA, Pandey RK. Mixed mode crack growth under static and cyclic loading in A1-alloy sheets. Engng Fract Mech 1991;40(2):371–85. https://doi.org/
10.1016/0013-7944(91)90271-2.
[54] Abdel Mageed AM, Pandey RK. Studies on cyclic crack path and the mixed-mode crack closure behaviour in Al alloy. Int J Fatigue 1992;14(1):21–9. https://doi.
org/10.1016/0142-1123(92)90149-7.
[55] Wang J, Ren L, Xie LZ, Xie HP, Ai T. Maximum mean principal stress criterion for three-dimensional brittle fracture. Int J Solids Struct 2016;102–103:142–54.
https://doi.org/10.1016/j.ijsolstr.2016.10.009.
[56] Salimi-Majd D, Shahabi F, Mohammadi B. Effective local stress intensity factor criterion for prediction of crack growth trajectory under mixed mode fracture
conditions. Theor Appl Fract Mech 2016;85:207–16. https://doi.org/10.1016/j.tafmec.2016.01.009.
[57] Swankie TD, Smith DJ. Low temperature mixed mode fracture of a pressure vessel steel subject to prior loading. Engng Fract Mech 1998;61(3):387–405.
https://doi.org/10.1016/S0013-7944(98)00065-4.
[58] Maccagno TM, Knott JF. The low temperature brittle fracture behaviour of steel in mixed modes I and II. Engng Fract Mech 1991;38(2):111–28. https://doi.org/
10.1016/0013-7944(91)90076-D.
[59] Pook LP. Comments on Fatigue Crack Growth Under Mixed Modes I and III and Pure Mode III Loading. In: Miller KJ, Brown MW, editors. West Conshohocken,
PA: ASTM International; 1985. p. 249-63.
[60] Pook LP. The significance of mode 1 branch cracks for combined mode failure. In: Radon JC, editor. Fracture and Fatigue: Pergamon; 1980. p. 143-53.
[61] Kong XM, Schlüter N, Dahl W. Effect of triaxial stress on mixed-mode fracture. Engng Fract Mech 1995;52(2):379–88. https://doi.org/10.1016/0013-7944(94)
00228-A.
[62] Richard HA, Eberlein A. Experiments on cracks under spatial loading. ICF132013.
[63] Maccagno TM, Knott JF. The mixed mode I/II fracture behaviour of lightly tempered HY130 steel at room temperature. Engng Fract Mech 1992;41(6):805–20.
https://doi.org/10.1016/0013-7944(92)90233-5.
[64] Li J. Estimation of the mixity parameter of a plane strain elastic–plastic crack by using the associated J-integral. Engng Fract Mech 1998;61(3):355–68. https://
doi.org/10.1016/S0013-7944(98)00066-6.
[65] Recho N, Ma S, Zhang X. Criteria for mixed-mode fracture prediction in ductile material. ECF15Stockolm 2004.
[66] Amstutz BE, Sutton MA, Dawicke DS, Boone ML. Effects of Mixed Mode I/II Loading and Grain Orientation on Crack Initiation and Stable Tearing in 2024-T3

23
Y. Wang, et al. Engineering Fracture Mechanics 235 (2020) 107126

Aluminum. In: Piascik RS, Newman JC, Dowling NE, editors. West Conshohocken, PA: ASTM International; 1997. p. 105-25.
[67] Amstutz BE, Sutton MA, Dawicke DS, Newman JC. An Experimental Study of CTOD for Mode I/Mode II Stable Crack Growth in Thin 2024-T3 Aluminum
Specimens. In: Reuter W, Underwood JH, Newman JC, editors. West Conshohocken, PA: ASTM International; 1995. p. 256-71.
[68] BSI-7910. Guide to methods for assessing the acceptability of flaws in metallic structures. London: British Standards Institution; 2013.
[69] R6. Assessment of the integrity of structure containing defect revision 4. London: British Energy Generation Ltd; 2001.
[70] SINTAP. Structural Integrity Assessment Procedures for European Industry. CORUS U.K. LTD; 1999.
[71] Hadley I, Lei Y. Outline of the fracture clauses of BS 7910:2013. Int J Press Vessels Pip 2018;168:289–300. https://doi.org/10.1016/j.ijpvp.2018.11.004.
[72] Smith DJ, Swankie TD, Pavier MJ, Smith MC. The effect of specimen dimensions on mixed mode ductile fracture. Engng Fract Mech 2008;75(15):4394–409.
https://doi.org/10.1016/j.engfracmech.2008.04.005.
[73] Ghosal AK, Narasimhan R. A finite element analysis of mixed-mode fracture initiation by ductile failure mechanisms. J Mech Phys Solids 1994;42(6):953–78.
https://doi.org/10.1016/0022-5096(94)90080-9.
[74] Pirondi A. Suitability of mixed-mode I/II assessment methods for implementation into the SINTAP procedure. Engng Fract Mech 2003;70(13):1597–609.
https://doi.org/10.1016/S0013-7944(02)00201-1.
[75] ASTM-E1820. Standard Test Method for Measurement of Fracture Toughness. West Conshohocken, PA: ASTM International; 2018.
[76] ISO-12135. Metallic materials—unified method of test for the determination of quasistatic fracture toughness. The International Organization for
Standardization; 2016.
[77] Laukkanen A, Wallin K, Rintamaa R. Evaluation of the Effects of Mixed Mode I-II Loading on Elastic-Plastic Ductile Fracture of Metallic Materials. In: Miller KJ,
McDowell DL, editors. West Conshohocken, PA: ASTM International; 1999. p. 3-20.
[78] Aoki S, Kishimoto K, Yoshida T, Sakata M, Richard HA. Elastic-plastic fracture behavior of an aluminum alloy under mixed mode loading. J Mech Phys Solids
1990;38(2):195–213. https://doi.org/10.1016/0022-5096(90)90034-2.
[79] Ayatollahi MR, Saboori B. A new fixture for fracture tests under mixed mode I/III loading. Eur J Mech A Solids 2015;51:67–76. https://doi.org/10.1016/j.
euromechsol.2014.09.012.
[80] Saboori B, Ayatollahi MR. A novel test configuration designed for investigating mixed mode II/III fracture. Engng Fract Mech 2018;197:248–58. https://doi.
org/10.1016/j.engfracmech.2018.04.048.
[81] Richard HA, Kuna M. Theoretical and experimental study of superimposed fracture modes I, II and III. Eng Fracture Mechanics 1990;35(6):949–60. https://doi.
org/10.1016/0013-7944(90)90124-Y.
[82] Ikeda K, Yoshimi M, Miki C. Electrical potential drop method for evaluating crack depth. Int J Fract 1991;47(1):25–38. https://doi.org/10.1007/BF00037037.
[83] Aronson G, Ritchie R. Optimization of the electrical potential technique for crack growth monitoring in compact test pieces using finite element analysis. J Test
Eval 1979;7(4):208–15. https://doi.org/10.1520/JTE11382J.
[84] Gajji A, Sasikala G. Potential drop method for online crack length measurement during fracture testing: development of a correction procedure. Engng Fract
Mech 2017;180:148–60. https://doi.org/10.1016/j.engfracmech.2017.05.033.
[85] Manoharan M, Hirth J, Rosenfield A. A suggested procedure for combined mode I — mode III fracture toughness testing. J Test Eval 1990;18(2):106–14.
https://doi.org/10.1520/JTE12460J.
[86] Kamat SV, Srinivas M, Rama Rao P. Mixed mode I/III fracture toughness of Armco iron. Acta Mater 1998;46(14):4985–92. https://doi.org/10.1016/S1359-
6454(98)00170-0.
[87] Donne C, Pirondi A. J-integral evaluation of single-edge notched specimens under mixed-mode I/II loading. J Test Eval 2001;29(3):239–45. https://doi.org/10.
1520/JTE12251J.
[88] Gordon JA, Hirth JP, Kumar AM, Moody NE. Effects of hydrogen on the mixed mode I/III toughness of a high-purity rotor steel. Metall Trans A
1992;23(3):1013. https://doi.org/10.1007/BF02675576.
[89] Srinivas M, Kamat SV, Rao PR. Influence of mixed mode I/III loading on fracture toughness of mild steel at various strain rates. Mater Sci Technol
2004;20(2):235–42. https://doi.org/10.1179/026708304225011955.
[90] Kamat S, Srinivas M, Rama Rao P. On measuring mixed mode I/III fracture toughness of ductile materials using the critical stretch zone width. J Test Eval
2006;34(2):111–20. https://doi.org/10.1520/JTE12586.
[91] Yoda M. The effect of the notch root radius on the J-integral fracture toughness under modes I, II and III loadings. Engng Fract Mech 1987;26(3):425–31.
https://doi.org/10.1016/0013-7944(87)90023-3.
[92] Jen WK, Lin HC, Itua K. Calculation of stress intensity factors for combined mode bend specimens. ICF4, Waterloo, Canada1977.
[93] Shahani AR, Tabatabaei SA. Computation of mixed mode stress intensity factors in a four-point bend specimen. Appl Math Model 2008;32(7):1281–8. https://
doi.org/10.1016/j.apm.2007.04.001.
[94] Tada H, Paris PC, Irwin GR, Tada H. The stress analysis of cracks handbook. ASME press New York; 2000.
[95] Al Laham S. Stress intensity factor and limit load handbook. British Energy Generation Limited 1998.
[96] Murakami Y, Keer LM. Stress intensity factors handbook, Vol. 3. J Appl Mech 1993;60(4):1063.
[97] Liu S, Chao YJ, Zhu X. Tensile-shear transition in mixed mode I/III fracture. Int J Solids Struct 2004;41(22):6147–72. https://doi.org/10.1016/j.ijsolstr.2004.
04.044.
[98] Richard HA. A new compact shear specimen. Int J Fract 1981;17(5):R105–7. https://doi.org/10.1007/BF00033347.
[99] Pirondi A, Dalle Donne C. Mixed Mode Fracture of a Ferritic Steel: J-Integral against CTOD. The 5th International Conference on Biaxial/Multiaxial Fatigue and
Fracture1997. p. 555-76.
[100] Shih CF, Moran B, Nakamura T. Energy release rate along a three-dimensional crack front in a thermally stressed body. Int J Fract 1986;30(2):79–102. https://
doi.org/10.1007/BF00034019.
[101] Banks-Sills L, Sherman D. JII fracture testing of a plastically deforming material. Int J Fract 1991;50(1):15–26. https://doi.org/10.1007/BF00035166.
[102] Yusof F. Three-dimensional assessments of crack tip constraint. Theor Appl Fract Mech 2019;101:1–16. https://doi.org/10.1016/j.tafmec.2019.01.025.
[103] Xu L, Zhang X, Zhao L, Han Y, Jing H. Quantifying the creep crack-tip constraint effects using a load-independent constraint parameter Q*. Int J Mech Sci
2016;119:320–32. https://doi.org/10.1016/j.ijmecsci.2016.11.002.
[104] Betegón C, Hancock JW. Two-Parameter Characterization of Elastic-Plastic Crack-Tip Fields. Journal of Applied Mechanics 1991;58(1):104-10. 10.1115/1.
2897135.
[105] Du ZZ, Hancock JW. The effect of non-singular stresses on crack-tip constraint. J Mech Phys Solids 1991;39(4):555–67. https://doi.org/10.1016/0022-
5096(91)90041-L.
[106] Arun Roy Y, Narasimhan R. J-dominance in mixed mode ductile fracture specimens. Int J Fract 1997;88(3):259–79. https://doi.org/10.1023/
A:1007475708898.
[107] Oskui Ah, Soltani N, Rajabi M, Schmauder S. Mixed-mode fracture behavior of AM60 magnesium alloy using two parameter fracture mechanics. Engng Fract
Mech 2019;218:106566https://doi.org/10.1016/j.engfracmech.2019.106566.
[108] O'Dowd NP, Shih CF. Family of crack-tip fields characterized by a triaxiality parameter—II. Fracture applications. J Mech Phys Solids 1992;40(5):939–63.
https://doi.org/10.1016/0022-5096(92)90057-9.
[109] Demir O, Siriç S, Ayhan AO, Lekesiz H. Investigation of mixed mode - I/II fracture problems - Part 1: computational and experimental analyses. Frattura ed
Integrità Strutturale 2015;10(35):pages 330-9. 10.3221/IGF-ESIS.35.38.
[110] Smith DJ, Ayatollahi MR, Pavier MJ. The role of T-stress in brittle fracture for linear elastic materials under mixed-mode loading. Fatigue Fract Engng Mater
Struct 2001;24(2):137–50. https://doi.org/10.1046/j.1460-2695.2001.00377.x.
[111] Saghafi H, Ayatollahi MR, Sistaninia M. A modified MTS criterion (MMTS) for mixed-mode fracture toughness assessment of brittle materials. Mater Sci Engng,
A 2010;527(21):5624–30. https://doi.org/10.1016/j.msea.2010.05.014.
[112] Razavi SMJ, Berto F. A new fixture for fracture tests under mixed mode I/II/III loading. 2019;42(9):1874–88. https://doi.org/10.1111/ffe.13033.

24
Y. Wang, et al. Engineering Fracture Mechanics 235 (2020) 107126

[113] Koo JM, Choy YS. A new mixed mode fracture criterion: maximum tangential strain energy density criterion. Engng Fract Mech 1991;39(3):443–9. https://doi.
org/10.1016/0013-7944(91)90057-8.
[114] Kaung Jain C. On the maximum strain criterion—a new approach to the angled crack problem. Engng Fract Mech 1981;14(1):107–24. https://doi.org/10.
1016/0013-7944(81)90021-7.
[115] Williams JG, Ewing PD. Fracture under complex stress — The angled crack problem. Int J Fract 1984;26(4):346–51. https://doi.org/10.1007/BF00962967.
[116] Papadopoulos GA. The stationary value of the third stress invariant as a local fracture parameter (Det.-criterion). Engng Fract Mech 1987;27(6):643–52.
https://doi.org/10.1016/0013-7944(87)90156-1.
[117] Matvienko YG. Maximum average tangential stress criterion for prediction of the crack path. Int J Fract 2012;176(1):113–8. https://doi.org/10.1007/s10704-
012-9715-1.
[118] Sajjadi SH, Ostad Ahmad Ghorabi MJ, Salimi-Majd D. A novel mixed-mode brittle fracture criterion for crack growth path prediction under static and fatigue
loading. Fatigue Fract Engng Mater Struct 2015;38(11):1372–82. https://doi.org/10.1111/ffe.12320.
[119] Xu Z, Li Y. A novel method in determination of dynamic fracture toughness under mixed mode I/II impact loading. Int J Solids Struct 2012;49(2):366–76.
https://doi.org/10.1016/j.ijsolstr.2011.10.011.
[120] Miao X-T, Yu Q, Zhou C-Y, Li J, Wang Y-Z, He X-H. Experimental and numerical investigation on fracture behavior of I-II mixed mode crack for commercially
pure Titanium. Theor Appl Fract Mech 2018;96:202–15. https://doi.org/10.1016/j.tafmec.2018.04.012.
[121] Greer JM, Galyon Dorman SE, Hammond MJ. Some comments on the Arcan mixed-mode (I/II) test specimen. Engng Fract Mech 2011;78(9):2088–94. https://
doi.org/10.1016/j.engfracmech.2011.03.017.
[122] Huang Y, Huan L, Xiao L. Prediction of low cycle fatigue crack growth under mixed-mode loading conditions using cohesive zone models. Procedia Engng
2015;99:1317–22. https://doi.org/10.1016/j.proeng.2014.12.665.
[123] Demir O, Ayhan AO, İriç S. A new specimen for mixed mode-I/II fracture tests: modeling, experiments and criteria development. Engng Fract Mech
2017;178:457–76. https://doi.org/10.1016/j.engfracmech.2017.02.019.
[124] Schroth JG, Hoagland RG, Hirth JP, Rosenfield AR. Tensile and shear fracture of an HSLA steel. Scr Metall 1985;19(2):215–9. https://doi.org/10.1016/0036-
9748(85)90185-1.

25

You might also like