You are on page 1of 28

Review

2021, Vol. 50(8) 1165–1192


! The Author(s) 2019

Mode I fracture toughness Article reuse guidelines:


sagepub.com/journals-permissions
DOI: 10.1177/1528083719858767
of fiber-reinforced journals.sagepub.com/home/jit

polymer composites:
A review

Amna Siddique1 , Sharjeel Abid2, Faizan Shafiq1,


Yasir Nawab2, Hailou Wang1, Baohui Shi1, Sidra Saleemi3 and
Baozhong Sun1

Abstract
Composite materials are known for their high stiffness and strength at lower weight as
compared to conventional structural materials. Recently, there has been a growing
interest in finding the new ways to decrease delamination failure, which is a life limiting
factor of laminated composites. This review paper emphasizes on the effects of different
reinforcement structures on mode I fracture toughness and possible ways to improve
fracture toughness. A brief description on intrinsic and extrinsic mechanisms of crack
growth has been discussed along with the earlier investigations and recent develop-
ments for mode I fracture toughness testing. Factors that affect the fracture toughness
are also discussed. A brief knowledge of mode I fracture toughness of traditional and
advanced fiber-reinforced composites is given, which could help researchers to under-
stand fracture behaviors of composites and thus, it can help engineers to design com-
posites with higher interlaminar strength.

Keywords
Fiber-reinforced composites, fracture toughness, delamination, mode I, interlaminar,
2D composites, 3D composites

1
Donghua University, Shanghai, China
2
National Textile University, Faisalabad, Pakistan
3
University of the Punjab, Lahore, Pakistan

Corresponding author:
Amna Siddique, Donghua University, No.2999 North Renmin Rd., Songjiang District, Shanghai 201620, China.
Email: amnasiddique104@hotmail.com
1166 Journal of Industrial Textiles 50(8)

Introduction
The high specific modulus, also known as stiffness to weight ratio, of fiber-rein-
forced polymer composites (FRPCs) has resulted in the widespread use of these
composites as key structural engineering materials in automotive, aerospace [1–3],
marine, transportation, infrastructure, civil engineering applications [4–8] and
motorsport industries [4,9,10]. Since, there has been an ever-increasing demand
for composite materials and new materials are being introduced in the market
frequently, so it is challenging to assess their properties in order to forecast service
life and failure behaviors. The performance evaluation of the advanced reinforcing
fibers (glass [11], carbon [12,13], Kevlar [14,15] and PBO (poly(p-phenylene-2, 6-
benzobisoxazole) [16,17] fibers) and resins (Epoxy [18–21], vinyl ester [11], poly-
imide (PI) [22] and cyanate ester [23]) in final composite is necessary for their safe
application.
Compared to other damage categories, delamination evolution is a dominant
life-limiting failure mode for composite structures [13,24–28]. It may be introduced
either during manufacturing or caused by damaging events during service, for
example, impact damage [29,30], large hailstones [31,32] and bird strike events
[33,34], thus, hampering structural integrity and durability [35,36]. Keeping in
view the adverse impact of delamination on the composite structures, it is necessary
to consider in structural design process as well as in verification testing, to ensure
safe application with prolonged life of these composites.
Property which defines the ability of a material to resist fracture is known as
fracture toughness [37,38]. It indicates the amount of stress required to propagate a
pre-existing thin crack. Damage tolerance is the desired basic property for various
structures depending upon the end application [39]. Durability is an economic life-
cycle design concern, whereas, damage tolerance tells about the structure’s ability to
securely withstand flaws until these flaws are removed or fixed. Fracture toughness
characterization of composites is still on the way of growth as compared to metals.
Fracture toughness depends on many factors. Many researchers reported that in
FRPCs, the performance of fibers depends on the reinforcement’s architecture
[13,28,40], which is discussed in the next sections. Hence, structural effects of
reinforcement as well as effects of other factors (manufacturing process, loading
type, loading rate and temperature) on interlaminar fracture toughness are worth
studying. Furthermore, the knowledge of damage and its mechanism should be
extended to design composites with higher interlaminar strength for longer life span.

Fracture toughness
Fracture toughness is a distinguished property for many structural designs and
applications to ensure reliability [41,42]. The value of this property is typically
denoted by the strain energy release rate (GC ).
In composites, strain energy release rate (GC ) denotes the material resistance to
interlaminar fracture. The value of strain energy release rate is denoted as GIC , GIIC ,
Siddique et al. 1167

Figure 1. Various cracking modes for fracture toughness classification.

or GIIIC with respect to cracking mode being tested (Figure 1). The GC value at which
the delamination essentially starts to spread differs largely depending on the mode of
loading [5]. As the material is being tested and the crack begins to propagate, the
stiffness and force on the material begin to decrease. The decrease in the load means
that the strain energy stored in the material is also reducing or being released.

Modes of fracture
The manner in which a crack propagates through a material gives insight into the
mode of fracture. Thus, there are different modes of fracture for these composite
materials. Cracking mode I is considered to be the crack opening or tensile mode of
delamination. It is the most common form of fracture failure as its motion is like
pulling plies of material away from each other. The crack faces undergo opening
displacements relative to one another as it propagates. The corresponding material
property GIC (fracture toughness), under mode I loading, is commonly obtained by
the double cantilever beam (DCB) test method [43].
Cracking mode II is the in-plane shear mode of delamination. A shear stress acts
parallel to the plane of the crack and perpendicular to the crack front. This is
classified by two separated plies of material sliding above each other in the path
of crack growth. This cracking mode is less common than the crack opening tensile
mode, but it is still relevant to designs where force is not particularly down the
center of a structural component. In cracking mode III, known as tearing mode, a
shear stress acts parallel to the plane of the crack and parallel to the crack front.
This is also termed as out of plane shear mode.

Types of preform used in textile-reinforced polymer


composites and their impact on fracture toughness
The high specific stiffness and strength of composites have been of great deal for
their utilization in structures where higher weight is a life-threatening variable.
Hence, the development of low weight yet high strength composites are used in
aerospace [44,45] and defense fields [46]. The advanced textile fibers such as
1168 Journal of Industrial Textiles 50(8)

Dyneema, Kevlar, glass, carbon and Zylon are known for their outstanding per-
formance in protective applications. The performance of textile fibers varies
depending on their architecture when used as reinforcement in composites.
Hence, mechanical characterization in different structural forms [13,47–49],
under several static and dynamic loading situations, is the area of interest in
recent years.

Laminated composites
Most of the FRPCs are made from laminates. Interlaminar delamination presents
one of the most important life limiting failure mode that restricts their applications
[5,8–10,28,50,51] especially in primary aircraft structures [4,52]. As reported by
Tamuzs et al. [53], the critical energy release rate of laminated composites does
not usually exceed from 0.2 to 0.4 kJ/m2. These composite materials show a worry-
ing weakness towards the presence and development of cracks between the layers
[10] especially in mode I loading. [40]
Delamination can be initiated due to the occurrence of flaws during fabrication
process or produced in structural components during service due to interlaminar
tension and shear. It may also develop due to a multiple factors such as impact of
falling objects, structural discontinuities, free edge effects, differences in moisture
and temperature [7,10]. The growth of delamination results in stiffness loss and
could ultimately result in catastrophic failure [50,54,55]. Crack growth is a result of
competition among micro structural intrinsic and extrinsic mechanisms. Intrinsic
mechanisms perform in front of crack tip and promote crack evolution. Extrinsic
mechanisms perform behind crack tip and impede crack evolution [56].
In fibre-reinforced composites, the intrinsic mechanisms are structural voids
[57], material dislocation [58] or debonding [59,60] generated during manufacture
or in-service. The shielding mechanism, also known as external mechanism, results
from the closure traction between crack faces in the crack opening due to unbroken
fibres or yarns and by deflection of crack path [13,61]. Extrinsic mechanisms give
rise to resistive curve (R-curve) behaviour and thus, increase the force required for
crack growth [13,28,62]. In this regard, various methods have been introduced to
improve interlaminar strength [63]. Z-pinning [64–68], stitching [39,68–70], 3D
weaving [71] and braiding [9,68] are popular methods to improve fracture tough-
ness of textile composites. Advanced textile composites prepared by these technol-
ogies have shown good impact damage tolerance [72–74].

Stitched laminated composites


Some experimental studies have been conducted on stitched unidirectional or woven
FRPC composites to find interlaminar delamination. It was found that the stitched
FRPC laminates present significantly larger interlaminar toughness as compared to
un-stitched laminates [4,74–79], and that an increase in stitch density caused an
increase in mode I strain energy release rate [4,74–76]. A recent study [80] has been
Siddique et al. 1169

carried out on flax fiber-epoxy composite laminates to find the effect of stitching in-
through-thickness direction on mode I fracture toughness. At least 10% improve-
ment in fracture toughness at the lowest fraction of stitch fiber was reported.
But, it is problematic to stitch big and complex structures by using existing
stitching machines. It requires automated, intricate multi-needle stitching
machines, which are costly for composite manufacturers. Meanwhile, the in-
plane yarns can be distorted or damaged because of stitching leaving a negative
impact on the in-plane mechanical properties [4,75].

Knitted fabric composites


Mode I fracture studies performed on knitted composites reported that these com-
posites have extraordinary fracture toughness properties, being their GIC value 2 to
10 times greater than traditional composites [74,81,82]. The origin of the high GIC
values, which are superior to woven or UD laminates, lies in the very complex
fabric architecture [82]. Three-dimensional loop structure induces various energy-
consuming mechanisms such as crack branching, friction, yarn bridging and break-
age, which increase their fracture resistance.

Z-anchor-reinforced composites
Z-anchor (z-pinning) is a novel through-the-thickness reinforcement, which is used
to improve the interlaminar strength of composite laminates [83–88]. In the z-
anchor process, in-plane yarns are entangled each other by sticking with special
needles as shown in Figure 2. Z-pinned composites have significantly higher mode I
delamination resistance than two-dimensional (2D) equivalents [9,88–91]. Kusaka
et al. [88] experimentally characterized the mode I interlaminar fracture behavior of
z-anchor-reinforced composite laminates with different z-anchor densities. Mode I
interlaminar fracture behavior of z-anchor-reinforced composite laminates
increased almost linearly with the z-anchor density. Pingkarawat et al. [92] studied

Figure 2. Z-pinned reinforced composite [87].


1170 Journal of Industrial Textiles 50(8)

the effects of length, diameter and the volume content of z-pins on interlaminar
fracture toughness. It was reported that fracture toughness increased with an
increase in volume content and length of z-pins and by decrease in diameter.
During crack propagation process, z-anchor reinforcement induced a large
amount of fibers bridging resulting in the improvement of the mode I fracture
toughness of composite laminate [88,92].

Interlaminar z-directional flock-fibers-reinforced composites


Another technique of manufacturing composites with greater delamination resist-
ance is to introduce interlaminar z-directional flock-fibers [93]. Jia et al. carried out
a study to compare the mechanical properties, specifically, the mode I fracture
toughness of z-directional flock fiber-reinforced composite with 3D braided com-
posites. It was reported that the GIC value of 3D braided composite was higher
than z-directional flock fiber-reinforced composite due to difference in toughening
mechanisms as shown in Figure 3. But, 3D braiding has low productivity as com-
pared to flocking technology. Moreover, z-reinforced flocking composite technol-
ogy is found to be more cost effective [50].

Chemical methods to improve fracture toughness of laminated


composites
A common technique that has been adopted to improve the fracture toughness of
laminated composites is to use toughened matrix material [39,94–97]. Later it was
found that at higher temperatures, these resins show instability [50].
Another method to increase the interlaminar toughness of laminated composites
is the addition of whiskers. In this regard, Yamashita et al. [98] and Sohn et al. [99]
conducted experimental tests and found significant improvement in mode I fracture
toughness of the composites. Wang et al. [4] also reported a significant improve-
ment in mode I and mode II fracture toughness of CFRP laminates by dispensing
whiskers at interface during the lay-up procedure. However, it was found that
complications in the fabrication process and high cost are major hurdles in the
practical application of this method [4].

Figure 3. (a) 3D braiding composite (b) z-directional flocked fibers-reinforced composite.


Siddique et al. 1171

Each of the above-mentioned methods proposed a way that can result in


improved interlaminar fracture toughness and damage resistance. However,
three-dimensional (3D) textile technologies are supposed to be the most effective
ones. Numerous varieties of 3D fabrics have been developed which contribute
distinctive features [5,9,13,100].

3D Woven textile composites


3D woven composites offer two major benefits compared to the 2D conventional
laminated composites. Firstly, by altering warp and weft yarns, weave design,
amount of binder, their mechanical properties can be customized to fulfill particular
requirements without laminating [101,102]. Secondly and possibly more essentially,
the z-yarns in through-thickness direction are interweaved with warp and weft yarns
which prevent the growth of delamination [6] and increase post-impact mechanical
properties [103]. In consideration of the angle of binder yarns, angle-interlock and
orthogonal weaves can be identified as shown in Figure 4. In angle-interlock weave,
binder tows are placed at an angle of h < 90 , which is called undulation angle. In
orthogonal-interlock weaves, the tows are placed at h ¼ 90 . When z-yarns are lim-
ited to an adjacent layer or a few layers, we call it layer-to-layer (or ply-to-ply)
interlock weave [13], and when they go all the way through-the-thickness and inter-
lace with all the layers, we call it through-thickness weave [105].
3D woven fabric composites have the ability to provide superior interlaminar
fracture toughness and greater impact damage resistance than 2D conventional

Figure 4. 3D woven fabrics [104] (a) Orthogonal woven fabric. (b) Layer by layer angle inter-
lock. (c) Through thickness angle interlock.
1172 Journal of Industrial Textiles 50(8)

laminated composites [28,72,101,102]. Other desirable attributes include


weight reduction, improved fatigue and fracture resistance and fully integrated
assemblies.
3D woven composites are used for highly engineered applications such as
I-Beams for civil infrastructure [106], multi-hit composite armor systems [107],
aircraft frame work [105], leading edges of air craft wings [106] and engine fan
blades of aircrafts [105,108]. For these applications, high through-thickness mech-
anical properties and damage tolerance are inevitable. Numerous experimental
investigations on interlaminar toughness have been reported for 3D-reinforced
composites [5,6,100]. Enhanced damage tolerance is attained due to bridging for-
mation by z-binders in the crack wake. It impedes the growth of impact created
damage, and thus minimizes the depletion of the post-impact properties [109]. It
was found that a small volume content of binder yarns results in a big improvement
to the damage tolerance [109,110].

3D Orthogonal inter-lock woven composites


Watanabe et al. [5] studied the interlaminar fracture toughness of 3D orthogonal
interlocked fabric composites. They developed a numerical model for the charac-
terization of delamination. The influence of through-thickness yarns on delamin-
ation growth was quantified using that model. Guénon [100] carried out a study on
3D carbon/epoxy orthogonal fabric composite to find the effect of 3D fabric
geometery on interlaminr and in-plane fractuer behaviour of composite. The inter-
laminar fracture toughness results of 3D-reinforced composite were compared with
2D laminates having a compareable in-plane stacking arrangement and material
system. A 10 times gain of interlaminar toughness and 25% increase in the in-plane
toughness were attributed to the through-the-thickness fibers. Energy dissipitating
processes associated to the existence of z-direction fibers were: fracture debonding,
crack brancing and deviating, crack bridging caused by in-plane fibers and pull out
of the z-direction fibers.
Iwahori et al. [6] developed a mechanical model of DCB for 3D orthogonal
interlock fiber composite using a 2D finite element method (FEM). It was to simu-
late the DCB test results. They found a good agreement between simulation and
experimental test results. It was verified that the GIC usually achieved from DCB
tests can be assessed by interlaminar tension test and 2D FEM simulation. They
also reported the strong influence of the z-fibers on GIC .

3D angle interlock woven fabric composites


Among different 3D woven fabric composites, interlock fabric composites provide
a fibre architecture intended to improve interlaminar toughness, while, maintaining
in-plane performance by increasing only a small volume of through thickness
reinforcement [9,72,102,111,112]. According to Siddique et al. [13,28], fracture
toughness of 3D angle interlocked composites is supplemented by
Siddique et al. 1173

energy-absorbing mechanisms, such as matrix breakage, yarns bridging and pull


out and occurrences of secondary cracks as well as their propagation.

Comparisons of mode 1 fracture toughness properties


of traditional composites and advanced textile composites
The comparison of mode I interlaminar toughness of traditional composites made
from unidirectional or woven roving and composites manufactured from innova-
tive textile technologies of stitching, knitting, braiding, and through-the-thickness
weaving is given by Table 1. Their GIC values are at least twice as high as traditional
composites. The magnitude of the improvement in GIC depends upon the textile
processing factors, such as density and thickness of the through-thickness stitch/
binder, the knit yarn density and braiding angle [74]. The toughening of knitted
and braided composites is attributed to widespread crack splitting combined with
the zig-zag delamination path. The better interlaminar toughness in the stitched
and through-the thickness woven composites is attributed to crack bridging,
though crack splitting and stitch/binder pull out effects [74]. Figure 5 illustrates
the dominating mode I fracture mechanisms of traditional and advanced textile
composites.

Table 1. Overview of the mode I fracture toughness of composites.

Fiber/matrix Weave GIC (J/m2) References

Carbon/epoxy Unidirectional 629 [113]


Carbon/epoxy Unidirectional 820 [114]
Carbon/epoxy Unidirectional 751 [115]
Cross ply [0 /90 ] 1277
Carbon/epoxy UD tape 280 [116]
satin fabric 780
Carbon/epoxy UD woven fabric 890 [113]
Carbon/epoxy UD laminate 300 [118]
Carbon/epoxy [0 /90 ] 1295 [119]
Carbon/epoxy [0 /90 ] 282 [120]
E-glass/vinyl ester Plain woven 450 [11]
Carbon/epoxy Plain woven 508 [45]
Carbon/epoxy Laminate 216,000 [121]
Carbon/epoxy Multidirectional laminates [0 /5 ] 825 [122]
[45 /þ45 ] 835
[90 /þ90 ] 1000
Carbon/epoxy Woven laminate [0 ] 728 [123]
Carbon/epoxy 3D Angle interlock 704 [9]
3D Orthogonal 1569
1174 Journal of Industrial Textiles 50(8)

Figure 5. Fracture mechanisms (a) Plies separation (b) matrix cracking (c) yarns bridging
(d) crack deflection (e) secondary crack formation.

The analysis of the previous mode I interlaminar fracture studies shows that
advanced textile manufacturing methods are proficient to fabricate composites with
significantly increased delamination resistance.

Testing of fracture toughness


Delamination is a crucial failure mode of fiber-reinforced composite structures. For
the evaluation of delamination resistance, researchers have made attempts, which
resulted in the improvement of various test methods [124].
ASTM 5528 test is recommended for mode I fracture toughness characterization
(GIC ) of laminated reinforced polymer composites, using the DCB specimen
[10,125].For mode II fracture toughness (GIIC), the end notch flexure test is com-
monly used [126,127]. For pure mode III fracture toughness (GIIIC), the edge crack
torsion test was proposed by Ratcliffe [128]. ASTM D6671 recommends mixed-
mode bending test to measure the fracture toughness under various mode I and
mode II loading combinations [127]. There is still a lack of standard test methods to
measure fracture toughness under variety of loading conditions and materials.
Most of the mode I fracture tests use a split cantilever beam specimen as shown
in Figure 6(a). However, in case of materials with very high toughness (e.g. due to
through-the-thickness reinforcement) or with very low thickness, the failure of the
specimen arms has been observed prior delamination growth, thus, making
a delamination fracture toughness measurement almost impossible [124]. The
Siddique et al. 1175

Figure 6. Geometery of specimens for mode I testing (a) Double cantilever beam [145].
(b) Single edge notch beam tension specimen. (c) Compact tension specimen [142].

failure of beam arms normally starts on the outer surface of the specimen due to
the concentration of compressive stress [124]. Normally, the fiber-reinforced com-
posites have lower strength under compression as compared to tension, and the
lowest compressive strength is particularly observed at the outer surface of the
material where the fibers have less support [124]. Hence, it leads to failure of the
specimen arm at surface, before delamination extension occurs [70,124]. In this
situation, tabbed DCB specimen is better choice.

Types of specimens
Single edge notch tension specimen
The geometry of single edge notch tension specimen for Mode I tests [129] is shown
in Figure 6(b). It has been used under static [129–131] and impact [73] loading
1176 Journal of Industrial Textiles 50(8)

conditions. This type of specimen offers lower stresses in front of crack tip, and
thus results in self similar crack propagation and lower failure loads [129,132].

Compact tension specimen


Another specimen type used for mode I testing is compact tension specimen
[133–144]. Figure 6(c) shows configuration of compact tension specimen. It is pre-
ferred for materials having limited orthotropic behavior such as: particulates, short
fibers, nanoparticles and whiskers-containing composites. Moreover, it is used to
get fracture toughness just at damage initiation because of its limited ligament
length to find fracture propagation. Rikards [125] used a modified compact tension
specimen for fracture toughness testing of glass-epoxy-laminated composites.

DCB specimen
DCB is an ideal specimen type in mode I interlaminar fracture tests
[4,9,41,45,114,120,123,127,146–154]. It comprises of a rectangular composite spe-
cimen with uniform thickness as shown in Figure 6(a). A mechanical cut is
machined [71], or a non-adhesive Teflon film is inserted in the middle plane
during production, which acts as delamination starter. Two loading masses are
attached on bottom and top surfaces of the end of DCB specimen arms. The
delamination starter end of the DCB specimen is moved apart by quasi-static or
dynamic loading. During the test, the delamination lengths are recorded. For more
accurate delaminating length measurements, a travelling microscope is recom-
mended by ASTM [127]. DCB specimen is used for quasi-static tests and has an
advantage of moderate cost of manufacturing.

Wedge insert fracture specimen


Mode I testing using WIF type specimen is also an accepted method [155,156]. WIF
is the modified form of DCB specimen and is mostly used for the dynamic fracture
toughness tests. The only disadvantage is to measure the coefficient of friction
between composite surface and wedge [157].

General experimental procedures


Rectangular shape specimen for mode I test can be manufactured from different
types of materials and methods as mentioned above. An initial crack can be formed
using non-adhesive film [8,115,120,158] or mechanical cutting using saw [13,71].
Screw-driven machines [117], servo-hydraulic [8,159] or tensile testing machines
[115,158,160] can be used to perform quasi-static test at constant cross-head dis-
placements [161]. According to the aim of investigation, the test materials and test
conditions can be changed. During test, crack length can be found visually or by
using travelling microscope that help to see crack growth path [117]. Fracture
Siddique et al. 1177

process can be recorded using camera to analysis crack initiation and propagation.
Fracture toughness calculations can be done using schemes mentioned in the next
section (Data reduction methods). Scanning electron micro-scope images can be
used to determine the micro damages in the specimens [113,160,162].

Data reduction methods


To determine the mode I fracture toughness according to the ASTM D5528 stand-
ard, three methods are used [78]: the modified beam method (MBT) [163–165], the
compliance calibration (CC) method and the modified compliance calibration
(MCC) method. The difference between them by means of fracture results is not
more than 3.1%. Among these, the MBT gives the most conservative value of three
methods as reported by O’Brien and Martin [7].

. Modified beam theory

3P
GI ¼ ð1Þ
2bða þ Þ

where P and  represent the load applied and displacement. a, b and  repre-
sent crack length, specimen width and correction for crack length, respectively.
. CC

nP
GI ¼ ð2Þ
2ba
y
where n is the exponent form x , whereas, y and x are defined in Figure 7.
. MCC
3P2 C23
GI ¼ ð3Þ
2A1 bh
where A1 represents the slope of line and h is the thickness of specimen.

Figure 7. Graphs of data reduction theories. (a) Modified beam theory. (b) Compliance cali-
bration. (c) Modified compliance calibration.
1178 Journal of Industrial Textiles 50(8)

Factors effecting mode I fracture behaviour


Matrix toughness
Bradley [96] and Jordan et al. [97] found that toughness of matrix shows an import-
ant role in the interlaminar fracture behavior of composites. Decrease in the yield
strength of the matrix results an increase in the delamination fracture energy by
increasing the size of non-linear visco-elastic region or plastic deformation ahead
crack tip [13,166,167]. Finally, it results in higher load redistribution far from the
crack tip and therefore, more crack-tip blunting is seen. Crack tip blunting
increases the radius of crack tip [168,169], and ultimately increases the fracture
toughness by decreasing the stress intensity at crack tip. [13]

Weave type
Some investigations have been done to find the effect of weave pattern on fracture
toughness (GIC ) of laminated composites by mode I fracture toughness test, and it
showed some significant influence [11,170]. Suppakul and Bandyopadhyay [11]
compared the mode I fracture toughness of twill, satin and plain weave laminates
and reported that twill weave showed the highest GIC value followed by satin and
plain weave, respectively. Fibre bridging is reported as dominant mechanism which
contributed to the fracture resistance of twill and satin woven laminates, which was
absent in plain woven laminated composite. Fishpool et al. [9] compared the inter-
laminar fracture behaviour of three geometries of woven carbon fibre composites
layer to layer, orthogonal and angle interlock using DCB specimen, and found
significant effect of weave architecture. Orthogonal weave was found to be the most
effective in resisting delamination propagation, whereas, angle interlock weaves
were the least influential. In orthogonal weave, delamination resistance was con-
tributed by through-the-thickness fiber tows bridging, fibers and resin pull out
behind the crack front. In angle interlock, weave architecture as fiber tows was
staggered across the width of specimen, and thus, it resulted in twisting of the
specimen arms. Guénon [100] and Tanzawa et al. [5,61] also reported that in
orthogonal interlocked composites, many complicated mechanisms such as
debonding of z-fiber tows from in-plane layer, crack branching, pull-out of z-
fiber tows contribute for their improved delamination resistance.

Effect of manufacturing process


Composites manufactured by different moulding methods present different fracture
toughness values. Interlaminar fracture toughness can be affected by both compos-
ite manufacturing process and specimen thickness [171]. Specimens manufactured
by hand lay-up method showed decrease in GIC with an increase in specimen thick-
ness. Contrary to this, specimens manufactured by resin transfer moulding tech-
nique showed opposite trend [171]. Composites manufactured by resin transfer
moulding technique presented higher fracture resistance as compared to
Siddique et al. 1179

composites manufactured by hand lay-up. During testing, specimens undergo


mixed mode loading and twisting due to the lack of reinforcement symmetry
caused by lay-up moulding [172]. Carraro et al. [173] studied the effect of two
different resin infusion methods on mode I toughness and damage mechanisms
of carbon-epoxy laminates. Schober et al. [174] studied the effect of process par-
ameters (consolidation pressure and temperature) on glass fibre-reinforced thermo-
plastic tapes. Both studies [173,174] reported a negligible influence of
manufacturing process on mode I toughness. Kim and White [153] also reported
a negligible effect of staged curing cycles on mode I toughness of composite lamin-
ates. Contrary to this, Hunt et al. [154] reported cure path-dependent mode I
fracture toughness of particle interleaf laminated composites. Zhang and Fox
[175] reported the influence of manufacturing method (Quick step curing and auto-
clave curing) on mode I interlaminar toughness of carbon-epoxy laminated com-
posite. They reported that composites manufactured by quick step cure process
have greater mode I toughness as compared to those manufactured by autoclave
curing.

Specimen thickness
Mode I fracture characterization using DCB test exhibits large scale fibers bridging
phenomena (when bridging zone length becomes comparable to length to speci-
men) [176]. In this scenario, R-curve depends on specimen geometry. Thus, mode I
fracture toughness dependence on specimen thickness has gained particular interest
of many researchers [149,176–179]. Hojo and Aoki [180] investigated the effects of
DCB thickness for carbon-PEEK composites and measured initiation GIC values in
the range of 1100 to 1300 J m2, essentially, independent of the specimen thickness.
In contrast to the initiation values, the higher propagation GIC values were
reported with an increase in the specimen’s thickness. Manshadi et al. [179] and
Farmand et al. [149] also investigated the effect of specimen’s thickness on mode I
toughness in a glass fiber-polyester and carbon fiber-epoxy laminated composites.
There is consensus that an increase in specimen thickness results an increase in
fracture toughness, due to enhanced bridging zone length and the energy release
arte (ERR) at the plateau level by increasing of specimen scaling.

Loading rate
Hug et al. [181], Smiley and Pipes [182], and Gillespie et al. [183] conducted mode I
interlaminar fracture toughness tests on carbon-epoxy composites to find the influ-
ence of the loading rates on the interlaminar fracture toughness. They observed
stable crack propagation independent of testing rates with a little effect of rate on
GIC . Contrary to these, several works [126,184–190] report dependence of GIC on
loading rate. Kusaka et al. [155] observed that carbon-epoxy laminated composites
showed a negative rate-dependent fracture behavior (i.e. increasing loading rate
results decrease in fracture toughness). An increase in loading rate changed the
1180 Journal of Industrial Textiles 50(8)

crack propagation behavior from an unsteady manner to a steady continuous


growth, and consequently decreased GIC . Aliyu and Daniel [191] reported that in
graphite-epoxy composite, increasing the loading rate results the increment in frac-
ture toughness within a loading rate of 0.5 to 500 mm min1. In contrast, Smiley
and Pipes [192] and Gillespie et al. [183] observed that graphite-PEEK showed
negative rate-dependent fracture behavior at loading rate of 0.25–40,000 mm
min1. This was due to the transition of ductile to brittle crack growth behavior
with an increase in load rate. Mall et al. [192] also observed the negative rate-
dependent fracture response of carbon-PEEK woven composites with unstable
crack growth behavior at a range of loading rate from 0.005 to 1000 mm min1.
It was explained by the decrease in plastic deformation zone of matrix beneath
surface by increase in loading rate.

Temperature
Several researchers [193–196] reported that mode I fracture toughness increases
with temperature. This response has been attributed to the ductility of the
matrix at higher temperature. However, some researchers [197,198] reported a
complicated behavior of composites at different temperatures. Coronado and his
co-workers [199] also analyzed the effect of temperature on mode I delamination in
a carbon-epoxy unidirectional composite under static and fatigue loadings. During
static tests, the material performed better at room temperature. During fatigue
loading at 90oC, the ductility of the matrix increased, which enhanced resistance
to delamination during dynamic propagation [193]. They carried out another simi-
lar study at low temperature (60oC to 20oC) and found that toughness decreases
with decreasing test temperature [7].
Generally, the temperature-dependent fracture toughness of a composite is
attributed to mechanical properties of matrix, its structure and fiber-matrix inter-
facial strength. Thus, still there has been much dispute about the mechanisms of
temperature dependence of composites [200,201].

Summary and future directions


Summarizing the papers published on mode I facture toughness, significant pro-
gress on fracture characterization is noticed in the last decade. For laminated
composites, standard test methods, including specimen size and shape, are pub-
lished for fracture characterization but, for 3D-reinforced composites, researchers
are trying to find standards. Many articles have been published on the effect of
reinforcement geometry on fracture toughness and proposed ways to improve frac-
ture by reinforcement deign.
We discussed the two main classes of toughening mechanisms, intrinsic and
extrinsic, which are more and less related to the microstructure of materials.
Traditionally, a single value of GIC for crack initiation is considered as material
fracture toughness, which may or may not adequately capture the toughening of
Siddique et al. 1181

material as in case of multidirectional and 3D-reinforced composites where a prom-


inent R-curve behavior is observed. In order to improve fracture toughness, extrin-
sic shielding approach is found to be effective. This is attained by designing
microstructure in such a way that can offer a complex crack propagation path,
and thus can dissipate more strain energy and less crack growth. In this regard,
multidirectional laminated composites and 3D-reinforced composites are very
effective. Moreover, for fracture characterization of 3D-reinforced composites,
lack of standard test methods needs attention of researchers.
The various factors and mechanisms by which these influence the mode I frac-
ture toughness of FRPCs’ are reviewed. Loading rate, weave type and temperature
have dominating effect on toughness. The effect of loading rate and temperature on
mode I toughness is mainly attributed to the properties of matrix, its micro-struc-
ture and fiber–matrix interfacial strength. However, there are discrepancies on the
mechanisms of influence, that need further attention of researchers.

Declaration of conflicting interests


The author(s) declared no potential conflicts of interest with respect to the research, author-
ship, and/or publication of this article.

Funding
The author(s) disclosed receipt of the following financial support for the research, author-
ship, and/or publication of this article: The authors acknowledge the financial supports from
National Science Foundation of China (Grant Number 11572085 and 51675095). The
Fundamental Research Funds for the Central Universities of China (2232018G-02) are
also gratefully acknowledged.

ORCID iD
Amna Siddique https://orcid.org/0000-0001-6373-0076

References
[1] Couture A, Laliberte J and Li C. Mode I fracture toughness of aerospace polymer
composites exposed to fresh and salt water. Chem Mater Eng 2013; 1: 8–17.
[2] Döbrich O, Gereke T and Cherif C. Modeling the mechanical properties of textile-
reinforced composites with a near micro-scale approach. Compos Struct 2016; 135: 1–7.
[3] Elias A, Laurin F, Kaminski M, et al. Experimental and numerical investigations of low
energy/velocity impact damage generated in 3D woven composite with polymer matrix.
Compos Struct 2017; 159: 228–239.
[4] Wang WX, Takao Y, Matsubara T, et al. Improvement of the interlaminar fracture
toughness of composite laminates by whisker reinforced interlamination. Compos Sci
Technol 2002; 62: 767–774.
[5] Tanzawa Y, Watanabe N and Ishikawa T. FEM simulation of a modified DCB test for
3-D orthogonal interlocked fabric composites. Compos Sci Technol 2001; 61:
1097–1107.
1182 Journal of Industrial Textiles 50(8)

[6] Iwahori Y, Nakane K and Watanabe N. DCB test simulation of stitched CFRP lamin-
ates using interlaminar tension test results. Compos Sci Technol 2009; 69: 2315–2322.
[7] Coronado P, Arguelles A, Vina J, et al. Influence of low temperatures on the phenom-
enon of delamination of mode I fracture in carbon-fibre/epoxy composites under fati-
gue loading. Compos Struct 2014; 112: 188–193.
[8] Borowski E, Soliman E, Kandil U, et al. Interlaminar fracture toughness of CFRP
laminates incorporating multi-walled carbon nanotubes. Polymers 2015; 7: 1020–1045.
[9] Fishpool DT, Rezai A, Baker D, et al. Interlaminar toughness characterisation of 3D
woven carbon fibre composites. Plast Rubber Compos 2013; 42: 108–114.
[10] Arguelles A, Vina J, Canteli AF, et al. Interlaminar crack initiation and growth under
modes I and II in a carbon-fibre epoxy composite subjected to fatigue. WIT transactions
on the built environment. England, UK: WIT Press, 2008, pp.295–303.
[11] Suppakul P and Bandyopadhyay S. The effect of weave pattern on the mode-I inter-
laminar fracture energy of E-glass/vinyl ester composites. Compos Sci Technol 2002; 62:
709–717.
[12] Zhao H, Chen L, Yun J, et al. Improved thermal stabilities, ablation and mechanical
properties for carbon fibers/phenolic resins laminated composites modified by silicon-
containing polyborazine. Eng Sci 2018; 2: 57–66.
[13] Siddique A, Sun B and Gu B. Structural influences of two-dimensional and three-
dimensional carbon/epoxy composites on mode I fracture toughness behaviors with
rate effects on damage evolution. J Ind Text 2020; 50: 23–45. DOI: 10.1177/
1528083718819871.
[14] Kim S-C, Kim JS and Yoon H-J. Experimental and numerical investigations of mode I
delamination behaviors of woven fabric composites with carbon, Kevlar and their
hybrid fibers. Int Prec Eng Manuf 2011; 12: 321–329.
[15] Tang L, Dang J, He M, et al. Preparation and properties of cyanate-based wave-
transparent laminated composites reinforced by dopamine/POSS functionalized
Kevlar cloth. Compos Sci Technol 2019; 169: 120–126.
[16] Tang Y, Dong W, Tang L, et al. Fabrication and investigations on the polydopamine/
KH-560 functionalized PBO fibers/cyanate ester wave-transparent composites. Compos
Commun 2018; 8: 36–41.
[17] Gu J, Dong W, Xu S, et al. Development of wave-transparent, light-weight composites
combined with superior dielectric performance and desirable thermal stabilities.
Compos Sci Technol 2017; 144: 185–192.
[18] Huangfu Y, Liang C, Han Y, et al. Fabrication and investigation on the Fe3O4/ther-
mally annealed graphene aerogel/epoxy electromagnetic interference shielding nano-
composites. Compos Sci Technol 2019; 169: 70–75.
[19] Liang C, Qiu H, Han Y, et al. Superior electromagnetic interference shielding 3D
graphene nanoplatelets/reduced graphene oxide foam/epoxy nanocomposites with
high thermal conductivity. J Mater Chem C 2019; 7: 2725–2733.
[20] Gu J, Liang C, Zhao X, et al. Highly thermally conductive flame-retardant epoxy
nanocomposites with reduced ignitability and excellent electrical conductivities.
Compos Sci Technol 2017; 139: 83–89.
[21] Huangfu Y, Ruan K, Qiu H, et al. Fabrication and investigation on the PANI/
MWCNT/thermally annealed graphene aerogel/epoxy electromagnetic interference
shielding nanocomposites. Compos Part A 2019; 121: 265–272.
Siddique et al. 1183

[22] Gu J, Lv Z, Wu Y, et al. Dielectric thermally conductive boron nitride/polyimide


composites with outstanding thermal stabilities via in-situ polymerization-electrospin-
ning-hot press method. Compos Part A 2017; 94: 209–216.
[23] Li Y, Xu G, Guo Y, et al. Fabrication, proposed model and simulation predictions on
thermally conductive hybrid cyanate ester composites with boron nitride fillers. Compos
Part A 2018; 107: 570–578.
[24] Sebaey TA, Blanco N, Costa J, et al. Characterization of crack propagation in mode I
delamination of multidirectional CFRP laminates. Compos Sci Technol 2012; 72:
1251–1256.
[25] Tanzawa Y, Watanabe N and Ishikawa T. Interlaminar fracture toughness of 3-D
orthogonal interlocked fabric composites. Compos Sci Techn 1999; 59: 1261–1270.
[26] Bolotin VV. Delaminations in composite structures: its origin, buckling, growth and
stability. Compos Part B 1996; 27: 129–145.
[27] Raju IS and O’Brien TK. 1 – fracture mechanics concepts, stress fields, strain energy
release rates, delamination initiation and growth criteria. In: Sridharan S, (ed.)
Delamination behaviour of composites. Sawston, UK: Woodhead Publishing, 2008,
pp.3–27.
[28] Siddique A, Sun B and Gu B. Finite element modeling on fracture toughness of 3D
angle-interlock woven carbon/epoxy composites at microstructure level. Mech Adv
Mater Struct. Epub ahead of print 11 April 2019. DOI: 10.1080/
15376494.2019.1602235.
[29] Cantwell WJ and Morton J. The impact resistance of composite materials – a review.
Composites 1991; 22: 347–362.
[30] Heimbs S, Nogueira AC, Hombergsmeier E, et al. Failure behaviour of composite T-
joints with novel metallic arrow-pin reinforcement. Compos Struct 2014; 110: 16–28.
[31] Appleby-Thomas GJ, Hazell PJ and Dahini G. On the response of two commercially-
important CFRP structures to multiple ice impacts. Compos Struct 2011; 93:
2619–2627.
[32] Rhymer J, Kim H and Roach D. The damage resistance of quasi-isotropic carbon/
epoxy composite tape laminates impacted by high velocity ice. Compos Part A 2012; 43:
1134–1144.
[33] Heimbs S and Bergmann T. High-velocity impact behaviour of prestressed composite
plates under bird strike loading. Int J Aerosp Eng 2012; 2012: 11.
[34] Johnson AF and Holzapfel M. Influence of delamination on impact damage in com-
posite structures. Compos Sci Technol 2006; 66: 807–815.
[35] Jin L, Yao Y, Yu Y, et al. Structural effects of three-dimensional angle-interlock woven
composite undergoing bending cyclic loading. Sci China Phys Mech Astronom 2013; 57:
501–511.
[36] Selzer R and Krey J. Fractography of interlaminar fracture surfaces of CFPI and
CFBMI composites. J Mater Sci 1994; 29: 2951–2954.
[37] Zhu XK. Advances in fracture toughness test methods for ductile materials in low-
constraint conditions. Proc Eng 2015; 130: 784–802.
[38] Rubio C, Wang J, Martı́nez J, et al. Dynamic Fracture Toughness of Composite
Materials. In: Tamin M (ed.) Damage and Fracture of Composite Materials and
Structures. Advanced Structured Materials, vol 17. Berlin, Heidelberg, Germany:
Springer, 2012, pp. 143–156.
1184 Journal of Industrial Textiles 50(8)

[39] Zhang J, Deng S, Ye L, et al. Composites with matrices modified by nano-silica and
CTBN rubber. In: 13th international conference on fracture, Beijing, China, 16–21 June
2013, pp.1–7. USA: Curran Associates, Inc.
[40] Pappas G, Joncas S, Michaud V, et al. The influence of through-thickness reinforce-
ment geometry and pattern on delamination of fiber-reinforced composites: part I –
experimental results. Compos Struct 2018; 184: 924–934.
[41] Brunner AJ, Blackman BRK and Davies P. A status report on delamination resistance
testing of polymer–matrix composites. Eng Fract Mech 2008; 75: 2779–2794.
[42] Sato N, Hojo M and Nishikawa M. Novel test method for accurate characteriza-
tion of intralaminar fracture toughness in CFRP laminates. Compos Part B 2014; 65:
89–98.
[43] Hug G, Thévenet P, Fitoussi J, et al. Effect of the loading rate on mode I interlaminar
fracture toughness of laminated composites. Eng Fract Mech 2006; 73: 2456–2462.
[44] Oshima S, Yoshimura A, Hirano Y, et al. Experimental method for mode I fracture
toughness of composite laminates using wedge loaded double cantilever beam speci-
mens. Compos Part A 2018; 112: 119–125.
[45] Simon I, Banks-Sills L and Fourman V. Mode I delamination propagation and R-ratio
effects in woven composite DCB specimens for a multi-directional layup. Int J Fatigue
2017; 96: 237–251.
[46] Roylance D and Wang S-S. Influence of fibre properties on ballistic penetration of
textile panels. Fibre Sci Technol 1981; 14: 183–190.
[47] Behera BK and Dash BP. Mechanical behavior of 3D woven composites. Mater Des
2015; 67: 261–271.
[48] Wang X, Hu B, Feng Y, et al. Low velocity impact properties of 3D woven basalt/
aramid hybrid composites. Compos Sci Technol 2008; 68: 444–450.
[49] de Vasconcellos DS, Sarasini F, Touchard F, et al. Influence of low velocity impact on
fatigue behaviour of woven hemp fibre reinforced epoxy composites. Compos Part B
2014; 66: 46–57.
[50] Ren J, Kim YK and Rice J. Comparing the fracture toughness of 3-D braided preform
composites with z-fiber-reinforced laminar composites. Text Res J 2011; 81: 335–343.
[51] Davis DC and Whelan BD. An experimental study of interlaminar shear fracture
toughness of a nanotube reinforced composite. Compos Part B 2011; 42: 105–116.
[52] Paul D, Kelly L, Venkayya V, et al. Evolution of US military aircraft structures tech-
nology. J Aircraft 2002; 39: 18–29.
[53] Tamuzs V, Tarasovs S and Vilks U. Delamination properties of translaminar-rein-
forced composites. Compos Sci Technol 2003; 63: 1423–1431.
[54] Reedy ED, Mello FJ and Guess TR. Modeling the initiation and growth of delamin-
ations in composite structures. J Compos Mater 1997; 31: 812–831.
[55] Barikani M, Saidpour H and Sezen M. Mode-I interlaminar fracture toughness in
unidirectional carbon-fibre/epoxy composites. Iran Polym J 2002; 11: 413–23.
[56] Ritchie RO. Mechanisms of fatigue-crack propagation in ductile and brittle solids. Int J
Fract 1999; 100: 55–83.
[57] Gao X, Wang T and Kim J. On ductile fracture initiation toughness: effects of void
volume fraction, void shape and void distribution. Int J Solids Struct 2005; 42:
5097–5117.
[58] Kumar S and Curtin WA. Crack interaction with microstructure. Mater Today 2007;
10: 34–44.
Siddique et al. 1185

[59] Canal LP, González C, Segurado J, et al. Intraply fracture of fiber-reinforced compos-
ites: microscopic mechanisms and modeling. Compos Sci Technol 2012; 72: 1223–1232.
[60] Madhukar MS and Drzal LT. Fiber-matrix adhesion and its effect on composite mech-
anical properties: IV. Mode I and mode II fracture toughness of graphite/epoxy com-
posites. J Compos Mater 1992; 26: 936–968.
[61] Tanzawa W. Ishikawa Interlaminar fracture toughness of 3-D orthogonal interlocked
fabric composites. Compos Sci Technol 1999; 59: 1261–1270.
[62] Launey ME and Ritchie RO. On the fracture toughness of advanced materials. Adv
Mater 2009; 21: 2103–2110.
[63] Gillespie JJW. Damage tolerance of composite structures: the role of interlaminar frac-
ture mechanics. J Offshore Mech Arctic Eng 1991; 113: 247–252.
[64] Hong-Yuan L and Yiu-Wing M. Delamination fracture mechanics of composite lamin-
ates with through-thickness pinning. Strength Fract Complex 2003; 1: 139–146.
[65] Lenzi F, Riccio A, Clarke A, et al. Coupon tests on z-pinned and unpinned composite
samples for damage resistant applications. Macromol Sympos 2007; 247: 230–237.
[66] Byrd LW and Birman V. Effectiveness of z-pins in preventing delamination of co-cured
composite joints on the example of a double cantilever test. Compos Part B 2006; 37:
365–378.
[67] Liu HY and Mai YW. Effects of Z-pin reinforcement on interlaminar mode I delam-
ination. In: Proceedings of ICCM-13, Beijing, China, 25–29 June 2001.
[68] Dransfield K, Baillie C and Mai Y-W. Improving the delamination resistance of CFRP
by stitching – a review. Compos Sci Technol 1994; 50: 305–317.
[69] Shu DW and Mai YW. Effect of stitching on interlaminar delamination extension in
composite laminates. Compos Sci Technol 1993; 49: 165–171.
[70] Chen L, Ifju PG and Sankar BV. A novel double cantilever beam test for stitched
composite laminates. J Compos Mater 2001; 35: 1137–1149.
[71] Siddique A, Sun B and Gu B. Effect of pre-crack length on mode I fracture toughness
of 3-D angle-interlock woven composites from finite element analyses. J Ind Text 2019;
110: 1145–1458. DOI: 10.1080/00405000.2019.1613028.
[72] Ren C, Siddique A, Sun B, et al. Differences of transverse impact damages in 3D angle-
interlock woven composites between warp and weft directions. Int J Damage Mech
2019; 28: 1203–1227. DOI: 10.1177/1056789518823053.
[73] Shi B, Liu S, Siddique A, et al. Impact fracture behaviors of three-dimensional braided
composite U-notch beam subjected to three-point bending. Int J Damage Mech 2019;
28: 404–426.
[74] Mouritza AP, Bannisterb MK, Falzonb PJ, et al. Review of applications for advanced
three-dimensional fibre textile composites. Composites 1999; 30: 1445–1461.
[75] Mouritz AP and Cox BN. A mechanistic approach to the properties of stitched lamin-
ates. Composites Part A 2000; 31: 1–27.
[76] Yutaka Iwahori TI, Youichi H and Naoyuki W. Study of interlaminar fracture toughness
improvement on stitched CFRP laminates. J Jpn Soc Compos Mater 2000; 26: 90–100.
[77] Dransfield KA, Jain LK and Mai YW. On the effects of stitching in CFRPs – I. Mode I
delamination toughness. Compos Sci Technol 1998; 58: 815–827.
[78] Göktaş D, Kennon WR and Potluri P. Improvement of mode I interlaminar fracture
toughness of stitched glass/epoxy composites. Appl Compos Mater 2017; 24: 351–375.
[79] Jain LK and Mai Y-W. On the effect of stitching on mode I delamination toughness of
laminated composites. Compos Sci Technol 1994; 51: 331–345.
1186 Journal of Industrial Textiles 50(8)

[80] Ravandi M, Teo WS, Tran LQN, et al. The effects of through-the-thickness stitching
on theMode I interlaminar fracture toughness of flax/epoxy composite laminates.
Mater Des 2016; 109: 659–669.
[81] Kim K-Y and Ye L. Interlaminar fracture properties of weft-knitted/woven fabric
interply hybrid composite materials. J Mater Sci 2012; 47: 7280–7290.
[82] Falconnet D, Bourban PE, Pandita S, et al. Fracture toughness of weft-knitted fabric
composites. Compos Part B 2002; 33: 579–588.
[83] Kusaka T. 9 – toughening mechanisms in Zanchor-reinforced composites. In: Qin Q
and Ye J (eds) Toughening mechanisms in composite materials. Sawston, UK:
Woodhead Publishing, 2015, pp.235–261.
[84] Kusaka T, Watanabe K, Hojo M, et al. Fracture behavior and toughening mechanism
in Zanchor reinforced composites under mode II loading. Compos Sci Technol 2009;
69: 2323–2330.
[85] Partridge IK and Cartié DDR. Delamination resistant laminates by Z-FiberÕ pinning:
part I manufacture and fracture performance. Compos Part A 2005; 36: 55–64.
[86] Byrd LW and Birman V. The estimate of the effect of z-pins on the strain release rate,
fracture and fatigue in a composite co-cured z-pinned double cantilever beam. Compos
Struct 2005; 68: 53–63.
[87] Robinson P and Das S. Mode I DCB testing of composite laminates reinforced with
z-direction pins: a simple model for the investigation of data reduction strategies. Eng
Fract Mech 2004; 71: 2561.
[88] Kusaka T, Watanabe K, Hojo M, et al. Fracture behaviour and toughening mechan-
ism in Zanchor reinforced composites under mode I loading. Eng Fract Mech 2012; 96:
433–446.
[89] Pegorin F, Pingkarawat K and Mouritz AP. Comparative study of the mode I and
mode II delamination fatigue properties of z-pinned aircraft composites. Mater Design
2015; 65: 139–146.
[90] Pegorin F, Pingkarawat K, Daynes S, et al. Influence of z-pin length on the delamin-
ation fracture toughness and fatigue resistance of pinned composites. Compos Part B
2015; 78: 298–307.
[91] Yan W, Liu H-Y and Mai YW. Numerical study on the mode I delamination tough-
ness of Z-pinned laminates. Compos Sci Technol 2003; 63: 1481–1493.
[92] Pingkarawat K and Mouritz AP. Improving the mode I delamination fatigue resist-
ance of composites using z-pins. Compos Sci Technol 2014; 92: 70–76.
[93] Kim YK and Hoskote S. Fracture toughness of flock fiber reinforced layered com-
posites. In: 1st international industrial simulation conference, Valencia, Spain, 9–11
June 2003, pp.477–482.
[94] Arai M, Hirokawa J-i, Hanamura Y, et al. Characteristic of mode I fatigue crack
propagation of CFRP laminates toughened with CNF interlayer. Compos Part B 2014;
65: 26–33.
[95] Takeda N, Kobayashi S, Ogihara S, et al. Effects of toughened interlaminar layers on
fatigue damage progress in quasi-isotropic CFRP laminates. Int J Fatigue 1999; 21:
235–242.
[96] Bradley WL. Chapter 5 – relationship of matrix toughness to interlaminar fracture
toughness. In: Friedrich K (ed) Composite materials series. Amsterdam, Elsevier, 1989,
pp.159–187.
Siddique et al. 1187

[97] Jordan WM, Bradley WL and Moulton RJ. Relating resin mechanical properties to
composite delamination fracture toughness. J Compos Mater 1989; 23: 923–943.
[98] Yamashita S, Hatta H, Takei T, et al. Interlaminar reinforcement of laminated compos-
ites by addition of oriented whiskers in the matrix. J Compos Mater 1992; 26: 1254–1268.
[99] Sohn MS, Hu XZ, Kim JK, et al. Impact damage characterisation of carbon fibre/
epoxy composites with multi-layer reinforcement. Compos Part B 2000; 31: 681–691.
[100] Guénon CG. Toughness properties of a three-dimensional carbon-epoxy composite.
J Mater Sci 1989; 24: 4168–4175.
[101] Yu B, Bradley RS, Soutis C, et al. 2D and 3D imaging of fatigue failure mechanisms of
3D woven composites. Compos Part A 2015; 77: 37–49.
[102] Ren C, Liu T, Siddique A, et al. High-speed visualizing and mesoscale modeling for
deformation and damage of 3D angle-interlock woven composites subjected to trans-
verse impacts. Int J Mech Sci 2018; 140: 119–132.
[103] Jia X, Sun B and Gu B. Ballistic penetration of conically cylindrical steel projectile
into 3D orthogonal woven composite: a finite element study at microstructure level.
J Compos Mater 2011; 45: 965–987.
[104] Hallal A, Younes R, Fardoun F, et al. Improved analytical model to predict the
effective elastic properties of 2.5D interlock woven fabrics composite. Compos
Struct 2012; 94: 3009–3028.
[105] Huang T, Wang Y and Wang G. Review of the mechanical properties of a 3D woven
composite and its applications. Polym Plast Technol Eng 2018; 57: 740–756.
[106] Tong L, Mouritz AP and Bannister MK. Chapter 1 – introduction. In: Tong L,
Mouritz AP, Bannister MK (eds) 3D fibre reinforced polymer composites. Oxford:
Elsevier Science, 2002, pp. 1–12.
[107] Bogdanovich AE. Advancements in manufacturing and applications of 3-D woven
preforms and composites. In: ICCM international conferences on composite materials,
Kyoto, Japan, 21 June 2006.
[108] De Luycker E, Morestin F, Boisse P, et al. Simulation of 3D interlock composite
preforming. Compos Struct 2009; 88: 615–623.
[109] Rudov-Clark S and Mouritz AP. Tensile fatigue properties of a 3D orthogonal woven
composite. Compos Part A 2008; 39: 1018–1024.
[110] Davallo M. Factors affecting fracture behaviour of composite materials. Int J Chem
Tech Res 2010; 2: 2125–2130.
[111] Jin L, Jin BC, Kar N, et al. Tension–tension fatigue behavior of layer-to-layer 3-D
angle-interlock woven composites. Mater Chem Phys 2013; 140: 183–190.
[112] Wang M, Cao M, Wang H, et al. Drop-weight impact behaviors of 3-D angle interlock
woven composites after thermal oxidative aging. Compos Struct 2017; 166: 239–255.
[113] Rehan MSBM, Rousseau J, Fontaine S, et al. Experimental study of the influence of
ply orientation on DCB mode-I delamination behavior by using multidirectional fully
isotropic carbon/epoxy laminates. Compos Struct 2017; 161: 1–7.
[114] Zhao L, Wang Y, Zhang J, et al. XFEM-based model for simulating zigzag delamin-
ation growth in laminated composites under mode I loading. Compos Struct 2017; 160:
1155–1162.
[115] Farmand-Ashtiani E, Alanis D, et al. Delamination in cross-ply laminates: identifica-
tion of traction–separation relations and cohesive zone modeling. Compos Sci Technol
2015; 119: 85–92.
1188 Journal of Industrial Textiles 50(8)

[116] Fanteria D, Lazzeri L, Panettieri E, et al. Experimental characterization of the inter-


laminar fracture toughness of a woven and a unidirectional carbon/epoxy composite.
Compos Sci Technol 2017; 142: 20–29.
[117] Filatovs GJ, Sadler RL and Elshiekh AHM. Fracture-behavior of a 3-D braid gra-
phite-epoxy composite. J Compos Mater 1994; 28: 526–42.
[118] Canal LP, Pappas G and Botsis J. Large scale fiber bridging in mode I intralaminar
fracture. An embedded cell approach. Compos Sci Technol 2016; 126: 52–59.
[119] de Morais AB, de Moura MF, Marques AT, et al. Mode-I interlaminar fracture of
carbon/epoxy cross-ply composites. Compos Sci Technol 2002; 62: 679–686.
[120] Reis PNB, Ferreira JAM, Antunes FV, et al. Initial crack length on the interlaminar
fracture of woven carbon/epoxy laminates. Fibers Polym 2015; 16: 894–901.
[121] Bergan A, Dávila C, Leone F, et al. Mode I cohesive law characterization procedure
for through-the-thickness crack propagation in composite laminates. Compos Part B
2016; 94: 338–349.
[122] Zhao L, Gong Y, Zhang J, et al. Simulation of delamination growth in multidirec-
tional laminates under mode I and mixed mode I/II loadings using cohesive elements.
Compos Struct 2014; 116: 509–522.
[123] Navarro P, Aubry J, Pascal F, et al. Influence of the stacking sequence and crack
velocity on fracture toughness of woven composite laminates in mode I. Eng Fract
Mech 2014; 131: 340–348.
[124] Reeder JR, Demarco K and Whitley KS. The use of doubler reinforcement in delam-
ination toughness testing. Compos Part A 2004; 35: 1337–1344.
[125] de Morais AB and Pereira AB. Application of the effective crack method to mode I
and mode II interlaminar fracture of carbon/epoxy unidirectional laminates. Compos
Part A 2007; 38: 785–794.
[126] Sjogren A and Asp LE. Effects of temperature on delamination growth in a carbon/
epoxy composite under fatigue loading. Int J Fatigue 2002; 24: 179–184.
[127] Prasad MSS, Venkatesha CS and Jayaraju T. Experimental methods of determining
fracture toughness of fiber reinforced polymer composites under various loading con-
ditions. J Miner Mater Charact Eng 2011; 10: 1263–1275.
[128] Ratcliffe JG. Characterization of the edge crack torsion (ECT) test for mode III
fracture toughness measurement of laminated composites. NASA-TM-2004-213269,
2004.
[129] Rakshit D and Chakraborty S. Determination of fracture parameters of FRP com-
posites: a combined experimental and numerical investigation. J Compos Mater 2015;
49: 231–241.
[130] Alfaseeh M, Kueh A, Abo Sabah S, et al. Influence of tows waviness and anisotropy
on effective Mode I fracture toughness of triaxially woven fabric composites. Eng
Fract Mech 2017; 182: 521–536.
[131] Martin JT and Lambert SB. Analysis of constraint in single edge notch tension spe-
cimens using the T-stress. Int J Pressure Vessels Piping 1996; 65: 13–19.
[132] El-Hajjar R and Haj-Ali R. Mode-I fracture toughness testing of thick section FRP
composites using the ESE(T) specimen. Eng Fract Mech 2005; 72: 631–643.
[133] Aly-Hassan MS, Hatta H, Wakayama S, et al. Comparison of 2D and 3D carbon/
carbon composites with respect to damage and fracture resistance. Carbon 2003; 41:
1069–1078.
Siddique et al. 1189

[134] Li S, Thouless MD, Waas AM, et al. Use of a cohesive-zone model to analyze the
fracture of a fiber-reinforced polymer–matrix composite. Compos Sci Technol 2005;
65: 537–549.
[135] Kongshavn I and Poursartip A. Experimental investigation of a strain-softening
approach to predicting failure in notched fibre-reinforced composite laminates.
Compos Sci Technol 1999; 59: 29–40.
[136] Pinho ST, Robinson P and Iannucci L. Fracture toughness of the tensile and com-
pressive fibre failure modes in laminated composites. Compos Sci Technol 2006; 66:
2069–2079.
[137] Gonzáles L and Knauss WG. Scaling global fracture behavior of structures-sized
laminated composites. Int J Fract 2002; 118: 363–394.
[138] Catalanotti G, Camanho PP, Xavier J, et al. Measurement of resistance curves in the
longitudinal failure of composites using digital image correlation. Compos Sci Technol
2010; 70: 1986–1993.
[139] Laffan MJ, Pinho ST, Robinson P, et al. Translaminar fracture toughness: the critical
notch tip radius of 0 plies in CFRP. Compos Sci Technol 2011; 72: 97–102.
[140] Blanco N, Trias D, Pinho ST, et al. Intralaminar fracture toughness characterisation
of woven composite laminates. Part I: design and analysis of a compact tension (CT)
specimen. Eng Fract Mech 2014; 131: 349–360.
[141] Li X, Hallett SR, Wisnom MR, et al. Experimental study of damage propagation in
over-height compact tension tests. Compos Part A 2009; 40: 1891–1899.
[142] Farahani BV, Tavares PJ, Belinha J, et al. A fracture mechanics study of a compact
tension specimen: digital image correlation, finite element and meshless methods. In:
2nd international conference on structural integrity, Funchal, Madeira, Portugal, 4–7
September 2017, pp.920–927. Procedia Structural Integrity, Elsevier B.V.
[143] Yuan F-G and Yang S. Fracture behavior of stitched warp-knit fabric composites. Int
J Fract 2001; 108: 73–94.
[144] Kumar Naik P, Londe DN, Yogesha B, et al. Mode I fracture characterization of
banana fibre reinforced polymer composite. IOP Conf Series 2018; 376: 012041.
[145] ISO. Fiber-reinforecd plastic composites. Determination of mode I inter-laminar fracture
toughness, GIc, for unidirectionally reinforced materials. [ISO 15024: 2001], 2001.
[146] Davies P, Blackman BRK and Brunner AJ. Standard test methods for delamination
resistance of composite materials: current status. Appl Compos Mater 1998; 5: 345–364.
[147] Iwamoto M, Ni QQ, Fujiwara T, et al. Intralaminar fracture mechanism in unidirec-
tional CFRP composites – part II: analysis. Eng Fract Mech 1999; 64: 747–764.
[148] Nairn JA. Fracture-mechanics of unidirectional composites using the shear-lag model-
II – experiment. J Compos Mater 1988; 22: 589–600.
[149] Farmand-Ashtiani E, Cugnoni J and Botsis J. Specimen thickness dependence of large
scale fiber bridging in mode I interlaminar fracture of carbon epoxy composite. Int J
Solids Struct 2015; 55: 58–65.
[150] Manshadi BD, Farmand-Ashtiani E, Botsis J, et al. An iterative analytical/experimen-
tal study of bridging in delamination of the double cantilever beam specimen. Compos
Part A 2014; 61: 43–50.
[151] Jiang Z, Wan S, Zhong Z, et al. Determination of mode-I fracture toughness and non-
uniformity for GFRP double cantilever beam specimens with an adhesive layer. Eng
Fract Mech 2014; 128: 139–156.
1190 Journal of Industrial Textiles 50(8)

[152] Wang Y and Zhao D. Characterization of interlaminar fracture behaviour of woven


fabric reinforced polymeric composites. Composites 1995; 26: 115–124.
[153] White SR and Kim YK. Staged curing of composite materials. Compos Part A 1996;
27: 219–227.
[154] Hunt C, Kratz J and Partridge IK. Cure path dependency of mode I fracture tough-
ness in thermoplastic particle interleaf toughened prepreg laminates. Compos Part A
2016; 87: 109–14.
[155] Kusaka T, Hojo M, Mai Y-W, et al. Rate dependence of mode I fracture behaviour in
carbon-fibre/epoxy composite laminates. Compos Sci Technol 1998; 58: 591–602.
[156] Kusaka T. 107 interlaminar fracture behavior of polymer matrix composite materials
under impact loading. JSME Mater Process Conf 2002; 10: 52–7.
[157] May M. Measuring the rate-dependent mode I fracture toughness of composites – A
review. Compos Part A 2016; 81: 1–12.
[158] Goktas D, Kennon WR and Potluri P. Improvement of mode I interlaminar fracture
toughness of stitched Glass/Epoxy composites. Appl Compos Mater 2016; 24: 351–75.
[159] Wang L, Wu J, Chen C, et al. Progressive failure analysis of 2D woven composites at
the meso-micro scale. Compos Struct 2017; 178: 395–405.
[160] Cholake ST, Moran G, Joe B, et al. Improved mode I fracture resistance of CFRP
composites by reinforcing epoxy matrix with recycled short milled carbon fibre.
Construct Build Mater 2016; 111: 399–407.
[161] Goto K, Hatta H, Takahashi H, et al. Effect of shear damage on the fracture behavior
of carbon-carbon composites. J Am Ceram Soc 2001; 84: 1327–1333.
[162] Zabala H, Aretxabaleta L, Castillo G, et al. Loading rate dependency on mode I
interlaminar fracture toughness of unidirectional and woven carbon fibre epoxy com-
posites. Compos Struct 2015; 121: 75–82.
[163] Williams JG. End corrections for orthotropic DCB specimens. Compos Sci Technol
1989; 35: 367–376.
[164] Hashemi S, Kinloch AJ and Williams JG. Corrections needed in double-cantilever
beam tests for assessing the interlaminar failure of fibre-composites. J Mater Sci Lett
1989; 8: 125–129.
[165] Hashemi S, Kinloch AJ and Williams JG. Mechanics and mechanisms of delamination
in a poly(ether sulphone)-fibre composite. Compos Sci Technol 1990; 37: 429–462.
[166] Kishi H, Shi YB, Huang J, et al. Ductility and toughenability study of epoxy resins
under multiaxial stress states. J Mater Sci Lett 1998; 33: 3479–3488.
[167] Jordan WM and Bradley WL. The relationship between resin mechanical properties
and Mode I delamination fracture toughness. J Mater Sci Lett 1988; 7: 1362–1364.
[168] Salazar A, Patel Y and Williams JG. Influence of crack sharpness on the fracture
toughness of epoxy resins. In: 13th international conference on fracture 2013, Beijing,
China, 16–21 June 2013, pp.4057–4066. USA: Curran Associates, Inc.
[169] Chakachery E A and Bradley W L. A comparison of the crack tip damage zone for
fracture of Hexcel F185 neat resin and T6T145/F185 composite. Polym Eng Sci 1987;
27: 33–40.
[170] Nasuha N, Azmi A and Lih T. A review on mode-I interlaminar fracture toughness of
fibre reinforced composites. J Physics 2017; 908: 012024.
[171] Davallo M. Factors affecting fracture behaviour of composite materials. Int Chem
Tech Res 2010; 2: 2125–2130.
Siddique et al. 1191

[172] Davies P and Benzeggagh ML. Chapter 3 – interlaminar mode-I fracture testing.
In: Friedrich K (ed.) Composite materials series. Amsterdam: Elsevier, 1989,
pp. 81–112.
[173] Carraro PA, Maragoni L and Quaresimin M. Influence of manufacturing induced
defects on damage initiation and propagation in carbon/epoxy NCF laminates. Adv
Manuf 2015; 1: 44–53.
[174] Schober M, Kuboki T, Ameri E, et al. Effects of process parameters on the interla-
minar fracture toughness of GF-PA6-tapes. PAMM 2017; 17: 273–274.
[175] Zhang J and Fox BL. Manufacturing influence on the delamination fracture behavior
of the T800H/3900-2 carbon fiber reinforced polymer composites. Mater Manuf Proc
2007; 22: 768–772.
[176] Sørensen BF and Jacobsen TK. Large-scale bridging in composites: R-curves and
bridging laws. Compos Part A 1998; 29: 1443–1451.
[177] Davies P, Kausch HH, Williams JG, et al. Round-robin interlaminar fracture testing
of carbon-fibre-reinforced epoxy and PEEK composites. Compos Sci Technol 1992; 43:
129–136.
[178] Agrawal A and Ben Jar PY. Analysis of specimen thickness effect on interlaminar
fracture toughness of fibre composites using finite element models. Compos Sci
Technol 2003; 63: 1393–402.
[179] Manshadi BD, Vassilopoulos AP and Botsis J. A combined experimental/numerical
study of the scaling effects on mode I delamination of GFRP. Compos Sci Technol
2013; 83: 32–9.
[180] Hojo M and Aoki T. Thickness effect of double cantilever beam specimen on interla-
minar fracture toughness of AS4/PEEK and T800/epoxy laminates. ASTM STP 1156,
1993, p.281-98.
[181] Hug G, Thevenet P, Fitoussi J and Baptiste D. Effect of the loading rate on mode I
interlaminar fracture toughness of laminated composites. Engineering Fracture
Mechanics 2006; 73: 2456–62.
[182] Smiley AJ and Pipes RB. Rate effects on mode-I interlaminar fracture toughness in
composite materials. J Compos Mater 1987; 21: 670–687.
[183] Gillespie JW, Carlsson LA and Smiley AJ. Rate dependent Mode I interlaminar crack
growth mechanisms in graphite epoxy and graphite PEEK. Compos Sci Technol 1987;
28: 1–15.
[184] Ashcroft IA, Hughes DJ and Shaw SJ. Mode I fracture of epoxy bonded composite
joints: 1. Quasi-static loading. Int J Adhes Adhes 2001; 21: 87–99.
[185] Hojo M, Ando T, Tanaka M, et al. Modes I and II interlaminar fracture toughness
and fatigue delamination of CF/epoxy laminates with self-same epoxy interleaf. Int J
Fatigue 2006; 28: 1154–1165.
[186] Shivakumar K, Chen H, Abali F, et al. A total fatigue life model for mode I delami-
nated composite laminates. Int J Fatigue 2006; 28: 33–42.
[187] Mahdi S, Kim H-J, Gama BA, et al. A comparison of oven-cured and induction-cured
adhesively bonded composite joints. J Compos Mater 2003; 37: 519–542.
[188] Matsubara G, Ono H and Tanaka K. Mode II fatigue crack growth from delamin-
ation in unidirectional tape and satin-woven fabric laminates of high strength GFRP.
Int J Fatigue 2006; 28: 1177–1186.
1192 Journal of Industrial Textiles 50(8)

[189] Hansen U and Gillespie JW. Dependence of intralaminar fracture toughness on dir-
ection of crack propagation in unidirectional composites. J Compos Technol Res 1998;
20: 89–99.
[190] Brunner AJ. Experimental aspects of mode I and mode II fracture toughness testing of
fibre-reinforced polymer-matrix composites. Comput Meth Appl Mech Eng 2000; 185:
1612–172.
[191] Aliyu AA and Daniel IM. Effects of strain rate on delamination fracture toughness of
graphite/epoxy. West Conshohocken: ASTM Special Technical Publication, 1985,
pp.336–348.
[192] Mall S, Law GE and Katouzian M. Loading rate effect on interlaminar fracture
toughness of a thermoplastic composite. J Compos Mater 1987; 21: 569–579.
[193] Coronado P, Argüelles A, Viña J, et al. Influence of temperature on a carbon–fibre
epoxy composite subjected to static and fatigue loading under mode-I delamination.
Int J Solids Struct 2012; 49: 2934–2940.
[194] Hunston DL and Bascom WD. Effects of lay-up, temperature, and loading rate in
double cantilever beam tests of interlaminar crack growth. Compos Technol Rev 1983;
5: 118–119.
[195] Hine PJ, Brew B, Duckett RA, et al. Failure mechanisms in continuous carbon-fibre
reinforced PEEK composites. Compos Sci Technol 1989; 35: 31–51.
[196] Czabaj MW and Davidson BD. Determination of the mode I, mode II, and mixed-
mode I-II delamination toughness of a graphite/polyimide composite at room and
elevated temperatures. J Compos Mater 2016; 50: 2235–2253.
[197] Jiang B and Levi D. Spatial interval discrimination in two dimensions. Invest Ophthal
Visual Sci 1990; 31: 326.
[198] Hartwig G, Kneifel B and Poehlmann K. Fracture properties of polymers and compos-
ites at cryogenic temperatures. New York, NY, USA: Plenum Press, 1986, pp.169–177.
[199] Argüelles A, Viña J, Fernández-Canteli A, et al. Influence of the temperature in the
delamination under mode I of fracture and dynamic loading of two carbon-epoxy
composites. Compos Part B 2015; 68: 207–214.
[200] Xiaoping H, Shenliang H and Liang Y. A study on dynamic fracture toughness of
composite laminates at different temperatures. Compos Sci Technol 2003; 63: 155–159.
[201] Rikards R. Interlaminar fracture behaviour of laminated composites. Comput Struct
2000; 76: 11–18.

You might also like