Failure Mechanisms in FRC 2

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Downloaded from http://rsta.royalsocietypublishing.

org/ on June 4, 2016

Physical modelling of failure


in composites
rsta.royalsocietypublishing.org
Ramesh Talreja1,2,3
1 Department of Aerospace Engineering, and 2 Department of

Materials Science and Engineering, Texas A&M University,


Research College Station, TX 77843, USA
3 Department of Engineering Sciences and Mathematics,
Cite this article: Talreja R. 2016 Physical
modelling of failure in composites. Phil. Trans. Luleå University of Technology, 971 87 Luleå, Sweden
R. Soc. A 374: 20150280. RT, 0000-0003-1871-4020
http://dx.doi.org/10.1098/rsta.2015.0280
Structural integrity of composite materials is governed
by failure mechanisms that initiate at the scale of
Accepted: 4 March 2016
the microstructure. The local stress fields evolve with
the progression of the failure mechanisms. Within
One contribution of 22 to a Theo Murphy the full span from initiation to criticality of the
meeting issue ‘Multiscale modelling of the failure mechanisms, the governing length scales in
structural integrity of composite materials’. a fibre-reinforced composite change from the fibre
size to the characteristic fibre-architecture sizes, and
Subject Areas: eventually to a structural size, depending on the
composite configuration and structural geometry as
materials science
well as the imposed loading environment. Thus, a
physical modelling of failure in composites must
Keywords: necessarily be of multi-scale nature, although not
composite materials, fatigue, failure, always with the same hierarchy for each failure
failure theories, physical modelling mode. With this background, the paper examines the
currently available main composite failure theories to
assess their ability to capture the essential features
Author for correspondence:
of failure. A case is made for an alternative in
Ramesh Talreja the form of physical modelling and its skeleton
e-mail: talreja@aero.tamu.edu is constructed based on physical observations and
systematic analysis of the basic failure modes and
associated stress fields and energy balances.
This article is part of the themed issue ‘Multiscale
modelling of the structural integrity of composite
materials’.

1. Introduction
The potential of composite materials in providing
high stiffness, strength and toughness in lightweight
structural applications is accompanied by the richness in
the mechanisms underlying those properties. Inadequate
account of those mechanisms in design procedures
can lead to wasteful use of material and unreasonable

2016 The Author(s) Published by the Royal Society. All rights reserved.
Downloaded from http://rsta.royalsocietypublishing.org/ on June 4, 2016

reliance on testing. Historically, however, many analyses of composite materials were inspired
2
and motivated by experience with metals, which led to the use of homogenization procedures
that did not allow explicit account of failure mechanisms at length scales that were obliterated

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 374: 20150280


.........................................................
by the homogenization. While the homogenized models accurately predict the deformational
response of composites prior to initiation of failure mechanisms, such models do not handle well
the progression of the mechanisms and the criticality they induce.
The high and increasing computational power in recent years has opened up possibilities of
simulating microstructure-level failure events by numerical stress analysis. However, accurate
prediction of failure response requires use of appropriate physics of failure in the simulation
procedures. Even here, metals-based modelling concepts are likely to result in inaccurate
predictions, as will be discussed later.
The exposition to follow will first give a brief account of the homogenized models proposed
over the years to describe failure of composites. The inherent inadequacy of the models to
incorporate the physics of composite failure will be pointed out. Efforts to remedy the lack of
predictive capability of the models, while keeping the single scale of homogenized composites,
will be scrutinized. A case will be made to develop physics-based structural integrity models
for composites on their own without relying on metals-based concepts of multi-scale modelling.
A schematic failure analysis process for composite structures will then be outlined.

2. Failure theories for homogenized composites


The first notable effort to describe failure of a homogenized unidirectional (UD) composite
appeared in 1965 by Azzi & Tsai [1], which became known as the Tsai–Hill theory. Hill [2] in
1948 proposed a generalization of the yield criterion for isotropic metals to orthotropic metals
by assuming six independent yield stresses: three for normal stresses in the three symmetry
directions and three for shear stresses in the three planes of symmetry. It is noteworthy that while
the von Mises criterion expressed in stresses is derivable from the distortional energy density, this
is not the case for the Hill criterion. Thus, Hill’s generalization of the von Mises criterion is simply
a mathematical generalization. Still, one can argue that it is relevant to yielding of metals. Hill,
in fact, motivated his criterion from the need to describe yielding in metals that develop texture
from sheet forming by rolling, resulting in orthotropic symmetry. The Azzi–Tsai adaptation of
the Hill criterion to UD composites was motivated by assumed similarity of the rolling direction
in metal sheets and the reinforcing direction in UD composites. The final equation, referred to
as the Tsai–Hill criterion, was derived by equating the six yield stresses in the Hill criterion to
the corresponding ‘strengths’ of a UD composite, and further assuming transverse isotropy in
the composite cross-sectional plane. Because the UD composites used as layers in laminates are
usually thin, the criterion is commonly used in its two-dimensional version, given by
 σ 2  σ σ   σ 2  σ 2
1 1 2 2 12
− + + = 1, (2.1)
X X Y S
where σ1 and σ2 are the normal stresses in the fibre and transverse directions, respectively, σ12 is
the in-plane shear stress, and X, Y and S are the strength values of σ1 , σ2 and σ12 , respectively. For
unequal strengths in tension and compression, a variation of equation (2.1) is easily derived.
Although initial verification of equation (2.1) was indicated for the case of UD composites
loaded in tension at various angles to the fibre direction, many cases later did not support
the criterion.
The next noteworthy development of failure criteria for UD composites appeared in a paper
by Tsai & Wu [3] in 1971 based on a formulation proposed by Goldenblat & Kopnov [4] in 1965.
The proposed formulation is a scalar function expressed as a polynomial of all stress tensor
components. The coefficients of the polynomial terms represent the strength constants. The Tsai–
Wu modification of this function was its simplification to a quadratic expression that allowed
an interpretation as a quadric surface using analytical geometry. Thus, the ellipsoidal surface of
Downloaded from http://rsta.royalsocietypublishing.org/ on June 4, 2016

‘failure’ of a UD composite under in-plane stresses took the form


3
Fp σp + Fpq σp σq = 1, (2.2)

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 374: 20150280


.........................................................
where indices p and q take values 1, 2 and 6, and the summation convention for repeated indices
is implied. The index 6 on the stress components stands for shear, and is the same as 12 in
equation (2.1). The coefficient terms in the equation represent strength values. Coefficients of the
linear terms are non-zero when the strengths for positive and negative stress components differ.
Thus, F6 is always zero, and for the same reason, F16 and F26 vanish. Of the remaining coefficients,
F1 , F2 , F11 , F22 and F66 can be expressed in terms of the normal strength values in the fibre and
transverse directions and the shear strength in the plane of the composite. Finally, the coefficient
F12 can, in principle, be obtained by a biaxial test, but the non-uniqueness of this procedure
is a source of ambiguity in the criterion. It can be argued that because the Tsai–Wu criterion,
equation (2.2), is not connected to the specifics of the failure in composites, its applicability will
always be limited by the uncertainty of determining the F12 value.
At this point, it can be said that the two criteria discussed above are severely handicapped by
their inherent limitations: the Tsai–Hill criterion is motivated by an incorrect failure mechanism
(yielding, which is not the failure mechanism in UD composites with a non-metallic matrix) and
the Tsai–Wu criterion has no specifics of a failure mechanism, rendering it simply a quadratic
curve-fitting framework.
Numerous efforts to modify the type of failure criteria exemplified by the two criteria
discussed above, including applying these to the case of fatigue failure, have not produced
consistent improvements [5,6]. Other ways to approach formulation of failure criteria, while still
keeping the composite material homogenized, were initiated by Hashin [7] in 1980. These will be
briefly summarized next.
Hashin [7] pointed out that the single ellipsoid represented by equation (2.2) led to physically
unacceptable interactions between stress components in some cases. He suggested to introduce
piecewise smooth surfaces to describe critical states and additionally to recognize that composite
failure involving fibre breakage was governed by stresses differently from when the failure occurs
in the matrix only. Assuming the two failure modes to be independent, Hashin proceeded to
formulate criteria for them separately. Thus, for example, the fibre failure mode when the fibre
axial stress is tensile was assumed to be given by
 σ 2  σ 2
1 6
+ = 1. (2.3)
X S
Thus, the transverse normal stress was assumed not to affect the fibre failure mode, and the
quadratic form of the failure criterion was retained.
For the matrix failure mode, Hashin [7] proposed the notion of a failure plane not intersecting
fibres in a UD composite. Assuming the fibre direction stress not to have influence on this plane,
Hashin proposed that the other stress components would determine the inclination of the plane.
He outlined the difficulties involved in determining the angle of this inclination and proposed
that some (unknown) extremum principle would govern it. Notwithstanding Hashin’s word of
caution, Puck and co-workers [8,9] launched an elaborate procedure to develop failure criteria
incorporating the matrix failure plane idea. The resulting so-called Puck failure criteria for a UD
composite require determination of seven material constants.
Without conducting a detailed and exhaustive scrutiny of all proposed failure theories
that operate on the homogenized description of composites, one can reasonably put forth the
argument that the inherent limitation in such theories would still be the inability to incorporate
the conditions for initiation of the first failure event and for the sequential development
of subsequent failure events. These failure events necessarily occur at scales that have been
obliterated in the composite homogenization. The final failure event associated with criticality
(e.g. unstable progression of failure events or transition to another failure mode) cannot
generally be analysed without delineation of the subcritical path of the prior failure events.
These arguments point to an alternative approach to failure assessment of composites where
Downloaded from http://rsta.royalsocietypublishing.org/ on June 4, 2016

resort is made to the size scales of the composite constituents and the configurations in which
4
the constituents have been put together. The multi-scale approach to structural integrity of
composites will be discussed next.

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 374: 20150280


.........................................................
3. Physical modelling of failure: mechanisms
The failure theories for homogenized composites, discussed above, served their purpose when
understanding of the physical mechanisms was not adequate and the computer power needed
to simulate the initiation and progression of failure events in different failure modes was not
available. Today, there is sufficient basis for conducting failure analysis with a multi-scale
approach that connects the failure process at the governing scales with the structural scale
response characteristics needed for design. What remains to do is to develop efficient models
that capture the physics of failure without overly complicating the failure analysis to be used
by designers. The first step in this process is to have a clear understanding of the underlying
failure mechanisms. Because the failure mechanisms differ significantly depending on whether
fibre failures are involved and whether the driving forces for failure are tensile, compressive or
shear, the discussion to follow will be arranged to account for these aspects.

(a) Fibre failure mode in tension


Fibre failure within a UD composite under imposed overall axial tension has been a subject of
study for many years owing to the importance of this failure mode to the ultimate composite
failure. Direct and in situ observations of the failure mechanisms have been made difficult by
the limited applicability of microscopy. Early observations by scanning electron microscopy
combined with ultrasonic measurements [10] and by fractography [11], both on carbon–epoxy UD
composites, suggested a local failure process in clusters of fibres. More recent work using high-
resolution computed tomography [12] provided more details of the fibre failures and reported
statistics of the broken fibre clusters (figure 1). They reported that at low load levels, fibre breaks
appear singly and in clusters of two neighbouring fibres, while close to final failure, clusters of
four or more fibre breaks were found. They did not see larger clusters forming from accumulation
of smaller clusters, as the load was increased, but instead found them forming independently at
random locations.
The observations reported have focused on fibre failures, but have not sufficiently clarified the
roles of matrix and the fibre–matrix interface in transferring load from the broken to intact fibres,
except in trying to find evidence to support the proposed models for the load transfer.

(b) Fibre failure mode in compression


Although the strength of a UD composite under axial compression is characterized as a fibre
failure mode, as by Hashin [7], and subsequently by many, the failure mechanisms involved in
this case depend significantly on the matrix behaviour. As described by Jelf & Fleck [13], the
fibre failure in compression may be categorized as elastic microbuckling or plastic microbuckling,
depending on whether the matrix stress–strain behaviour is linear or nonlinear. For the latter
case, the compressive strength is predicted well by Budiansky’s kink band model [14], which has
the fibre misalignment and matrix shear yield stress as parameters. Kyriakidis et al. [15] further
verified the dependence of the compressive strength on fibre misalignment angle and generalized
it to fibre waviness as the microstructure (defect) scale. The microstructural features of the kink
bands and mechanisms of their formation were clarified by these authors. Their analysis found
that in the presence of fibre waviness, localization of shear deformation occurs in the matrix. This
forms bands and the flow of the matrix in the bands results in bending of fibres and eventual
breakage, which results in the formation of the observed kink bands. A fairly comprehensive
review of the role of various parameters on the compressive strength can be found in Waas &
Shultheisz [16]. Lee & Waas [17] conducted experimental and numerical studies of the influence
Downloaded from http://rsta.royalsocietypublishing.org/ on June 4, 2016

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 374: 20150280


.........................................................
singlets (1) (4)
increasing stress

doublets (2) formation of a critical cluster N = N* (5)

K-plet (3) catastrophic failure (6)

Figure 1. X-ray radiograph sequence shows the sequence of fibre failures under increasing axial tensile stress [12].

of a wide range of parameters such as fibre stiffness, fibre volume fraction and fibre misalignment
on the compressive strength.
Berbinau et al. [18] analysed the role of the fibre failure itself in the criticality of failure in
compression and showed that the earlier contention that the fibres fail by tension in bending
within the kink band was not plausible.

(c) Matrix and fibre–matrix interface failure in transverse tension


The cracks observed on loading a UD composite under tension normal to fibres have commonly
been referred to as ‘transverse cracks’. The appearance of these cracks is typified by the images
shown in figure 2 [19]. Owing to the brittle nature of such cracks (i.e. unstable growth), it is
common to study these in a constrained environment, e.g. in a laminate, often a cross ply laminate,
in order to have them stopped at the lamina/lamina interfaces. Furthermore, by applying cyclic
loading of appropriate level, the crack formation can be slowed down for closer examination. The
cracking images in figure 2 were obtained in cross ply laminates under cyclic transverse tension.
In the early work on composite failure, as described in common textbooks on introduction
to composite materials [20,21], the failure initiation in transverse tension is attributed to fibre–
matrix de-bonding. While observations seem to imply this, a closer analysis suggests that a failure
mechanism in the matrix could be a precursor of the observed de-bonding. This will be discussed
further in §4.
Downloaded from http://rsta.royalsocietypublishing.org/ on June 4, 2016

(a)
6

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 374: 20150280


.........................................................
20 mm

(b)

20 mm

Figure 2. (a,b) Images illustrative of crack formation under tension normal to fibres (adapted from Gamstedt & Sjögren [19]).
The initial fibre–matrix de-bonds coalesce into larger transverse cracks.

(d) Matrix failure in transverse compression


Studies of failure in transverse compression of thin UD composites (lamina) are few owing to
difficulties of in situ observations. The earliest observations of transverse compression failure
can be found reported in Agarwal et al. [20] on thick UD composites that show failure on
planes parallel to fibres but otherwise inclined to the loading direction. This seems to suggest an
influence of the shear stress in the failure process. More recent observations of this failure mode
have been reported in Gonzalez & Llorca [22], reproduced in figure 3. As seen in figure 3, the
failure plane is inclined (in this case at 56◦ ), confirming earlier observations. Their observations
of the fracture surface showed the presence of hackles, supporting previous suggestions that local
shear in the matrix plays a role in this failure mode.

(e) Matrix failure in in-plane shear


Applying in-plane shear to a thin and flat UD composite causes difficulties of producing a
uniform stress state owing to the influence from gripping ends. The preferred specimen is
therefore a thin-walled tube in torsion with the fibres running in the circumferential (hoop)
direction. Quaresimin & Carraro [23] and Carraro & Quaresimin [24] used this specimen
and loading mode. Their microscopic observations of failure induced by cyclic in-plane shear
confirmed earlier observations by Redon [25] (figure 4) and by Plumtree & Shi [26]. As seen in
figure 4, microcracks develop in the matrix along planes inclined to the fibres in a UD composite
subjected to in-plane shear. These cracks tend to turn and grow in the fibre direction, merging
together to form ‘axial’ cracks.
Downloaded from http://rsta.royalsocietypublishing.org/ on June 4, 2016

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 374: 20150280


.........................................................
56°

100 mm

Figure 3. Scanning electron micrograph of the cross-sectional plane of a UD carbon–epoxy specimen loaded in transverse
compression (adapted from Gonzalez & Llorca [22]).

t
0.1 mm

Figure 4. Cracks formed in matrix planes inclined to the fibre under an in-plane shear stress τ [25].

(f) Failure modes in combined loading


The individual failure initiation mechanisms described in the previous sections are governed by
the appropriate local conditions. The criticality of the local conditions in each case will be affected
when the loading modes are combined.
For fibre failure in tension, presence of the transverse stress (tension or compression) or the
in-plane shear stress, or both, will aid or abate the formation of failed fibre clusters depending
on how the local load transfer from a failed fibre to its neighbouring fibres is affected. Systematic
experimental observations of these effects do not seem to be available.
In the case of axial compression in the presence of in-plane shear, Jelf & Fleck [27] reported
that for a carbon–epoxy UD composite, produced as pultruded thin tubes, superposition of
torque on axial compression did not change the mechanism of plastic microbuckling resulting
in kink band formation. Although applying torque alone caused splitting parallel to fibres, when
it was combined with axial compression, failure initiated as plastic microbuckling. Vogler et al.
[28] conducted tests on a UD carbon–polyetheretherketone produced as flat specimens subjected
to axial compression and in-plane shear. Their experimental observations indicated kink band
formation to be the governing mechanism in compression, the effect of in-plane shear being
rotation and bending of the fibres within the band. The cause of fibre breakage was found to
be the axial compression.
For matrix failure under combined action of transverse tension and in-plane shear, the failure
initiation mechanism to occur first will depend on whether the local stress field causes the
critical conditions for fibre–matrix de-bonding or for producing cracks on planes inclined to
the fibres. Observations by Quaresimin & Carraro [23] on glass–epoxy tubes with fibres in the
circumferential direction subjected to simultaneous cyclic axial tension and torsion indicated
Downloaded from http://rsta.royalsocietypublishing.org/ on June 4, 2016

that two regimes of interaction existed: for low-to-moderate values of in-plane shear, the
8
dominating mechanism was fibre–matrix de-bonding, whereas for higher values of shear, the
cracks forming along the fibres resulted from growth and merger of the cracks illustrated in

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 374: 20150280


.........................................................
figure 2. Plumtree & Shi [26] reported similar observations on flat UD coupons of carbon–epoxy
loaded in off-axis tension.
Observations on UD composite failure under transverse compression were noted above.
Combining this loading mode with in-plane shear has experimental challenges. There seems to
be no data reported for this case.

4. Physical modelling of failure: multi-scale approaches


A general review of multi-scale approaches to composite materials can be found in [29]. Our
focus here is to use the multi-scale methodology to support physical modelling, by which is
meant to capture understanding of the observed failure mechanisms into models that have at
the minimum two scales: a scale where the failure events occur and a scale where description of
the structural failure could be given. In this context, the following subsections describe multi-scale
approaches with respect to the individual failure modes. Section 4a describes an overall strategy
for failure assessment.

(a) Fibre failure mode in tension


The observed failure process in this case, as described in §3a, suggests that the first scale at which
a failure event occurs is the fibre diameter. A broken fibre either exists from the manufacturing
process or it results from loading as breakage at the weakest point. The next event is failure of
fibres in the neighbourhood of a broken fibre caused by the stress enhancement induced by the
first broken fibre. Observations indicate localized failure in clusters of fibres, one or more of which
grow unstably, triggering ultimate failure of the UD composite. A refinement in the failure process
is multiple failed fibre clusters interconnecting to a larger failed plane growing unstably to failure.
Most literature has focused on analysing conditions for formation of a cluster of broken
fibres. The scale at which the stress and failure analysis is conducted consists of a representative
region containing sufficient fibres to simulate a fibre-cluster failure. An example of this is in
Nedele & Wisnom [30] where a representative region is modelled as an assembly of five concentric
cylinders. This axisymmetrical model is illustrated in figure 5.
The Nedele–Wisnom analysis calculated the stress enhancement in fibres neighbouring the
central broken fibre and the probability of failure in these fibres accounting for the statistical
fibre strength distribution. The matrix in the model was assumed as either elastic or perfectly
plastic and consideration was given to a frictional stress transfer from the de-bonded length of
the broken fibre. No account was taken of cracking of the matrix in the interfibre region. This has
been done in an unpublished work [31]. Leaving the details of the stress and failure analysis to
[31], the main results are as follows. The broken fibre placed at the centre of the axisymmetrical
model initiates matrix cracking or fibre–matrix de-bonding depending on whether the broken
fibre exists before loading is applied or if it breaks under loading. The subsequent failure events
consist of failure of fibres neighbouring the existing broken fibre. The stress enhancement in the
neighbouring fibres is affected by the matrix crack emanating from the broken fibre end or as a
result of kinking out of the de-bond crack. The stress enhancement on the neighbouring fibres
extends over an axial distance and its intensity depends on the extent to which the matrix crack
has grown, and if the broken central fibre has de-bonded from the matrix, then the extent of the
de-bonding matters.

(b) Fibre failure mode in compression


As described in §3b, the critical mechanism for this mode is kink-band formation, as
observed in most glass–epoxy and carbon–epoxy composites. Budiansky & Fleck [32] review
Downloaded from http://rsta.royalsocietypublishing.org/ on June 4, 2016

(a) (b)
C 9

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 374: 20150280


.........................................................
M
M F F

cross section of fracture plane cross section of simplified fracture plane


: initial broken fibre M: matrix
: neighbouring fibres near broken fibre F: fibre
: surrounding matrix C: effective composite
other white areas are effective composite phase

Figure 5. A fibre cluster of focus for composite failure analysis (a) is modelled as a five-cylinder axisymmetrical assembly (b).
The central broken fibre is surrounded in the model by a cylinder of matrix, which in turn is surrounded by a cylinder of smeared-
out fibres. The cylinder of the outer matrix is placed to maintain the composite fibre volume fraction and the homogenized
composite cylinder beyond represents the surrounding composite. Adapted from [31].

analytical treatments of kink bands, where the initial fibre misalignment angle appears as the
microstructural parameter. Kyriakidis et al. [15] modelled the UD composite as alternating
layers of fibres and matrix and introduced the axial misalignment as waviness, which for long
wavelength approximates to the misalignment. This microstructural model clarified the roles of
matrix properties as well as the sensitivity of the compressive strength (limit load) to the initial
imperfections.

(c) Matrix and fibre–matrix interface failure in transverse tension


It is important to recognize that in the absence of a crack, or a weak plane that can potentially
form a crack, failure in a given material will be a ‘point process’, i.e. it will initiate at a material
point when critical conditions for the particular failure mode are met at that point. Subsequent
events will depend on the region in which the failed material point exists. If the material point
lies in a brittle region, which has little capacity to deform inelastically (i.e. irreversibly), then the
consequence of failure can be formation of a brittle crack, if appropriate conditions exist, such as
a critical release of energy along a preferred plane. On the other hand, if the region is a polymer
that deforms inelastically, then crazes or shear bands, or both, can form [33], leading eventually
to ductile failure.
Before examining whether the failure ensued is to be brittle or ductile, it is appropriate to look
at the stress field under which failure initiates. It is a common error to assume that a polymer
that shows extensive inelastic deformation under a uniaxial test in the laboratory also behaves in-
elastically within a composite with stiff reinforcement. Asp et al. [34] found that the behaviour of
an epoxy polymer subjected to the so-called composite-like stress state was brittle, in spite of
extensive ductility displayed under uniaxial stress. This gave a plausible explanation of the
observed low strain to failure under transverse tension of UD composites and the insensitivity
of this strain to modifications of the polymer morphology.
Downloaded from http://rsta.royalsocietypublishing.org/ on June 4, 2016

Asp et al. [35,36] pursued the line of thinking in [34] and developed a criterion for cavitation-
10
induced brittle failure in the matrix polymer within composites. According to their work, in
regions of matrix close to the stiff fibres in a UD composite, the stress state developed under

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 374: 20150280


.........................................................
remote transverse tension is triaxial with nearly equal principal stresses. Independent tests
conducted on an epoxy polymer produced a critical value of the dilatational energy density, which
when used for the same epoxy within the composite, predicted the composite failure stress well.
The observed mechanism in the unreinforced polymer under hydrostatic tension was cavitation,
and it was reasonable to assume that the expansion of the cavity formed becomes unstable at a
material-specific value associated with the initiation of the instability.
A remarkable implication of the Asp et al. theory [35,36] is that what is commonly assumed
to be fibre–matrix de-bonding may actually be a consequence of the brittle failure induced by the
cavity ‘burst’ when the critical dilatational energy density is reached. Because this happens close
to the fibre surface (at specific locations), where the largest dilatational energy density exists, the
consequent fibre–matrix interface breakage appears as de-bonding.
The fibre–matrix de-bonding mechanism is also possible without the matrix cavitation failure
as a precursor. This will be the case if favourable conditions for cavitation do not exist or if the
fibre–matrix interface is sufficiently weak or is sufficiently weakened by defects. In that case, the
radial tensile stress, and possibly this stress combined with the shear stress on the fibre surface,
will break the fibre–matrix interface bonds, initiating de-bonding. For modelling purposes, it is
difficult to know what interface failure properties to use, as these depend on the actual quality of
the bond formed during the composite manufacturing process, which is not the same for model
composites with one or few filaments that are used for experimental determination of the bond
characteristics. The interface bond strength cannot be determined accurately by theoretical means.
Several experimental methods have therefore been devised (see Zhandarov & Mader [37] for a
review). These are either stress-based (strength) or energy-based (toughness) methods. As noted
above, the characteristics of the fibre–matrix interface obtained by the experimental methods
are commonly under simple conditions (e.g. single-fibre tests). Using the material constants
thus obtained for interfaces that are under constrained conditions within composites would
require caution. Applying the interface toughness criterion for evaluating initiation of fibre–
matrix de-bonding also faces difficulties owing to the uncertainty of knowing the flaw size and
its variability.
For ductile failure to occur under transverse tension of a UD composite, conditions must
exist for inelastic deformation of the polymer matrix within the composite. Because the inelastic
deformation requires shear stress to drive it, the distortional part of the strain energy density at
a point must be sufficient to initiate what may be called ‘yielding’ (although this term originates
from crystalline metal behaviour). For isotropic metals, the initiation of yielding is satisfactorily
given by the critical value of the distortional energy density obtained experimentally for the
given metal. Equivalently, the yield criterion for metals can be expressed in terms of the second
invariant of the deviatoric stress tensor, as is the case for the von Mises criterion. The initiation
of yielding in polymers is governed by molecular phenomena that differ significantly from the
dislocation motion underlying yielding of crystalline metals. Still, it is common to describe the
onset of inelastic deformation in polymers by the approaches used for metal yielding. In contrast
to metals, the inelastic response of glassy polymers displays pressure sensitivity, as discussed by
Rottler & Robbins [38]. This has prompted modifying the metal yield criteria by including the
hydrostatic stress, e.g. by adding to the threshold of the octahedral shear stress τ0 a constant α
times the mean pressure p, as
y
τoct = τ0 + αp, (4.1)

where p is the average of the three principal stresses, p = −(σ1 + σ2 + σ3 )/3.


Equation (4.1) is the modified von Mises yield criterion, which in energy terms states that the
dilatational energy density contributes as well to the shear-driven onset of inelastic deformation
in polymers. Additionally, temperature and strain rate are also found to affect the inelastic
deformation [39].
Downloaded from http://rsta.royalsocietypublishing.org/ on June 4, 2016

In a polymer matrix within a UD composite, the stress triaxiality is generally high except in
11
resin-rich regions. The inelastic deformation will thus tend to occur away from the fibre–matrix
interfaces. Once initiated, the inelastic deformation can lead to shear banding before crack

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 374: 20150280


.........................................................
formation. Estevez et al. [40] have studied the crack formation process in glassy polymers by
considering the competition between shear banding and crazing. Based on their study, it can be
stated that the role of the distortional part of the strain energy density at a point is to localize
inelastic deformation in shear bands, whereas the dilatational component is responsible for
cavitation leading to craze formation, craze widening and breakdown of craze fibrils. The mix
of the two energy components determines the ease or difficulty of ductile crack formation in the
matrix within the composite.
Models for ductile crack formation in polymers are complex and require many material
constants for implementation. Huang & Talreja [41] have shown that the Rice–Tracey ductile
fracture model [42] is capable of predicting ductile cracking in polymers with high accuracy. The
Rice–Tracey model assumes that pre-existing microvoids (from defects or inclusions) grow within
plastically deformed material to a critical size, at which point neighbouring microvoids coalesce
and cause fracture. The microvoid growth is controlled by the stress triaxiality and equivalent
plastic strain, given by    εp  
R 3σm
D = ln = α exp dεp , (4.2)
R0 εp0 2σp
where σm and σp are the mean stress and the von Mises equivalent stress, respectively, εp is the
equivalent plastic strain, and R and R0 are the current and initial void radii, respectively. The
lower limit of integration is often taken as zero for simplicity. In implementation, the damage
variable D = ln(R/R0 ) is used instead of the void size. When D reaches a critical value Dc at
the considered material point, a crack is assumed to occur or to pass through that point by
void coalescence. Rice & Tracey [42] calculated the constant α to be 0.283 under assumptions
of spherical voids and an infinite medium. The critical void growth ratio Dc is believed to
be a material constant that does not change with geometry or loading conditions. However, it
cannot be directly measured. Instead, a calibration procedure is carried out to obtain its value by
matching the simulation-predicted results with experimental measurements for either a smooth
or pre-cracked specimen under selected loading conditions. The value of Dc thus obtained
is then used to predict the ductile fracture of the same material under other geometries and
loading conditions.

(d) Matrix failure in transverse compression


As noted in §3d, observed failure of a UD composite under transverse compression indicates that
failure occurs on a plane inclined to the loading direction. This prompts the suggestion that the
failure mode is a manifestation of the matrix yielding in compression governed by cohesion and
friction acting on the plane. The criterion for this type of yielding is the classical Mohr–Coulomb
criterion, expressed by
τ = c − σ tan φ, (4.3)
where τ is the shear stress at yielding acting on the critical plane, σ is the normal stress on that
plane, and c and φ are material constants representing cohesion and friction angle, respectively.
This criterion is similar to equation (4.1), if c is considered yield stress in pure shear and φ is
viewed as the effect of hydrostatic stress. The angle made by the inclined failure plane with respect
to the normal to the compression loading direction is given by
π φ
β= + . (4.4)
4 2
In a UD composite, when compression is applied normal to the fibres, the inclination of the failure
plane will be altered by the triaxility of the local stress state [22,43].
The formation of the failure plane in a UD composite involves progression from yielding at
the most favourable point to connectivity of the yield (failure) planes at the subsequent points
Downloaded from http://rsta.royalsocietypublishing.org/ on June 4, 2016

of yielding. This connectivity is supposedly by bands in which ductile deformation and failure
12
mechanisms take place. These aspects cannot be treated in a homogenized composite, but require
representation of the microstructure for detailed local stress field calculations. Gonzalez & Llorca

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 374: 20150280


.........................................................
[22] have presented an approach of this nature. They include in their analysis also the effect of
fibre–matrix interface failure via an assumed cohesive zone model.

(e) Matrix failure in in-plane shear


The stress–strain behaviour of UD composites under pure in-plane shear is found to be nonlinear.
Part of this nonlinearity is attributed to the inelastic deformation of the matrix and the other
part to the formation of the type of cracks shown in figure 4. Unless defects in the matrix lie
with potential weak planes inclined to the fibre direction, assuming these cracks to form owing
to the tensile stress acting normal to those planes would not be reasonable. Puck & Shurmann
[44] offered, however, the explanation that these cracks form at 45◦ to the fibre direction under
the action of the tensile principal stress, assuming a pure shear stress in the matrix between the
fibres. Actual experimental evidence, such as in figure 4, shows these cracks to be not at 45◦ and
not straight. These cracks are curved (‘sigmoidal’, as described in [26]), multiple, and tend to
link up near the fibre surfaces, eventually forming a wavy crack (described in observations as a
transverse crack) with its plane running overall parallel to the fibres. Along the lines of Puck &
Shurmann [44], Carraro & Quaresimin [24] assumed the shear-induced crack to be formed by
the largest local (tensile) principal stress, which they calculated numerically. The direction of the
local maximum principal stress differed from 45◦ to the fibre axis, assumed in [44] because of the
triaxiality of the local stress field.
The assumption of a nucleation (fracture) plane in polymers warrants closer examination.
Contrary to metals, where specific crystalline planes exist, and can be expected to nucleate
fracture planes under appropriate stress conditions, in polymers no such planes exist. To form
a fracture plane, the molecules will need to be stretched beyond their breaking point in a manner
that breakage points are roughly aligned along a plane. Because the observed shear-induced
cracks do not appear to be aligned with the plane of maximum shear, assuming a ‘stretch and
break’ process, governed by tension, is reasonable.

(f) Failure modes in combined loading


Having considered the individual failure modes in single loads, i.e. tension or compression in
fibre or transverse directions, or shear in the plane of a UD composite, the next area of inquiry
is concerning failure when these loads are combined. The combined effect is perceived to be an
‘interaction’, meaning mutually enhanced degradation (or, exceptionally, strengthening).
The notion of interaction in combined loading has its history in yielding of metals. It is
conceivable that each loading mode applied separately, or two or more loading modes applied
together, may contribute to energy needed to drive the mechanisms underlying yielding. These
mechanisms are basically related to dislocation mobility, and therefore the relevant strain energy
density involved is distortional. This is a key reason for the von Mises criterion’s ability to
successfully describe the combined effect of the normal and shear stresses. The ensuing quadratic
expression of the criterion indicates the nature of the combined action of the individual stress
components in initiating yielding.
For combined action of the stress components in contributing to the failure of a composite,
there is no fundamental justification similar to that for metal yielding. First of all, there is no
single failure mode in composites such as metal yielding to warrant a quadratic combination
of the partial effect of each stress component. The conditions governing each failure mode are
different, as discussed above. Therefore, the only reasonable course of action in describing the
combined effects is to consider each failure mode separately and analyse how the individual
effects combine for that particular failure mode. This will be attempted next.
Downloaded from http://rsta.royalsocietypublishing.org/ on June 4, 2016

A general remark before considering the combined effects of loading modes is in order.
13
Because each loading mode generally induces a separate failure mechanism, it would be
convenient and appropriate to consider a given failure mode as being driven primarily by one

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 374: 20150280


.........................................................
loading mode, denoted as the ‘dominant’ loading mode, and any other loading mode that is acting
simultaneously with this loading mode as the ‘modifying’ loading mode. Thus, for example, the
remotely applied tension is the dominant loading mode for tensile failure of fibres and an in-plane
shear would be the modifying loading mode for this failure mode.
Let us begin with the fibre failure mode in tension, discussed in §§3a and 4a. The essential
feature of this failure mode is the local load sharing from one failed fibre to its neighbouring intact
fibres. The properties of the matrix and the fibre–matrix interface as well as the microstructural
parameters such as the interfibre spacing play roles in this load transfer process. If remotely
applied transverse tension (or compression) and in-plane shear stress are combined with the
axial tension, then these additional loads will affect the local load transfer by changing the local
stress field. Without fully analysing these modifying effects, it is commonly assumed that the
transverse load (tension or compression) has no effect and that the in-plane shear contributes to
the fibre failure mode in tension in the form of a quadratic interaction, as in equation (2.3). In
fact, this equation also suggests that if the in-plane shear is the dominant loading mode, then
the mechanisms of shear cracks leading to transverse crack formation, discussed in §§3e and 4e,
will be modified by the imposed axial tension in the same way. Obviously, this assertion is not
supported by any analysis at the microstructural level.
Consider next the fibre failure mode in compression, discussed in §§3b and 4b. Although
different mechanisms underlying this failure mode are possible, in the common case of glass–
epoxy and carbon–epoxy, the mechanism involved is plastic microbuckling leading to kink-band
formation. The roles of the local shear stress and fibre misalignment have been identified as the
important factors governing this mechanism. If transverse normal stress (tensile or compressive)
and the in-plane shear stress are applied in addition to the axial tension, then their effects must be
considered in terms of any changes induced in the local shear stress and in the fibre misalignment.
Experimental studies have focused on combining the axial compression with the in-plane shear
[27,28]. In those experiments, both these studies found that the fibre failure mode in compression
remained unchanged by the additional application of the in-plane shear. If this is the case, then
the contribution of the in-plane shear can be analysed in terms of the modification induced in
the local shear associated with the kink-band formation. It must be realized, however, that if the
applied axial compression does not trigger plastic microbuckling, then increasing the in-plane
shear would likely result in shear cracks followed by transverse cracks, the mechanisms that
come into play when the in-plane shear alone is applied. The role of the axial compression in
these mechanisms is as yet not clear.
As far as failure in the matrix polymer is concerned, all combinations of loads applied—
tension and compression normal to fibres and in-plane shear—induce mechanisms of yielding
and crack formation. In going from the unreinforced polymer to the same polymer within a
UD composite, care must be exercised in accounting for the stress triaxiality induced by the
presence of fibres and the potential failure surfaces at the fibre–matrix interfaces. In theories
developed for homogenized composites, such as by Puck & Schurmann [8], these aspects
cannot be accounted for directly. Various assumptions and curve-fitting parameters must be
introduced related to the fracture planes within the matrix between the fibres. Although such
theories facilitate simple procedures for design, certain fundamental objections to their validity
remain. For instance, what triggers the formation of the fracture planes within composites
where the local stress fields are essentially triaxial? In other words, if no specific weak
planes exist beforehand in the matrix, then what is the mechanism involved in getting to
the flat planes of fracture from discrete failures at points? It is true that fracture planes are
observed in experiments, but closer examination of the planes often suggests some precursor
mechanism(s). For instance, in the case of compression normal to fibres, the observed hackles
near the fibres on the inclined fracture planes indicate the role of the local stresses in initiating
discrete failures.
Downloaded from http://rsta.royalsocietypublishing.org/ on June 4, 2016

interior 14
delaminations

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 374: 20150280


.........................................................
interior
delaminations
axial
axial
splits
splits transverse
cracks

transverse
cracks

Figure 6. X-ray radiograph (top left) [45] shows interior details of failure events in a cross ply laminate. The arrows show
transverse cracks, axial splits and interior delaminations. The three-dimensional sketch (top right) depicts the same failure
events for clarity. Image (bottom left) (J. Varna 2016, personal communication, Luleå University of Technology, Luleå, Sweden)
shows transverse cracks with interface cracks at the transverse crack fronts. The bottom right image (J. Varna 2016, personal
communication, Luleå University of Technology, Luleå, Sweden) shows closer detail of coupling of transverse cracks through the
interface. (Online version in colour.)

5. Beyond unidirectional composite failure


UD composites form the building blocks in many composite structures, e.g. in stacked
configurations in multidirectional laminates. Failure ensued in a single layer is often the
beginning and not the end of the structural failure. In going beyond UD composite failure, two
aspects are noteworthy: one, that the UD composite failure within a layer is necessarily under
combined loading, and two, that this failure is constrained by the presence of the other differently
oriented layers. When failure in a UD composite lying within a laminate occurs, it is not an
ultimate failure in the sense of breaking apart as separate pieces, but rather as initiation of a crack
that has its plane aligned with the fibre direction in the composite layer. This failure, described
in the early literature as ‘first-ply failure’, has a progression consisting of multiple parallel cracks,
which increase in number, reducing their mutual spacing, until a saturation state is reached.
Failure in other plies progresses similarly, in a sequence determined by the criticality of their
stress states.
Although investigated thoroughly over the years, and familiar to most composite researchers,
the UD composite cracking process in a laminate was briefly outlined above to put in perspective
the theories and models discussed in previous sections concerning UD composite failure. Going
beyond these theories to treating composite laminate failure has additional challenges. Figure 6
illustrates the nature of these challenges by displaying the range of failure modes, within plies
and between plies. The images in figure 6 pertain to a cross-ply laminate tested in tension–
tension fatigue to the point when all of its ‘subcritical’ failure events have taken place, leaving
the axial plies to fail in what was described above as fibre failure mode in tension. The subcritical
failures are marked in the X-ray radiograph in the figure and for clarity are illustrated in a
Downloaded from http://rsta.royalsocietypublishing.org/ on June 4, 2016

lamina level failure analysis 15


(RVE-1)

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 374: 20150280


.........................................................
composite microstructure, matrix, interface fibre
manufacturing defects failure modes failure modes
(RIMS)

laminate level failure analysis


(RVE-2)

multiple lamina cracking,


delamination, fibre failure

Figure 7. A failure assessment roadmap for composite laminates. (Online version in colour.)

three-dimensional sketch on the side. Images at the bottom of the figure show the multiple ply
cracks (left) and their diversion into the ply interfaces is focused on in the image to the right.
Until significant coupling of the transverse cracks through the interfaces occurs, the individual
plies retain their load-bearing capacity in spite of sustaining a multitude of cracks. The
displacement of the crack surfaces under loading changes the average deformational response
of the laminate, however. This response has been treated fairly generally in Talreja [46,47], and
since then by many researchers. Our focus here is failure in the sense of loss of load-bearing
capacity, which for laminates degrades significantly when delamination of the plies disrupts the
initial load sharing among the plies with perfectly bonded ply–ply interfaces. Assessment of this
requires a fracture mechanics methodology for analysing advancement of multiply connected
cracks. Such a methodology is as yet unavailable. Current efforts using cohesive zones face
uncertainty of determining the properties (parameters characterizing the de-cohesion process)
by experimental means.

6. A laminate failure assessment methodology


While the physical modelling of failure in composites has still challenges to meet, a roadmap to
reaching the goal can be outlined. Such a map is laid out in figure 7. As indicated there, the first
step is to characterize the microstructure of the composite as it results from the manufacturing
process used. The characterization of the microstructure is to be at a scale needed for the failure
analysis. Thus, depending on which failure mode is being analysed, e.g. a fibre failure mode
in a UD composite, the choice of the microstructural region will be accordingly. In each case,
the microstructural region must also include the appropriate manufacturing defects (e.g. fibre
misalignment) within the region. This region, often called the representative volume element
(RVE), is denoted here as real initial material state (RIMS) to capture the fact that it represents
the ‘real’ material state as produced at the end of the manufacturing process, i.e. ‘initially’, when
the composite is placed in the service environment. The realization of RIMS is labelled as RVE-1
and RVE-2 at the lamina and the laminate levels, respectively. The stress and failure analysis of
the appropriate RVE will generate the conditions (criteria) for the particular failure mode. The
failure conditions for lamina failure modes will enter the RVE-based analyses of the subsequent
failure modes in a laminate.

7. Concluding remarks
A scrutiny of the fundamentals of the common failure theories advanced over the past many
years reveals the inherent deficiency of the theories to treat the range of failure mechanisms
observed. This deficiency is rooted in homogenizing the composite microstructure and thereby
losing the ability to explicitly analyse the initiation and progression of the failure events involved.
Downloaded from http://rsta.royalsocietypublishing.org/ on June 4, 2016

Current capabilities of observing the composite microstructure before, during and after failure,
16
aided by the computer power to simulate the failure process, calls for a different approach.
Such an approach has been discussed here. It has been motivated by the detailed physical

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 374: 20150280


.........................................................
understanding of the individual failure modes operating in UD composites under different
imposed elementary loading modes and their combinations. The models to treat the failure modes
have also been discussed.
The current state of composite failure assessment has advanced to the stage that UD composite
failure can be simulated well, although some improvements can still be made. Greater challenges
lie in dealing with failure in composite laminates, and subsequently in composite structures.
Hurdles are mainly in treating intralaminar cracks that have diverted into the lamina interfaces,
and the interconnection between these cracks that results from growth in the interfaces. The load-
bearing capacity of a laminate with this type of damage is yet to be determined reliably. Current
trends in treating this problem by the cohesive zone models are subject to uncertainty owing to
the non-uniqueness of the material parameters (strength and fracture toughness) entering such
models.
Competing interests. The author declares that there are no competing interests.
Funding. I received no funding for this study.

References
1. Azzi VD, Tsai SW. 1965 Anisotropic strength of composites. Exp. Mech. 5, 283–288. (doi:10.
1007/BF02326292)
2. Hill R. 1948 A theory of the yielding and plastic flow of anisotropic metals. Proc. R. Soc. Lond.
A 193, 281–297. (doi:10.1098/rspa.1948.0045)
3. Tsai SW, Wu EM. 1971 A general theory of strength for anisotropic materials. J. Compos. Mater.
5, 58–80. (doi:10.1177/002199837100500106)
4. Goldenblat II, Kopnov VA. 1965 A strength criterion for anisotropic materials (generalized
strength criterion for anisotropic materials with different tensile, compression and shear
strengths). Akademiia Nauk SSSR, Izvestiia, Mekhanika 5, 77–83.
5. Talreja R. 2014 Assessment of the fundamentals of failure theories for composite materials.
Compos. Sci. Technol. 105, 190–201. (doi:10.1016/j.compscitech.2014.10.014)
6. Quaresimin M, Susmel L, Talreja R. 2010 Fatigue behaviour and life assessment of
composite laminates under multiaxial loadings. Int. J. Fatigue 32, 2–16. (doi:10.1016/j.ijfatigue.
2009.02.012)
7. Hashin Z. 1980 Failure criteria for unidirectional fiber composites. J. Appl. Mech. 47, 329–334.
(doi:10.1115/1.3153664)
8. Puck A, Schürmann H. 1998 Failure analysis of FRP laminates by means of physically
based phenomenological models. Compos. Sci. Technol. 58, 1045–1067. (doi:10.1016/S0266-
3538(96)00140-6)
9. Puck A, Kopp J, Knops M. 2002 Guidelines for the determination of the parameters
in Puck’s action plane strength criterion. Compos. Sci. Technol. 62, 371–378. (doi:10.1016/
S0266-3538(01)00202-0)
10. Fuwa M, Bunsell AR, Harris B. 1975 Tensile failure mechanisms in carbon fibre reinforced
plastics. J. Mater. Sci. 10, 2062–2070. (doi:10.1007/BF00557484)
11. Purslow D. 1981 Some fundamental aspects of composites fractography. Composites 12,
241–247. (doi:10.1016/0010-4361(81)90012-4)
12. Aroush DR, Maire E, Gauthier C, Youssef S, Cloetens P, Wagner HD. 2006 A study
of fracture of unidirectional composites using in situ high-resolution synchrotron X-ray
microtomography. Compos. Sci. Technol. 66, 1348–1353. (doi:10.1016/j.compscitech.2005.09.010)
13. Budiansky B. 1983 Micromechanics. Comput. Struct. 16, 3–12. (doi:10.1016/0045-7949(83)
90141-4)
14. Jelf PM, Fleck NA. 1992 Compression failure mechanisms in unidirectional composites.
J. Compos. Mater. 26, 2706–2726. (doi:10.1177/002199839202601804)
15. Kyriakides S, Arseculeratne R, Perry EJ, Liechti KM. 1995 On the compressive failure of fiber
reinforced composites. Int. J. Solids Struct. 32, 689–738. (doi:10.1016/0020-7683(94)00157-R)
Downloaded from http://rsta.royalsocietypublishing.org/ on June 4, 2016

16. Waas AM, Schultheisz CR. 1996 Compressive failure of composites, part II: experimental
17
studies. Prog. Aerospace Sci. 32, 43–78. (doi:10.1016/0376-0421(94)00003-4)
17. Lee SH, Waas AM. 1999 Compressive response and failure of fiber reinforced unidirectional

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 374: 20150280


.........................................................
composites. Int. J. Fract. 100, 275–306. (doi:10.1023/A:1018779307931)
18. Berbinau P, Soutis C, Guz IA. 1999 Compressive failure of 0◦ unidirectional carbon-
fibre-reinforced plastic (CFRP) laminates by fibre microbuckling. Compos. Sci. Technol. 59,
1451–1455. (doi:10.1016/S0266-3538(98)00181-X)
19. Gamstedt EK, Sjögren BA. 1999 Micromechanisms in tension-compression fatigue of
composite laminates containing transverse plies. Compos. Sci. Technol. 59, 167–178. (doi:10.
1016/S0266-3538(98)00061-X)
20. Agarwal BD, Broutman LJ, Chandrashekhara K. 2006 Analysis and performance of fiber
composites. Hoboken, NJ: John Wiley & Sons.
21. Daniel IM, Ishai O. 1994 Engineering mechanics of composite materials. New York, NY: Oxford
University Press.
22. González C, LLorca J. 2007 Mechanical behavior of unidirectional fiber-reinforced polymers
under transverse compression: microscopic mechanisms and modeling. Compos. Sci. Technol.
67, 2795–2806. (doi:10.1016/j.compscitech.2007.02.001)
23. Quaresimin M, Carraro PA. 2014 Damage initiation and evolution in glass/epoxy tubes
subjected to combined tension–torsion fatigue loading. Int. J. Fatigue 63, 25–35. (doi:10.1016/j.
ijfatigue.2014.01.002)
24. Carraro PA, Quaresimin M. 2014 A damage based model for crack initiation in
unidirectional composites under multiaxial cyclic loading. Compos. Sci. Technol. 99, 154–163.
(doi:10.1016/j.compscitech.2014.05.012)
25. Redon O. 2000 Fatigue damage development and failure in unidirectional and angle-ply glass
fibre/carbon fibre hybrid laminates. Roskilde, Denmark: Risø National Laboratory.
26. Plumtree A, Shi L. 2002 Fatigue damage evolution in off-axis unidirectional CFRP. Int. J.
Fatigue 24, 155–159. (doi:10.1016/S0142-1123(01)00068-8)
27. Jelf PM, Fleck NA. 1994 The failure of composite tubes due to combined compression and
torsion. J. Mater. Sci. 29, 3080–3084. (doi:10.1007/BF01117623)
28. Vogler TJ, Hsu SY, Kyriakides S. 2000 Composite failure under combined compression and
shear. Int. J. Solids Struct. 37, 1765–1791. (doi:10.1016/S0020-7683(98)00323-0)
29. Kanouté P, Boso DP, Chaboche JL, Schrefler BA. 2009 Multiscale methods for composites: a
review. Arch. Comput. Methods Eng. 16, 31–75. (doi:10.1007/s11831-008-9028-8)
30. Nedele MR, Wisnom MR. 1994 Stress concentration factors around a broken fibre
in a unidirectional carbon fibre-reinforced epoxy. Composites 25, 549–557. (doi:10.1016/
0010-4361(94)90183-X)
31. Zhuang L, Talreja R, Varna J. Submitted. Tensile failure of a unidirectional composite from a
local fracture plane.
32. Budiansky B, Fleck NA. 1994 Compressive kinking of fiber composites: a topical review. Appl.
Mech. Rev. 47, S246–S250. (doi:10.1115/1.3124417)
33. Donald AM, Kramer EJ. 1982 Effect of molecular entanglements on craze microstructure
in glassy polymers. J. Polym. Sci. Polym. Phys. Ed. 20, 899–909. (doi:10.1002/pol.1982.18020
0512)
34. Asp LE, Berglund LA, Gudmundson P. 1995 Effects of a composite-like stress state on the
fracture of epoxies. Compos. Sci. Technol. 53, 27–37. (doi:10.1016/0266-3538(94)00075-1)
35. Asp LE, Berglund LA, Talreja R. 1996 Prediction of matrix-initiated transverse failure in
polymer composites. Compos. Sci. Technol. 56, 1089–1097. (doi:10.1016/0266-3538(96)00074-7)
36. Asp LE, Berglund LA, Talreja R. 1996 A criterion for crack initiation in glassy polymers
subjected to a composite-like stress state. Compos. Sci. Technol. 56, 1291–1301. (doi:10.
1016/S0266-3538(96)00090-5)
37. Zhandarov S, Mäder E. 2005 Characterization of fiber/matrix interface strength: applicability
of different tests, approaches and parameters. Compos. Sci. Technol. 65, 149–160. (doi:10.
1016/j.compscitech.2004.07.003)
38. Rottler J, Robbins MO. 2001 Yield conditions for deformation of amorphous polymer glasses.
Phys. Rev. E 64, 051801. (doi:10.1103/PhysRevE.64.051801)
39. Arruda EM, Boyce MC, Jayachandran R. 1995 Effects of strain rate, temperature and
thermomechanical coupling on the finite strain deformation of glassy polymers. Mech. Mater.
19, 193–212. (doi:10.1016/0167-6636(94)00034-E)
Downloaded from http://rsta.royalsocietypublishing.org/ on June 4, 2016

40. Estevez R, Tijssens MG, Van der Giessen E. 2000 Modeling of the competition between
18
shear yielding and crazing in glassy polymers. J. Mech. Phys. Solids 48, 2585–2617. (doi:10.
1016/S0022-5096(00)00016-8)

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 374: 20150280


.........................................................
41. Huang H, Talreja R. 2006 Numerical simulation of matrix micro-cracking in short fiber
reinforced polymer composites: initiation and propagation. Compos. Sci. Technol. 66, 2743–
2757. (doi:10.1016/j.compscitech.2006.03.013)
42. Rice JR, Tracey DM. 1969 On the ductile enlargement of voids in triaxial stress fields. J. Mech.
Phys. Solids 17, 201–217. (doi:10.1016/0022-5096(69)90033-7)
43. Pinho ST, Robinson P, Iannucci L. 2006 Fracture toughness of the tensile and compressive fibre
failure modes in laminated composites. Compos. Sci. Technol. 66, 2069–2079. (doi:10.1016/j.
compscitech.2005.12.023)
44. Puck A, Schürmann H. 2002 Failure analysis of FRP laminates by means of physically
based phenomenological models. Compos. Sci. Technol. 62, 1633–1662. (doi:10.1016/
S0266-3538(01)00208-1)
45. Jamison RD, Schulte K, Reifsnider KL, Stinchcomb WW. 1984 Characterization and
analysis of damage mechanisms in tension-tension fatigue of graphite/epoxy laminates.
In Effects of defects in composite materials. West Conshohocken, PA: ASTM International.
(doi:10.1520/STP30196S)
46. Talreja R. 1985 A continuum mechanics characterization of damage in composite materials.
Proc. R. Soc. Lond. A 399, 195–216. (doi:10.1098/rspa.1985.0055)
47. Talreja R. 1985 Transverse cracking and stiffness reduction in composite laminates. J. Compos.
Mater. 19, 355–375. (doi:10.1177/002199838501900404)

You might also like