You are on page 1of 13

Finite element modeling of mechanically fastened composite-aluminum

joints in aircraft structures


Zlatan Kapidzic
a,b,
, Larsgunnar Nilsson
b
, Hans Ansell
a
a
Saab AB, SE-581 88 Linkping, Sweden
b
Division of Solid Mechanics, Linkping University, SE-581 83 Linkping, Sweden
a r t i c l e i n f o
Article history:
Available online 11 November 2013
Keywords:
Bolted joints
Composite-aluminum
Finite element modeling
Hybrid wing structures
a b s t r a c t
A three-dimensional, solid nite element model of a composite-aluminum single-lap bolted joint with a
countersunk titanium fastener is developed. The model includes progressive damage behavior of the
composite and a plasticity model for the metals. The response to static loading is compared to experimen-
tal results from the literature. It is shown that the model predicts the initiation and the development of
the damage well, up to failure load. The model is used to evaluate the local forcedisplacement responses
of a number of single-lap joints installed in a hybrid composite-aluminum wing-like structure. A
structural model is made where the fasteners are represented by two-node connector elements which
are assigned the forcedisplacement characteristics determined by local models. The behavior of the
wing box is simulated for bending and twisting loads applied together with an increased temperature
and the distribution of fastener forces and the progressive fastener failure is studied. It is shown that
the fastener forces caused by the temperature difference are of signicant magnitude and should be taken
into account in the design of hybrid aircraft structures. It is concluded that, the account of the non-linear
response of the joints results in a less conservative load distribution at ultimate failure load.
2013 Elsevier Ltd. All rights reserved.
1. Introduction
The proportion of ber-reinforced composites in modern air-
craft structures is constantly increasing due to their favorable
lightweight material properties. At the same time, metals are still
required in primary aircraft components e.g. when damage toler-
ance design is utilized. Consequently, the number of hybrid inter-
faces where composites and metals are joined, is growing. A
common type of such interface is a shear loaded, bolted joint with
a titanium bolt and carbon ber-reinforced polymer (CFRP) and
aluminum plates. This study deals with simulation of structures
that contain a large number of such joints.
Bolted joints are often weak parts of the structure. It is therefore
important that they are properly designed from static, fatigue and
damage tolerance points of view. From the weight point of view,
the joints must not be conservatively over-dimensioned. The need
for an effective and accurate analysis method is from these reasons
apparent. Traditionally, the development and design of joints have
been based on experimental testing of specimens, which has been
shown to be both expensive and time-consuming. In recent years,
the use of nite element (FE) methods to simulate the behavior of
composite joints has increased. In order to replace the experi-
ments, the analysis method should be able to accurately capture
a number of issues, such as: three-dimensional state of stress
and strain, material behavior, bolt pretension and clamping force,
bolt hole clearance, friction, secondary bending effects, contact be-
tween the surfaces etc. Another matter, that further complicates
the modeling, is that there is no obvious ready-to-use material
model for the failure of composites. The complex nature of CFRP
failure makes it very hard to derive a universal material model
with damage initiation and propagation that works well in all pos-
sible situations. Nevertheless, a large number of different criteria
for the initiation and development of damage have been proposed
over the years. A review with the most common methodologies for
modeling of failure can be found in [1].
Another review particularly concerning the failure of bolted
composite joints is presented in [2]. Since that review, a vast num-
ber of analytical studies of CFRP fastened joints has been pub-
lished. They range from two-dimensional models with rigid pins
representing the bolt [35], to three-dimensional models with or
without damage development [615]. In these studies, bolt preten-
sion, bolt clearance, the effect of countersunk/protruding bolt
heads, different failure initiation and damage progression models,
by-pass loading and implicit/explicit solution methods have been
studied. Also, a large number of experimental investigations has
been done, for instance in [16,17]. The level of detail in FE
0263-8223/$ - see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.compstruct.2013.10.056

Corresponding author at: Saab AB, SE-581 88 Linkping, Sweden. Tel.: +46
(0)13183884.
E-mail address: zlatan.kapidzic@saabgroup.com (Z. Kapidzic ).
Composite Structures 109 (2014) 198210
Contents lists available at ScienceDirect
Composite Structures
j our nal homepage: www. el sevi er . com/ l ocat e/ compst r uct
modeling, in the studies mentioned, requires an extensive com-
puter power in order for the computational time to be reasonable,
even for small joint installations. In structures with a very
large number of fasteners, the computational and modeling
burden tends to become unmanageable if no simplications are
introduced.
This has motivated several authors to develop different ap-
proaches suited for the incorporation of mechanically fastened
joints into FE analysis on a structural level. In most of these meth-
ods, the fasteners are represented by structural nite elements. For
example, Weyer et al. [18] used connector elements available in
the program Abaqus to represent self-piercing rivets in crash anal-
ysis. The modeling included elastic, plastic and damage behavior of
a single bolted metal sheet specimen. Gray and McCarhty [19]
developed a method using beam elements connected to a rigid sur-
face to represent the bolt, and shell elements to model the lami-
nate plates of the joint. Bolt-hole clearance, bolt-torque and
friction between the plates were considered, but the damage dissi-
pation in laminates was not accounted for. The functionality of the
model was demonstrated on a three-bolt single-lap joint. Another
efcient approach was developed by Gray and McCarthy [20], were
a user-dened element was implemented in Abaqus. The method
is capable of modeling non-linear behavior and failure of compos-
ite joints based on semi-empirical approach. A joint with twenty
fasteners was used for validation of the method. Ekh and Schn
[21] made use of beam elements to represent both bolt and lami-
nates and connector elements to account for bolt-hole clearance
and friction. The model was suited for optimization of load trans-
fer, but was limited to single-column joints. A two-dimensional
model of a steelconcrete beam with spring elements representing
the shear connections was developed by Wang and Chung [22]. A
non-linear behavior of the springs, based on experimental data
was utilized.
The motivation for this work comes from the intrinsic problems
of hybrid structures. If materials with different thermal and
mechanical properties are mixed in structures, issues that are usu-
ally absent in homogeneous structures can arise. Some examples of
material properties that can differ for composite laminates and
aluminum are: thermal expansion coefcients, failure and fracture
mechanisms, response to different types of loading, i.e. tensile
versus compressive, fatigue accumulation and scatter, impact
resistance, impact residual strength, degree of anisotropy, environ-
mental sensitivity etc. Based on these differences, composite and
aluminum materials used in aircraft structures are subject to dif-
ferent design and airworthiness requirements. The issues that arise
in the design and certication of hybrid aircraft structures as a re-
sult of this mismatch are: thermally induced loads and deforma-
tions, multiple failure modes in joints, allowance of buckling and
permanent deformations, determination of factors for statistical
scatter of material properties, determination of signicant load
states etc. These issues have to be taken into account in the design
and certication of hybrid aircraft structures. The recommended
practice for certication of composite assemblies is to begin ana-
lyzing and testing at a small specimen level. This way, the risks
in technology can be eliminated at an early stage before moving
onto testing more complex and expensive structural parts. This is
known as the Building Block Approach (BBA) [23]. When it comes
to hybrid constitutions this approach may not always be appropri-
ate. As a result of the material mismatch, a hybrid joint may be-
have in one way when tested as a specimen and in another way
when installed in a structure. Based on this argument, the analyzes
and tests should be done at a structural level rst in order to eval-
uate the true behavior of a structure. Such testing can be very
expensive and a lot of effort can be saved if these tests can be done
virtually, i.e. in a computer simulation. In this study, the behavior
of a large winglike hybrid box is examined by FE simulations.
Fatigue testing of this structure exposed to spectrum loading and
an applied temperature, as well as static testing, will be performed
in the near future and the results will be compared. The objective
of this work is to develop a methodology for simulation of struc-
tures that contain hybrid CFRP-aluminum shear loaded bolted
joints with damage and failure behavior.
The study is divided into two parts. In the rst part, the local
behavior of a CFRP-aluminum single bolt joint is modeled using
the FE code Abaqus [24]. A three-dimensional FE model, using solid
elements, is developed. For the composite plate, a progressive
damage model (PDM) is implemented and aluminum and titanium
parts are modeled using an elasto-plastic material behavior.
The resulting joint behavior is compared to results from the
experiments performed by Ireman et al. [16] in order to verify
the method. In the second part, the method from the rst part is
used to evaluate the local behavior of several joint conguration
that are present in the wing box. These behaviors are then assigned
to connector elements available in Abaqus and included into the
model of the wing box. Mechanical and thermal loads are applied
and the redistribution of fastener loads due to joint failure is
examined.
2. Modeling of composite-aluminum bolted joints
In this section, the mechanical behavior of a single-lap, shear
loaded bolted joint specimen shown in Fig. 1 is assessed. The spec-
imen was previously tested by Ireman et al. [16] where it was de-
noted AC6III. It contains one plate made of CFRP material HTA/
6376 with ply thickness of 0.13 mm and a quasi-isotropic stacking
sequence 45=0=90
4

S
. The other plate is made of aluminum
AA7475-T76. A single 6 mm countersunk bolt, made of titanium
Ti6Al4V STA is used with nger-tight pretension. The tting tol-
erance between the bolt and hole is according to ISO f7/H10, which
gives a clearance of 1070 lm.
The testing with an applied shear load was performed without
any lateral support, which means that secondary bending effects
were present in the joint. The loaddeection response was mon-
itored at the loading grips and the relative displacement of the
plates was obtained by extensiometers attached to the sides of
the specimen. The specimen was loaded until nal failure occurred
in a bearing failure mode. Ireman [12] also performed an FE anal-
ysis of the specimen testing but did neither include the failure
behavior of the composite nor the plasticity of the metals.
2.1. Finite element modeling
In this work, FE modeling of the AC6III specimen is performed
using eight-node brick elements with reduced integration, C3D8R
in Abaqus. The whole mesh is shown in Fig. 2 and a closer, cut-
through view of the bolt and composite plate is shown in Fig. 3.
As can be seen, the bolt, the washer and the nut are considered
to be one solid piece. The aluminum plate is meshed with 8 ele-
ments in the thickness direction, and in the CFRP plate each ply
is represented using one element. In both plates, the mesh is re-
ned close to the contact area between the bolt and the plates.
Some mesh convergence studies, based on stresses and strains in
this area are performed, but the results are not presented in this
paper. For stabilization of the initial loading procedure and to
eliminate rigid body modes, weak spring elements are connected
to the plates and the bolt in all three directions. Furthermore, a
linear spring element is connected in series with the specimen to
account for the exibility of the testing machine, as was done in
[25].
In order to eliminate hourglass modes in the reduced integrated
elements, Abaqus assigns an hourglass stiffness based on the
Z. Kapidzic et al. / Composite Structures 109 (2014) 198210 199
elasticity parameters of the element. However, when the material
stiffness is exposed to signicant reduction, as in this case, en-
hanced hourglass control is recommended to stabilize the zero en-
ergy modes. The articial energy introduced with the enhanced
hourglass control is checked to ensure that it is negligible and does
not affect the accuracy of the solution.
The contact between the surfaces in the joint is modeled using
the general contact algorithm available in Abaqus. It considers all
neighboring surfaces that might come into contact during the anal-
ysis and uses nite sliding with a surface-to-surface option. The
penalty approach, with default stiffness is used to enforce the con-
tact constraints. A Coulomb friction is assigned to all surfaces with
a friction coefcient of 0.2. This value is experimentally estimated
by Schn [26] and used by Ireman [12]. Initial clearance is
eliminated between the bolt and hole edge. The aluminum and
titanium are modeled using an elasto-plastic material model with
isotropic hardening. The assigned properties are shown in Table 1.
For CFRP material, a phenomenological procedure, known as the
ply-discount method (PDM), is implemented. A non-linear shear
stressstrain relation is assumed. The initial elastic properties
and unidirectional strengths are shown in Table 2 and details of
the implementation are explained in the next section.
As shown in Fig. 2, one free end of the specimen is clamped,
while a displacement is incrementally applied to the other free
end. The analysis is performed using large displacement formula-
tion and automatic incrementation.
2.2. Progressive damage modeling of CFRP
The ply-discount method is a progressive damage procedure
that aims to represent the accumulation of damage in a CFRP
Fig. 1. Dimensions in mm of the single-bolted joint specimen AC6III tested in [16].
Applied displacement
Aluminum
Composite
Clamped edge
(u=v=w=0)
(v=w=0)
z
x
y
Fig. 2. The FE model of single-bolted joint from [16].
Fig. 3. Cut-through view of the bolted joint mesh with the aluminum plate
removed.
Table 1
Elasto-plastic properties of AA7475-T76 and Ti6Al4V STA.
Property AA7475-T76 Ti6Al4V STA
Youngs modulus, E (MPa) 69,000 110,000
Poissons ratio m 0.28 0.29
Yield stress, ry (MPa) and ry p ry p
plastic strain, p 400 0 950 0
430 0.002 1034 0.002
490 0.09 1103 0.1
200 Z. Kapidzic et al. / Composite Structures 109 (2014) 198210
laminate. In a typical procedure, the load is increased until a cer-
tain failure criterion is fullled at a material point. Depending on
the activated failure mode, specic stiffness parameters are re-
duced at that material point. In this way, a decrease of load-bearing
capacity is achieved in a direction consistent with the triggered
failure mode. Although very simple and effective, the ply-discount
method suffers from a major drawback. Even if the failure criteria
employed may be physically based, the choice of degraded mate-
rial parameters and the level of degradation is completely arbi-
trary. A good understanding of the failure conditions and damage
development is required to produce accurate results. Also, the re-
sults predicted should preferably be correlated to experimental
evidence to ensure their validity. Despite its shortcomings, the
ply-discount method has been successfully implemented by
numerous authors to different joint congurations. The most com-
monly used failure criteria in three-dimensional models e.g. [10]
are the criteria proposed by Hashin [27]. In two-dimensional mod-
els, e.g. [35] the applied criteria used are the ones proposed by
Chang and Chang [28] and Chang and Lessard [29].
In the current study, the failure criteria proposed by Olmedo
and Santiuste [6] are utilized. These criteria are based on previ-
ously proposed criteria by Chang and Chang [28] and Chang and
Lessard [29], which were extended in [6] to include the out-of-
plane stresses. Both the extended and the original criteria include
the following non-linear relation between the in-plane shear strain
and the shear stress, as derived by Hahn and Tsai [30]
c
12

1
G
12
s
12
as
3
12
1
where G
12
is the in-plane shear modulus and a is an experimentally
determined constant. If a 0, the non-linear strainstress depen-
dency is eliminated. It is assumed that the non-linearity mainly is
caused by micro-cracking in the matrix. Therefore, similar relations
can be assumed to hold for the out-of-plane shear strains, as was
done by Olmedo and Santiuste [6]. In order to implement Eq. (1)
into the FE-model, the non-linear term is rewritten such that a
numerically stable algorithm is obtained, as described in [24]. The
shear stress at the end of the load increment i is then given as a lin-
ear function of the shear strain
s
i1
12
1 dG
12
c
i1
12
2
where
d
3aG
12
s
i
12

2
2as
i
12

3
=c
i
12
1 3aG
12
s
i
12

2
3
The parameter d can be seen as a damage parameter that de-
grades the initial shear modulus G
12
to its current value 1 dG
12
.
In [6], the proposed criteria include ber tensile and compres-
sive failure, matrix in-plane tensile and compressive failure, matrix
out-of-plane tensile and compressive failure and bermatrix
shearing failure.
Fiber tensile failure is assumed to have occurred in a material
point if the following condition is satised for r
11
P0

r
11
X
T

2

s
2
12
2G
12

3
4
as
4
12
S
2
12
2G
12

3
4
aS
4
12

s
2
13
2G
13

3
4
as
4
13
S
2
13
2G
13

3
4
aS
4
13
v
u
u
u
t 1 4
The third term under the square root is due to the contribution
from the out-of-plane shear stress to the ber failure. If a 0, and
the shear strength S
12
S
13
, the criterion reduces to the tensile -
ber mode criterion proposed by Hashin [27]. Originally, Chang and
Chang [28] assumed that S
12
used should be the in situ ply shear
strength measured from a cross-ply laminate, since that value
may differ signicantly from the strength measured from unidirec-
tional plies.
Fiber compressive failure is assumed to take place if the com-
pressive stress (r
11
0) is equal to the ber buckling strength
jr
11
j
X
C
1 5
This is the simplest and the most commonly used assumption,
where obviously, the inuence of the other stress components is
neglected.
The matrix in-plane failure criterion is met if

r
22
Y

2

s
2
12
2G
12

3
4
as
4
12
S
2
12
2G
12

3
4
aS
4
12

s
2
23
2G
23

3
4
as
4
23
S
2
23
2G
23

3
4
aS
4
23
v
u
u
u
t 1 6
where Y Y
T
for r
22
P0 and Y Y
C
for r
22
0. The same com-
ments following the ber tensile failure criterion are valid here, ex-
cept that the criterion reduces to the Hashin and Rotem criterion
[31] for a 0.
Matrix out-of-plane failure is considered by expanding the Ha-
shin [27] criterion for delamination initiation with the non-linear
shear stress/strain behavior, which gives

r
33
Z

2

s
2
13
2G
13

3
4
as
4
13
S
2
13
2G
13

3
4
aS
4
13

s
2
23
2G
23

3
4
as
4
23
S
2
23
2G
23

3
4
aS
4
23
v
u
u
u
t 1 7
where Z Z
T
for r
33
P0 and Z Z
C
for r
33
0. This criterion is not
present in the original set of criteria from [28,29]. It is included in
[6] to account for the damage caused by effects of tightening torque
in pretightened joints.
Fibermatrix shearing failure is initiated for r
11
0 when

r
11
X
C

2

s
2
12
2G
12

3
4
as
4
12
S
2
12
2G
12

3
4
aS
4
12

s
2
13
2G
13

3
4
as
4
13
S
2
13
2G
13

3
4
aS
4
13
v
u
u
u
t 1 8
As in Eq. (4) and (6) the out-of-plane shear stress is added to
the original criterion. It is obvious that the criteria (8) and (5)
are identical if the shear stresses vanish. Consequently, the latter
criterion should also enclose the ber compression. However, -
ber buckling may follow subsequent to bermatrix shearing
since only the shear stiffnesses degrade after bermatrix shear-
ing failure.
If a failure initiation criterion is met at a material point, certain
stiffness parameters of that material point are reduced. The sim-
plest alternative would be to reduce all parameters to zero, which
Table 2
Lamina elastic and strength properties for HTA7/6376.
Property
Longitudinal modulus E
11
(GPa) 141
Transverse modulus E
22
(GPa) 10
Out-of-plane modulus E
33
(GPa) 11
In-plane shear modulus G
12
(GPa) 5.2
Out-of-plane shear modulus G
13
(GPa) 5.2
Out-of-plane shear modulus G
23
(GPa) 3.9
In-plane Poissons ratio m
12
0.3
Out-of-plane Poissons ratio m
13
0.5
Out-of-plane Poissons ratio m
23
0.5
Longitudinal tensile strength X
T
(MPa) 2250
Longitudinal compressive strength X
C
(MPa) 1400
Transverse tensile strength Y
T
(MPa) 65
Transverse compressive strength Y
C
(MPa) 300
Out-of-plane tensile strength Z
T
(MPa) 30
Out-of-plane compressive strength Z
C
(MPa) 344
In-plane shear strength S
12
(MPa) 120
Out-of-plane shear strength S
13
(MPa) 80
Out-of-plane shear strength S
23
(MPa) 80
Z. Kapidzic et al. / Composite Structures 109 (2014) 198210 201
corresponds to a sudden brittle failure with no energy absorption
in the damage process. This is not a very realistic assumption, since
it has been conrmed in many experimental studies that the stress
degrades gradually after the onset of failure. In their work, Zang
and Rowland [11] and Tserpes et al. [25] concluded that the reduc-
tion of the parameters to zero gives a too severe degradation when
compared to experimental results. It also causes convergence prob-
lems in the FE-solution procedure because the elements in the area
of large contact pressure tend to become highly distorted. A more
realistic set of degradation rules were introduced by Tan [32] and
extended by Camanho [10] to three dimensions. Similar rules were
used in [4,6,11,25] all with good correlation to experimental re-
sults. The same degradation factors are used in this work and are
shown in Table 3.
The progressive damage model, with the failure criteria de-
scribed in this section, is implemented using the Abaqus subrou-
tine USDFLD. This subroutine allows the user to assign the
material properties as functions of user dened eld variables,
which in turn can be functions of eld quantities at material points.
In this case, these variables are dened as the failure initiation cri-
teria presented in the equations above and are functions of the
stress tensor components at material points. The analysis is driven
by incremental increase of the applied load and the strain and
stress elds are evaluated at the end of each increment. When
the stresses fulll a failure criterion at a material point, the mate-
rial properties are degraded. At this point, iterations should be per-
formed at the same load level in order to capture the redistribution
of the stresses to other material points which then also may fail.
However, if the load increment size is very small, these additional
iterations may be omitted [8]. The analyzes preformed in this study
utilize the automatic time stepping control with maximal load
increment of one per cent of the total applied load and are termi-
nated when severe convergence difculties are encountered.
2.3. Model validation
The results from the FE-analyzes are compared to experimental
results performed by Ireman et al. [16]. In the experimental set up,
the behavior of the joint was evaluated by measuring the applied
forcedisplacement response and the relative displacements of
the plates. The latter measurements were performed at low load
levels and are not used here for comparison with FE-results. In
Fig. 4, the applied forcedisplacement response from experiments
is compared with three different versions of FE-results. In the rst
version, only the linear elastic material model is utilized in all com-
ponents, i.e. no composite damage or metal plasticity is involved.
The second version includes the elasto-plastic material model for
metal parts and in the third version both the composite damage
model and the metal plasticity are present.
Initially, the response is linear, both in the experiment and in
the analyzes, indicating the absence of pretension and bolt hole
clearance effects. At approximately 7 kN, the experimental curve
and the curve with composite damage and metal plasticity start
to deect from the linear behavior. The other two curves continue
linearly until the curve with plasticity deects slightly fromthe lin-
ear one. This indicates that the effects of damage accumulation in
the composite are initiated at at about 7 kN. The experimental
curve reaches a maximal load just below 14 kN, where ultimate
bearing failure in the composite plate occurs and the load carrying
capacity starts to diminish. At the same point, the FE-analysis
starts having convergence difculties and the run is terminated
shortly afterwards. Fig. 4 also shows the load level at which the
ultimate shear failure of the bolt is expected to occur. This level
is simply the material shear strength times the area of the bolt
cross-section. It is evident that the load in the experiment never
quite reaches this level, which conrms that the joint failed in
the bearing failure mode. However, it is reasonable to expect that
the failure mode would be bolt shear failure for the same bolt
diameter and a signicantly thicker laminate. Fig. 5 shows the
cross-section of the composite plate at 14 kN and the elements
where the failure criteria are met and material properties
degraded.
It is concluded that the PDM material model used is able to cap-
ture the failure initiation point and the damage development in the
composite up to ultimate failure load. The progression of failure
beyond this point is not modeled. The degradation parameters
used here were successfully employed in other studies, e.g. in
[10], for similar quasi-isotropic laminates but for different types
of bolt installations. However, it is uncertain whether the same
parameters would be appropriate for other types of lay-ups and
laminate materials. Therefore, experimental verication of the cho-
sen parameters should always be performed for new laminate
congurations.
Alternative approaches to modeling damage and subsequent
loss of stiffness in composite materials have been developed in
the context of continuum damage mechanics (CDM). After failure
initiation, the damage development due to each failure mechanism
is typically controlled by a damage variable. The material stiffness
properties are successively reduced depending on the current
damage state, which in turn depends on a chosen damage evolu-
tion law. This law is thus a key parameter in the framework of
CDM. The choice of law is often based on some kind of physical
or energy-conserving ground. Usually, once the damage variable
has reached its maximal allowable value the material stiffness is
exhausted. A large number of CDM models are published in the lit-
erature. The implemented material models are often demonstrated
on regular coupons or specimens with open hole tension or com-
pression and they seem to work well. Bearing failure of bolted
joints, on the other hand, is problematic in the sense that the area
of damage concentration is also the area of contact between the
Table 3
Degradation rules in different failure modes.
Failure mode E
11
E
22
E
33
G
12
G
13
G
23
m
12
m
13
m
23
Matrix failure 0.4 0.4 0.2 0 0 0
Fiber failue 0.14 0.4 0.4 0.25 0.25 0.2 0 0 0
Fibermatrix shear 0.25 0.25 0 0
Shear non-linearity 1 d
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
0
2
4
6
8
10
12
14
16
18
Bolt shear failure
u [mm]
F

[
k
N
]
Experiment AC6III
FE, no damage, no plasticity
FE, no damage and plasticity
FE, damage and plasticity
Fig. 4. Comparison between experimental results and FE-results.
202 Z. Kapidzic et al. / Composite Structures 109 (2014) 198210
bolt and the hole edge. A complete degradation of stiffness and the
subsequent element deletion in the area of contact creates conver-
gence problems and yields poor results. From a physical point of
view and based on the experimental observations, material re-
moval is not very realistic. The material that has failed due to the
bearing failure mode seem to have some residual stiffness left
but there is no obvious way to model this kind of behavior and fur-
ther research is required.
3. Structural modeling
This section deals with the FE-modeling of a wing-like CFRP-
aluminum structure, that is to be fatigue tested at elevated tem-
perature. The purpose of designing, analyzing and testing such a
structure is to address the issues of hybrid structures described
in the introducing part of this article. The focus is directed on mod-
eling and studying the failure behavior of structural joints that
contain a large number of bolts and on the redistribution of loads
due to joint failure.
The structure is a box of dimensions 3300x630x150 mm and
consists of four CFRP skins, eight C-shaped aluminum spars, six
aluminum rib sections and one aluminum splice section, see
Fig. 6. The material properties are the same as in Table 1 and
Table 2. All parts are held together by bolted joints with titanium
bolts. All four skins have the same lay-ups and thicknesses and
each skin has three areas with different thickness: the skin
area overlapping the splice has the thickness of 8.32 mm, the area
between the splice and rib Section 6.24 mm and the area
outside the rib Section 7.28 mm. Each composite skin area is
quasi-isotropic and symmetrically stacked using a layup sequence
45=0=90
n

S
, where n is the number of sub-sequences. Notice
that this is the same layup as in the bolted joint in the previous sec-
tion. Between each skin area there is a ply drop-off region. Table 4
shows the thicknesses of the aluminum parts. All bolts have a
6 mm diameter except for the splice-skin bolts which are 8 mm.
The area between the ribs is the area of main interest in the exper-
iment. In this area, all the bolts in CFRP-aluminum joints are
countersunk.
In the experimental set up, the wing box will be placed on a
steel support base, simply supported at the ends on both upper
and lower side. The mechanical loads will be applied via two rect-
angular steel frames, clamped around the section of the box at rib
section positions. Four load actuators, two on each frame, will be
used to apply the load in the transverse direction. The diagonally
placed actuators will be assigned the same load, which will result
in a four point bending/twisting situation, see Fig. 7. Moreover, the
testing set up will at the same time be placed in a furnace that gen-
erates temperatures up to 90 C during the fatigue and static
testing.
3.1. The nite element model of the structure
The FE-model of the experimental set up is shown in Fig. 7.
All aluminumparts are represented as their middle surfaces and
meshed with shell elements S4R. The CFRP skins are modeled with
continuum shell elements, SC8R, that discretize the entire three-
dimensional geometry. One element with stacked lay-up formula-
tion is used in the thickness direction. The steel frames are
(Avg: 75%)
FV1
+0.0e+00
+5.0e01
+1.0e+00
(a) Matrix damage
(Avg: 75%)
FV2
+0.0e+00
+5.0e01
+1.0e+00
(b) Fiber-matrix shear damage
(Avg: 75%)
FV3
+0.0e+00
+5.0e01
+1.0e+00
(c) Fiber damage
Fig. 5. Cross-section of the composite plate at 14 kN and the damage eld variables.
Fig. 6. Studied wing box structure with dimensions 3300 630 150 [mm]. One of the upper skins is not shown and the other one is partially hidden for easier visibility of
the inner parts. The bolts connecting the skins with the inner structure are indicated in the gure.
Z. Kapidzic et al. / Composite Structures 109 (2014) 198210 203
modeled using solid brick elements and for the supports, rigid ele-
ments are used. All parts are assigned linear elastic material behav-
ior. Contact conditions are enforced between all parts in the model
with the general contact algorithm.
The solution of the FE-model is obtained in several consecutive
general steps in Abaqus/Standard. In the rst step, a pretension
load is applied to the vertical braces of the steel frames, while
the rigid supports are constrained in all directions. The purpose
of the pretension is to hold the frame in place while the actuator
loads are applied, so that the forces can be transferred to the wing
box via contact. For this purpose, a sufciently high pretension
load is chosen to ensure that the contact is maintained between
the frame and the box during loading. No estimation of the preten-
sion load is made in correlation with the experimental set up. In
the second step, a constant temperature difference DT = 50 C is
applied to all parts. Vertical point loads representing the actuator
forces are applied directly onto the steel frame in the third step,
see Fig. 7.
3.2. Bolted joints
Each bolt in the structure is represented by Abaqus connector
element CONN3D2 of type BUSHING. This is a two-node line ele-
ment, that has six degrees of freedom (DOF) in each node and con-
nects two parts at their interface. It is implemented by a Lagrange
multiplier technique, where the fastener forces are solved as the
constraint forces in the interconnections. The nodes of the connec-
tor element are not necessarily placed coincidentally with the
nodes at the interface of the bolted parts. Instead, the Abaqus fea-
ture *FASTENER is used to deploy distributing coupling elements
from the connector element nodes to the nodes of the parts that
are to be joined. In this way the connector elements representing
the bolts become independent of the surrounding mesh, which
makes the modeling work easier. There are two kinds of element
interfaces in the structural model: shell-to-shell (SS) interface
and continuum shell-to-shell (CS) interface. In the SS interface,
the shell elements placed at the middle surfaces of the plates are
linked by the connector elements. In the CS interface, the shell
elements placed at the middle of the aluminum plate are linked
to the lower side of the continuum shell element representing
the composite. This set-up is illustrated in Fig. 8.
Table 5 summarizes the different bolted joint congurations
and interfaces present in the FE-model of the wing box.
The local behavior for each connector element, in terms of
forcedisplacement characteristics, is determined from a local so-
lid FE-model of a single bolt joint. This model is principally the
same as the specimen model described previously. However, now
the plates are square with 4D long sides, where D is the bolt diam-
eter, as shown in Fig. 9. The small in-plane dimensions of the plates
are chosen in order to eliminate the secondary bending effects. It is
assumed that these effects also are negligible in the wing box, due
to the out-of-plane deection restrictions imposed on the joints by
lateral support of the surrounding structure. However, all single-
lap joints have an inherent load path eccentricity, even if the sec-
ondary bending is eliminated. This causes the bolt to tilt and bend,
Table 4
Aluminum component thicknesses. The anges of each component are the parts
connected to the skins, webs are the longitudinal parts and the stiffeners are the parts
that are perpendicular to the webs, see Fig. 6.
Flange Web Stiffener
Spar 4.5 3 2
Rib 4.5 3 3
Splice 6 3 3
Fig. 7. FE-model of the wing box structure. One of the skins is hidden for easier visibility of the inner structure.
Fig. 8. Shell-to-shell (SS) and continuum shell-to-shell (CS) interfaces.
204 Z. Kapidzic et al. / Composite Structures 109 (2014) 198210
which in turn gives an uneven contact pressure distribution over
the thickness which affects the damage distribution.
Another effect that is overlooked is the inuence of the by-pass
loads on the forcedisplacement characteristics of a joint. By-pass
loads certainly affect the damage propagation in the composite
plate [13] but it is unclear to what extent they also inuence the
exibility. Indeed, the by-pass loads are present in the wing box
joints. Moreover, the proportion of by-pass to transferred loads
vary for different bolts and depend on the applied load level. This
means that the connector element characteristics would have to
be bolt-dependent and solution-dependent in order to accommo-
date for any by-pass load effects on the exibility.
The bolts in the wing box will be torque tightened in the exper-
iment. Therefore, the pretension effects on the joint exibility in
the local model is briey studied. Fig. 10 shows the applied
forcedisplacement response of a local solid bolt model from the
skin-splice joint, with and without a pretension load on the bolt.
The applied pretension force is 15 kN and is included in the model
by an Abaqus built-in feature for pretension of sections. The gure
also shows the linear semi-empirically established response pro-
posed by Huth [33], which is often used in industrial applications
in load distribution analyzes. The load level, where the bolt shear
failure is expected, is also shown in the gure.
The effects of the tightening force on the forcedisplacement
behavior are apparent. In the initial loading, the pretension force
results in a higher load transfer for the same displacement due
to the friction forces. With further loading the two curves have
the same, almost linear, slope until the curve with no pretension
starts to deect. This is due to the initiation of the damage in the
composite plate. The curve with pretension departs from linear
at a higher load level. At the maximal load level, the two curves
converge to one, due to the loss of frictional contact. These effects
have also been conrmed experimentally by Ireman et al. [16]. It is
also interesting to notice that the load level, to which both curves
are converging, is the predicted bolt shear failure load. As antici-
pated, for thicker laminates the failure mode is bolt shear failure.
Since the ultimate failure level of the joint seems to be unaffected
by pretension, the pretension is neglected in connector element
characterization.
The displacement that is used to characterize the behavior of
the connector element is extracted form the local solid model
in different ways for the two element interface types. For the
SS type the connector element displacement is dened as the
relative displacement of the points A and C in Fig. 9, and for
the CS interface type the relative displacement of points A and
B is extracted. In this way the relative displacement of the con-
nector element nodes is related to the points where the element
is attached to the plates. The forcedisplacement curve extracted
from the solid model is given to the connector element as elasto-
plastic behavior assigned to the element shear resultant force as
function of the resultant in-plane displacement. The resultants
are formed as vector sums of the in-plane force and displacement
components. Remaining DOFs are assigned rigid behavior except
for the drilling DOF, i.e. rotation around the element line, which
is released. Once the load level of the shear resultant reaches
the bolt shear failure load, the connector damage functionality
in Abaqus is activated and the load is linearly decreased to zero
over a very short displacement range. This range is set to 1% of
the maximal plastic displacement. The sudden loss of load carry-
ing capacity signies the nature of fastener shear failure. When
the resultant force has dropped to zero the connector element
is removed from the model. Fig. 11 shows the forcedisplacement
curves for the three joint congurations present in the area be-
tween the rib sections. The connector elements outside the rib
sections are expected to be exposed to forces within the linear
elastic regime and are therefore assigned linear elastic behavior
according to Huth [33].
Damage behavior in the connector elements is characterized by
softening response that causes local instabilities and convergence
difculties in the implicit code. To overcome this issue two fea-
tures available in Abaqus/Standard are activated: viscous regulari-
zation and stabilization. The rst one is applied on connector
element level as a regularization of the elements constitutive re-
sponse by introducing viscosity into the damage evolution law.
The second feature provides stabilization of the solving process
by adding viscous forces to the global equilibrium equations. The
amount of viscosity introduced into the solution is in both cases
controlled by user-dened parameters, which have to be chosen
carefully. If the viscosity effect is too small it might not eliminate
the convergence problems, and too much of it might distort the
Table 5
The congurations of the bolted joints in the wing box.
Bolted
articles
Plate
materials
Plate thicknesses
(mm)
Bolt diameter
(mm)
Element
interface
Skin-Spar CFRP-Al 6.244.5 6 CS
Skin-Splice CFRP-Al 8.326 8 CS
Skin-Rib CFRP-Al 7.284.5 6 CS
Spar-Splice Al-Al 33 6 SS
Spar-Rib Al-Al 33 6 SS
A
F
F
B
C
Fig. 9. Local solid model of the single bolt joint.
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
0
5
10
15
20
25
30
35
Bolt shear failure
u [mm]
F

[
k
N
]
Pretension
No pretension
Semiempirical
Fig. 10. Response of the local solid bolt model in the skin-splice joint as the applied
force vs. the displacement of the loaded edge, see Table 5 for dimensions.
Z. Kapidzic et al. / Composite Structures 109 (2014) 198210 205
solution. Anyway, the energy dissipated by viscous regularization
and stabilization should be checked a posteriori to make sure that
it is negligible compared to the total energy.
3.3. Structural analysis results
This section presents the responses of three different bolt
groups in the wing-box, see Fig. 12(a).
The rst group is denoted Skin-Spar bolt group and connects
two of the skins to one of the outer spar pairs and to the splice. The
bolts are placed along x-direction in a single row containing 130
bolts. The second bolt group is the Skin-Splice bolt group and it
consists of four bolt rows in y-direction, with 16 bolts in each
row. This group connects two of the composite skins to the splice
ange. There are four bolts in this group that also belong to the rst
bolt group and these are the ones placed at the edge of the splice
ange, between the outer spars. Both of the mentioned bolt groups
are situated on the tensile loaded side of the wing box, which cor-
responds to the lower side of a real wing. The third bolt group is
the Spar-Splice bolt group and it connects each of the eight spars
to the splice stiffeners with 8 joints, each containing 5 bolts placed
in a row in the z-direction. Only 5 of these bolt rows are fully vis-
ible in Fig. 12(a) and the remaining 3 are indicated by arrows. All
fasteners and bolt rows are numbered within their own bolt group
and Fig. 12(b) shows the numbering in the splice area.
Fig. 13 shows the distribution of resultant shear bolt forces in
the three bolt groups due to an applied temperature of
DT = 50 C. Each bolt row in all three bolt groups displays a
U-shaped force distribution, which seems to be symmetric with
respect to the coordinate axis. In Fig. 13(a) disturbances in the
two U-shapes can be noticed. This is due to the expansion of the
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
0
5
10
15
20
25
30
35
u [mm]
F

[
k
N
]
Bolt shear failure, 8mm bolt
Bolt shear failure, 6mm bolt
SkinSplice
SkinSpar
SparSplice
Fig. 11. Elasto-plastic and damage behavior of the connector elements representing
the bolts in the area between the rib sections. For conguration of the joints, see
Table 5 and for the location, see Fig. 6.
Fig. 12. Studied bolt groups with row and fastener numbering.
206 Z. Kapidzic et al. / Composite Structures 109 (2014) 198210
aluminum ribs in the y-direction, whereby a force in the same
direction is transferred via the skin-spar bolts at rib positions.
From the overall maximal force levels and Fig. 11, the conclusion
is drawn that all bolts still are in the linear elastic regime at the
currently applied temperature.
Fig. 14 presents the distribution of bolt forces due to applied
temperature and mechanical loads. Each graph contains two
curves. The curve, denoted Failure onset, shows the force distri-
bution just before the rst fastener fails, when the elements in the
area between the ribs are assigned the behavior from Fig. 11. The
other curve shows the force distribution at the same applied load
level, but with linear elastic behavior assigned to all connector ele-
ments. The applied actuator load levels at the onset of rst bolt
failure are P
1
253:3 kN and P
2
130:6 kN, which gives the bend-
ing and twisting moments M
b
134:9 kN m and M
t
72:9 kN m
in the cross-section of the area between the ribs. These loads are
1.43 times higher than the dimensioning loads for ultimate failure
when Saab in-house methods for dimensioning are used.
It is evident that the effects of thermally induced loads cannot
be neglected. It can be seen from Fig. 14 that a linear elastic behav-
ior assigned to the fasteners leads to an overestimation of the max-
imal bolt loads at the onset of the rst fastener failure which
therefore is conservative. It is also noticed that several bolts in bolt
groups Skin-Splice and Spar-Splice approach their corresponding
ultimate failure force at 30 kN and 17 kN, respectively. Most of
the bolts in bolt group Skin-Spar are still in the linear elastic
regime, as anticipated.
A slight increase in the applied load leads to the rst ultimate
bolt failure, at bolt nr 16 and 32 in rows 1 and 2 at the edge of
the Skin-Splice bolt group, see Fig. 12(b). Fig. 15 shows the force
redistribution in the Skin-Splice and the Spar-Splice joints. As a re-
sult of loss of load carrying capacity in the failed bolts, the forces
are redistributed to the neighboring fasteners, i.e. the fasteners
close to the end of each row of Skin-Splice bolt group and in the
Spar-Splice joint, especially in row 4. The forces in the fasteners
at the very end of rows 3 and 4 in Skin-Splice bolt group however,
tend to decrease. The reason for this is that the loss of a load path
at the end of rows, 1 and 2 causes the load to redistribute towards
the middle of the splice, thus decreasing the transferred load
through the end of the rows 3 and 4.
Further progression of the fastener failure in the Skin-Splice
bolt group is shown in Fig. 16. It can be seen from Fig. 16(a) that
at this point, three fasteners in each of the rows 1 and 2 have com-
pletely failed and that two more are on their way to fail. The loads
in the fasteners at the end of rows 3 and 4 have decreased further.
The effects of load redistribution to other fasteners is more pro-
nounced. It is expected that the amount of total fastener shear load
transferred over the splice joint is the same before and after the
progressive fastener failure. However, there are several factors that
can disturb this equilibrium. One is that the stabilization option
used to damp the local instabilities dissipates some of the total en-
ergy and consequently can lower the transferred forces. This effect
is however negligible in this case, since the dissipated energy by
stabilization is conrmed to be very low, under 1% of the total
0 20 40 60 80 100 120
0
500
1000
1500
2000
2500
3000
3500
4000
Bolt nr
F

[
N
]
SkinSpar bolt group
(a) Skin-Spar bolt group, 1 bolt row in x-direction.
0 16 32 48 64
0
500
1000
1500
2000
2500
3000
3500
4000
Bolt nr
F

[
N
]
SkinSplice bolt group
row 1 row 2 row 3 row 4
(b) Skin-Splice bolt group, 4 bolt rows in
y-direction.
0 5 10 15 20 25 30 35 40
0
500
1000
1500
2000
2500
3000
3500
4000
Bolt nr
F

[
N
]
SparSplice bolt group
row 1 row 2 row 3 row 4 row 5 row 6 row 7 row 8
(c) Spar-Splice bolt group, 8 bolt rows in
z-direction.
Fig. 13. Distribution of resultant shear bolt force due to applied temperature of DT = 50 C in all three bolt groups studied.
Z. Kapidzic et al. / Composite Structures 109 (2014) 198210 207
strain energy. Another effect that can lower the total fastener load
is that the thermally induced forces decrease with fewer fasteners.
In other words, when ten of the outer fasteners are broken, some of
the energy stored in the joint by thermal expansion is released and
the remaining fastener forces decrease. Due to the loss of the fas-
teners at the end of rows 1 and 2, the third effect is that the outer
Spar-Splice starts twisting and joint row 4 starts transferring load
via the splice stiffener by contact and by axial bolt forces. The axial
force is not one of the components that contributes to the shear
resultant force and this effect is therefore not evident from
Fig. 16(b).
However, the inuence of shifted load transfer on load redistri-
bution to the fasteners outside the splice area can be seen in
Fig. 17. Due to the loss of Skin-Splice fasteners the load is redistrib-
uted to Spar-Splice joint row 4, which leads to increased shear
forces in the part of Skin-Spar joint, bolts 5063.
0 20 40 60 80 100 120
0
0.5
1
1.5
2
2.5
3
3.5
x 10
4
Bolt nr
F

[
N
]
SkinSpar bolt group
Linear elastic
connector
Failure onset
(a) Skin-Spar bolt group.
0 16 32 48 64
0
0.5
1
1.5
2
2.5
3
3.5
x 10
4
Bolt nr
F

[
N
]
SkinSplice bolt group
row 1 row 2 row 3 row 4
(b) Skin-Splice bolt group.
0 5 10 15 20 25 30 35 40
0
0.5
1
1.5
2
2.5
x 10
4
Bolt nr
F

[
N
]
SparSplice bolt group
row 1 row 2 row 3 row 4 row 5 row 6 row 7 row 8
(c) Spar-Splice bolt group.
Fig. 14. Distribution of resultant shear bolt force due to applied temperature of DT = 50 C and mechanical bending/twisting loads at rst bolt failure onset.
0 16 32 48 64
0
0.5
1
1.5
2
2.5
3
3.5
x 10
4
Bolt nr
SkinSplice bolt group
row 1 row 2 row 3 row 4
Failure onset
2 bolts failed
(a) Skin-Splice bolt group.
0 5 10 15 20 25 30 35 40
0
0.5
1
1.5
2
2.5
x 10
4
Bolt nr
SparSplice bolt group
row 1 row 2 row 3 row 4 row 5 row 6 row 7 row 8
(b) Spar-Splice bolt group.
F

[
N
]
F

[
N
]
Fig. 15. Distribution of resultant shear bolt force due to applied temperature of DT = 50 C and mechanical bending/twisting loads shortly after the rst bolt failure.
208 Z. Kapidzic et al. / Composite Structures 109 (2014) 198210
4. Conclusions
Local behavior of a bolted composite-aluminum shear lap joint
is modeled using PDM in the composite plate. The resulting force
displacement characteristics correlate reasonably well with the
experimental results found in the literature. It is concluded that
PDM, despite its shortcomings, can be used to determine the
behavior of composite failure in bearing failure mode. The initia-
tion of composite damage is observed locally at low load levels,
but the effect of the damage on the forcedisplacement curve is
not present until approximately half the maximal load. This behav-
ior should be carefully considered in the design situations, when
load levels for permanent deformations and ultimate failure are
determined. Mixing of aluminum and composite materials in
structures gives rise to effects that might be absent otherwise.
One of the effects is that thermally induced loads arise in hybrid
assemblages as a result of the difference in thermal expansion
properties of the materials. It is shown that the fastener forces
caused by temperature difference are of signicant magnitude
and should be considered in the design of hybrid structures. It is
also noticed that different material failure behaviors affect the lo-
cal stiffness of bolted joints and consequently the load distribution
in the structure. Because of the size of the numerical problem, solid
modeling including failure behavior is not suitable for modeling
larger structures including many joints. Some simplications are
necessary and in this work connector elements are used to repre-
sent the fasteners and the behavior that they exhibit. The force
displacement response of connector fasteners is determined from
a local solid fastener model including PDM for the composite plate.
In this study, the important effects of bolt hole clearance were not
included. However, connector elements do have a capability to in-
clude these kinds of behavior and this should be further looked
into. Also, the effects of by-pass loads on the local forcedisplace-
ment characteristics of hybrid joints need to be further studied.
The fastener load distribution due to bending/twisting and ther-
mally induced loads in a large hybrid wing box is studied. It is
shown that considering the non-linear response of the joints re-
sults in less conservative load distributions in typical wing struc-
tures on ultimate failure load level. On operational load levels,
the composite damage induced by repeated loading, might lead
to bolt hole widening and consequently, to load redistribution. This
phenomenon should be studied further.
Acknowledgments
The authors would like to acknowledge support from the HY-
BRIS-project, Swedish Defense Material Administration, and Saab
AB who funded the work presented in this article.
References
[1] Orici AC, Herszberg I, Thomson RS. Review of methodologies for composite
material modelling incorporating failure. Compos Struct 2008;86(1-3):
194210.
[2] Camanho PP, Mattews FL. Stress analysis and strength prediction of
mechanically fastened joints in FRP: a review. Compos A: Appl Sci Manuf
1997;28(6):52947.
[3] Dano ML, Gendron G, Picard A. Stress and failure of mechanically fastened
joints in composite laminates. Compos Struct 2000;50(3):28796.
[4] Dano ML, Kamal E, Gendron G. Analysis of bolted joints in composite
laminates: Strains and bearing stiffness predictions. Compos Struct
2007;79(4):56270.
[5] Xiao Y, Ishikawa T. Bearing strength and failure behavior of bolted composite
joints (part II: modeling and simulation). Compos Sci Technol 2005;65(7
8):103243.
[6] Olmedo A, Santiuste C. On the prediction of bolted single-lap composite joints.
Compos Struct 2012;94(6):21107.
[7] Hhne C, Zerbst A-K, Kuhlmann G, Steenbock C, Rolfes R. Progressive damage
analysis of composite bolted joints with liquid shim layers using constant and
continuous degradation models. Compos Struct 2010;92(2):189200.
[8] McCarthy CT, McCarthy MA, Lawlor VP. Progressive damage analysis of multi-
bolt composite joints with variable bolt-hole clearances. Composites: Part B
2005;36(4):290305.
[9] Linde P, Pleitner J, De Boer H, Carmone C. Modelling and simulation of ber
metal laminates. In: ABAQUS users conference; 2004.
[10] Camanho PP, Mattews FL. A progressive damage model for mechanically
fastened joints in composite laminates. J Compos Mater 1999;33:224880.
0 16 32 48 64
0
0.5
1
1.5
2
2.5
3
3.5
x 10
4
Bolt nr
F

[
N
]
SkinSplice bolt group
row 1 row 2 row 3 row 4
Failure onset
10 bolts failed
(a) Skin-Splice bolt group.
0 5 10 15 20 25 30 35 40
0
0.5
1
1.5
2
2.5
x 10
4
Bolt nr
F

[
N
]
SparSplice bolt group
row 1 row 2 row 3 row 4 row 5 row 6 row 7 row 8
(b) Spar-Splice bolt group.
Fig. 16. Distribution of resultant shear bolt force due to applied temperature of DT = 50 C and mechanical bending/twisting loads after failure of 10 fasteners in Skin-Splice
bolt group.
0 20 40 60 80 100 120
0
0.5
1
1.5
2
2.5
3
3.5
x 10
4
Bolt nr
F

[
N
]
SkinSpar bolt group
Failure onset
10 bolts failed
Fig. 17. Distribution of resultant shear bolt force in Skin-Spar bolt group due to
applied temperature of DT = 50 C and mechanical bending/twisting loads after
failure of 10 fasteners in Skin-Splice bolt group.
Z. Kapidzic et al. / Composite Structures 109 (2014) 198210 209
[11] Zhang J, Rowland J. Damage modeling of carbon-ber reinforced polymer
composite pin-joints at extreme temperatures. Compos Struct
2012;94(8):231425.
[12] Ireman T. Three-dimensional stress analysis of bolted single-lap composite
joints. Compos Struct 1998;43(3):195216.
[13] Rosales-Iriarte F, Fellows NA, Durodola JF. Failure prediction in carbon
composites subjected to bearing versus bypass loading. J Compos Mater
2012;46(15):185978.
[14] Egan B, McCarthy CT, McCarthy MA, Gray PJ, Frizzell RM. Modelling a single-
bolt countersunk composite joint using implicit and explicit nite element
analysis. Comput Mater Sci 2012;64:2038.
[15] Chishti M, Wang CH, Thomson RS, Orici AC. Numerical analysis of damage
progression and strength of countersunk composite joints. Compos Struct
2012;94(2):64353.
[16] Ireman T, Ranvik T, Eriksson I. On damage development in mechanically
fastened composite laminates. Compos Struct 2000;49(2):15171.
[17] Chishti M, Wang CH, Thomson RS, Orici AC. Experimental investigation of
damage progression and strength of countersunk composite joints. Compos
Struct 2012;94(3):86573.
[18] Weyer S, Hooputra H, Zhou F. Modeling of self-piercing rivets using fasteners
in crash analysis. In: ABAQUS users conference; 2006.
[19] Gray PJ, McCarthy CT. A global bolted joint model for nite element analysis of
load distributions in multi-bolt composite joints. Composites: Part B
2010;41(4):31725.
[20] Gray PJ, McCarthy CT. A highly efcient user-dened nite element for load
distribution analysis of large-scale bolted composite structures. Compos Sci
Technol 2011;71(12):151727.
[21] Ekh J, Schn J. Finite element modeling and optimization of load transfer in
multi-fastener joints using structural elements. Compos Struct
2008;82(2):24556.
[22] Wang AJ, Chung KF. Advanced nite element modelling of perforated
composite beams with exible shear connectors. Eng Struct
2008;30(10):272438.
[23] MIL-HDBK-17-3F. Military handbook, polymer matrix composites. U.S.
Department of Defense; 2000.
[24] Abaqus Analysis Users Manual, Simulia; 2011. <www.simulia.com>.
[25] Tserpes KI, Labeas G, Papanikos P, Kermanidis Th. Strength prediction of bolted
joints in graphite/epoxy composite laminates. Composites: Part B
2002;33:5219.
[26] Schn J. Coefcient of friction for aluminum in contact with carbon ber epoxy
composite. Tribol Int 2004;37(5):395404.
[27] Hashin Z. Failure criteria for unidirectional bre composites. J Appl Mech
1980;47:32934.
[28] Chang FK, Chang KY. A progressive damage model for laminated composites
containing stress concentrations. J Compos Mater 1987;21(9):83455.
[29] Chang FK, Lessard LB. Damage tolerance of laminated composites containing
an open hole and subject to compressive loadings: part I Analysis. J Compos
Mater 1991;25(1):243.
[30] Hahn HT, Tsai SW. Nonlinear elastic behavior of unidirectional composite
laminates. J Compos Mater 1973;7(1):10218.
[31] Hashin Z, Rotem A. A fatigue failure criterion for ber reinforced materials. J
Compos Mater 1973;7:44864.
[32] Tan SC. A progressive failure model for composite laminates containing
openings. J Compos Mater 1991;25(5):55677.
[33] Huth H. Inuence of fastener exibility on the prediction of load transfer
and fatigue life for multiple-row joints. Fatigue in mechanically
fastened composite and metallic joints. ASTM STP 927, Philadelphia; 1986.
p. 22150.
210 Z. Kapidzic et al. / Composite Structures 109 (2014) 198210

You might also like