You are on page 1of 54

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/241652953

A Numerical and Experimental Study of Woven Composite Pin-Joints

Article  in  Journal of Composite Materials · June 2000


DOI: 10.1177/002199830003401204

CITATIONS READS
95 446

3 authors, including:

Fabrice Pierron Michel Grédiac


University of Southampton Université Clermont Auvergne
349 PUBLICATIONS   7,551 CITATIONS    343 PUBLICATIONS   7,398 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Temperature Elevation in an Instrumented Phantom Insonated by B-Mode Imaging, Pulse Doppler and Shear Wave Elastography View project

3D Printing Technology in the Fabrication of Sand Molds and Cores for Shape Casting of Light Alloys View project

All content following this page was uploaded by Michel Grédiac on 10 November 2014.

The user has requested enhancement of the downloaded file.


A NUMERICAL AND EXPERIMENTAL STUDY OF
WOVEN COMPOSITE PIN-JOINTS
F. PIERRON∗ , F. CERISIER∗ and M. GREDIAC∗∗
∗ Department of Mechanical and Materials Engineering
Ecole des Mines de Saint-Etienne
158, cours Fauriel
42023 Saint-Etienne Cedex 02, France

∗∗ LERMES
Université Blaise Pascal Clermont II
24, avenue des Landais, BP 206
63174 Aubière Cedex, France

Submitted to Journal of Composite Materials


Revised version
March 29, 2006

1
Abstract

A numerical and experimental study was carried out to determine the stiffness

and the bearing strength of bolted woven composite joints. The main objective

was to investigate the possibility of predicting the properties of the joint from the

properties of the material measured with standard tests. A refined finite element

model was developed in which the nonlinearities due to both the material and the

contact angle between the pin and the hole were taken into account. Particular

attention was paid to account for the influence of the clearance which has been

shown to be very significant. In conclusion, good agreement beetween experimental

results and numerical predictions has been obtained.

1 INTRODUCTION

Many papers deal with the failure analysis of bolted composite joints. This is mainly

due to the fact that joining composite structural components often requires mechan-

ical fasteners which ultimate mechanical response is difficult to predict. Contrary to

many metallic structural parts, for which the strength of the joints is mainly governed

by the shear and the tensile strengths of the pins, composite joints present specific

failure modes due to their heterogeneity and anisotropy. Three main basic failure

modes are often described in the literature: bearing, net-tension and shear-out ([1, 2]

for instance). In the present work, the specimen geometry is adjusted so that the

bearing mode is enhanced. The main reason is that net-tension and shear-out can

be avoided by increasing the width and the end-distance of the structural part for a

given thickness. On the contrary, the bearing strength involves local effects which are

mainly influenced by the material properties and the contact area between the pin

and the hole. As a result, this type of failure cannot be avoided by any modification

of the geometry of the structural part for a given thickness and pin diameter and

can be considered an intrinsic property of the joint for given constitutive material,

stacking sequence, diameter and clearance.

The final objective of the present study is to develop a model for failure prediction

2
of the joints. However, many such models already exist in the literature [3, 4, 5, 1],

for instance, though generally the materials under study are carbon/epoxy laminated

composites. Nevertheless, it has appeared to the present authors that a common

weakness of the above papers is the lack of thorough experimental validation of the

numerical finite element models used to calculate the stress and strain fields around

the hole. Indeed, Hamada et al. [3] use a linear elastic finite element model with a

rather crude contact model for the pin/composite interface and by using a Yamada-

Sun failure criterion together with a caracteristic length approach [6], the strengths

of the pin-joints are predicted. Nevertheless, no experimental validation of the model

is given and their good agreement between predicted and measured failure loads does

seem rather fortuitious. A more rigorous approach is presented by Hung et al. [5] who

use a cumulative damage model to predict the response of the pinned and clamped

joints. An experimental validation of the finite element model is given in the form of

comparisons between strain gauge measurements and numerical results. These seem

rather satisfactory. However, the difference between measured and calculated strains

increase when the gauges get closer to the hole though the closest gauge remains one

hole radius away from the hole edge, which is too far away since failure occurs much

closer to the hole edge. Moreover, clearance between the pin and the composite is not

considered and the behaviour of the joint seems linear up to failure because of the

particular quasi-isotropic lay-up. Therefore, the validation presented in this paper

does not seem to be particularly conclusive.

The objective of the present paper is to study the behaviour of woven glass fibre

epoxy pin-joints both numerically and experimentally, with particular attention given

to the sensitivity of the model to different parameters (clearance, friction, non-linear

material behaviour). This work is aimed at a thorough experimental validation of the

numerical results so that the model can subsequently be used for the development of

a relevant predictive failure model.

3
2 TESTING
2.1 Joint testing configuration

The test specimen is shown in Figure 1. The width w and the end length e have been

chosen so that bearing failure occurs. The value of the length l has been chosen high

enough so that Saint-Venant’s effect due the grip is avoided. The diameter of the pin

is 16 mm. The diameter of the hole is between 16.1 and 18 mm, so as to observe the

influence of the clearance. This range of clearances corresponds to classical values

from the mechanical construction standards [7]. The specimens were cut from the

panels using a diamond coated blade. The hole is drilled in the composite plate with

a silicon carbide mill. Care was taken during the drilling process to avoid damage on

the back surface of the specimens by using wasted wedges. The material is a woven

glass fibre cloth embedded in an epoxy resin. The panels have been autoclaved from

a 7781/XE85AI Hexcel prepreg. The resin is a 120◦ C curing system. The final fibre

volume fraction lies around 45%. The panels were 6.5 mm [±45]12s . The reason for

this is that the present study is related to a wider project involving the design of a

composite structural component for the railway industry [8, 9] for which the fibres

orientations are ±45 ◦ .

A double lap fixture shown in Figure 2 was used to perform the tests. All parts

of the fixture are made of high strength steel. The influence of the clamping pressure

was not studied in the present work for the sake of simplicity, therefore, the bolt was

not tightened. The lower edge of the specimen was clamped with hydraulic grips to

control the pressure fixed at 4 MPa according to the manufacturer recommendations.

The magnitude of the applied load is measured by a load cell mounted on the testing

machine. The displacement of the bottom grip is also recorded during the tests. The

cross-head rate was 1 mm.mn−1 .

Preliminary testing has been performed on specimens with clearances of 0.1 mm,

0.5 mm, 1 mm, 1.5 mm and 2 mm. The idea was to check the influence of the

clearance on the joint behaviour as well as to have a first set of data to start the

modelling. The results are given in Figure 3 as load-displacement curves. As can

4
be seen on this figure, the behaviour of the joint is very dependent on the clearance

between the pin and the hole, particularly the failure load that is reduced by about

30% when the clearance varies from 0.1 mm to 2 mm. On the contrary, the joint

global stiffness does not vary very much (slopes of the curves in Figure 3), suggesting

that the influence of the clearance is concentrated locally near the hole. This result

is not surprising but implies that the prediction of the joint strength as a function

of the clearance will necessitate the development of a model able to describe local

effects of load distribution on the stress and strain fields near the hole. Many such

models have been proposed over the years, both analytical and numerical [10], but

the present authors have noticed that there was a lack of experimental validation of

these models. Nevertheless, such validations are essential to demonstrate the validity

of the failure prediction models that use as input values the computed stresses and

strains near the hole. The rest of the present paper is dedicated to the development

and validation of finite element calculations aimed at modelling the behaviour of the

pin-joints described above.

2.2 Material characterization

The first step to develop a finite element model of the joint is to characterize the

material behaviour. In order to do so, several mechanical tests have been performed.

2.2.1 Tension

Tensile tests have been performed on 6 rectangular specimens with fibres aligned

with the specimen long axis (0◦ ), according to the ASTM D3039-76 standard [11].

The specimens dimensions were 225 mm by 25.4 mm by 2.69 mm. Composite tabs

of 2.7 mm thickness cut from the tested panel have been used so that the final

gauge length was 150 mm. Each specimen has been instrumented by a 0/90 rosette

(HBM 6/120XY13) fixed in the centre of both faces of the specimen so that the

effect of parasitic bending could be eliminated. The grips used are MTS servohy-

draulic grips. The clamping pressure has been selected to 4 MPa according to the

manufacturer recommendations. The cross-head speed was fixed at 0.5 mm.mn−1 .

5
The stress strain curves are nearly linear up to failure. A limited non-linearity

appears towards the end of the life of the specimen, which is well known for fabric

reinforced composites because of the fibres alignment. Longitudinal modulus and

Poisson’s ratio have been calculated from the linear part of the curve. The results

are as reported in Table 1. It can be seen that no significant differences have been

noticed between the warp and weft directions.

2.2.2 Shear

The in-plane shear response of the material has been measured using the 45◦ off-

axis tensile test, following the ASTM D3518 standard [11]. The dimensions of the

specimens were equal to that of the tension tests. The shear specimens were equiped

as the tension ones. A typical shear response is represented in Figure 4. As can be

seen, the response is highly non-linear. The implications of this will be discussed

later in the modelling section. In terms of modulus and failure stress, the results are

reported in Table 1. It is to be noted that the in-plane shear failure stress is only an

approximation of the shear strength [12].

2.2.3 Compression

Compression tests have been performed using a slightly modified version of the ASTM

D3410-87 standard [13]. The specimens were 112 by 12 by 2.69 mm. They were

equipped with 50 mm long and 2 mm thick tapered tabs from the same material

as the tested one. The values found for the compressive failure stress was 365 MPa

in the warp direction and 355 MPa in the weft direction. These values were much

lower than that reported in the manufacturer technical data sheet (values reported

in Table 1). It is to be noted however that the behaviour of a compression specimen

is basically a structural one because of the instability driven failure mode [14, 15].

Therefore, a first approximation of the compressive strength will be taken as equal to

the tensile strength which is consistent with the values measured by Hexcel (Table 1).

Now that the material properties have been evaluated, the modelling of the joint

can be performed. This is the object of the next section.

6
3 FINITE ELEMENT MODELLING

In order to calculate the stress distribution around the hole of the composite joint,

a finite element model has been developed. The package used for this analysis is

ABAQUS 5.5. In order to reduce the number of parameters, only the two extreme

clearance values have been considered, that is 0.1 mm and 2 mm. The specimen

dimensions are that of Figure 1. The mesh used for the present study is represented

in Figure 5. The elements used for this study are eight-noded biquadratic plane stress

CPS8 elements.

In order to assess the influence of certain parameters on the joint response, several

levels of modelling have been considered. First, a linear elastic model has been

developed. The boundary conditions are as follow. The right hand side of the model

of Figure 5 is clamped and a pressure with a sine profile is applied on the left hand

side of the hole, simulating the pressure of the bolt on the composite hole. These are

classical boundary conditions used in the literature [16].

The second model considered is that taking the contact between the bolt and

the composite into account. In order to simulate this contact, a rigid surface has

been modelled to represent the bolt. ABAQUS uses an iterative surface interaction

technique to solve this problem. It is not the scope of the present paper to describe

this numerical approach. All details can be found in [17]. The local meshes used for

the two clearance values are represented in Figure 6. Within this model, the influence

of the friction coefficient has also been investigated.

Finally, because of the strongly non-linear in-plane shear behaviour of the ma-

terial, the contact model previously described has been refined by introducing this

material nonlinearity. In order to do so, the shear response in Figure 4 has been mod-

elled by a fifth order polynomial, assuming that the response is the same when the

shear stress changes sign. This is implemented by an ABAQUS UMAT procedure

programmed with the specific ABAQUS language, with the following stress-strain

7
relationship (for εs ¿0):

σs = 9.081014 ε5s − 1.411014 ∗ ε4s + 8.881012 ∗ ε3s


(1)
11
−2.9010 ∗ ε2s 9
+ 5.4910 ∗ εs

However, because of the debonding of the strain gauges after a 6% shear strain (see

Figure 4), the following part of the stress strain curve has been modelled by a straight

line in the continuity of the measured stress-strain curve:

σs = 4.14108 ∗ εs + 4.67107 (2)

The next step is now to validate the above models experimentally and to find out

which level of modelling is necessary to represent the pin-joint response correctly.

This is the objective of the next section.

4 EXPERIMENTAL VALIDATION
4.1 Global stiffness

The first validation that can easily be performed concerns the global stiffness of the

structure, that is, the load measured by the load cell divided by the displacement

of the pin. For a clearance of 0.1 mm, Figure 7 compares the global stiffness of

the joint obtained from the linear elastic model, the contact model (no friction)

and the contact plus the shear nonlinear material behaviour model. As can be seen

from these results, the material nonlinearity is essential to describe the response

of the joint. Moreover, the implementation of the contact elements does not seem

to change the joint stiffness significantly. This is consistent with the fact that the

influence of the pin contact pressure should be limited to the local stress and strain

fields. However, even for this 0.1 mm clearance, the classically assumed sinusoidal

pressure distribution is not representative of the action of the bolt on the composite.

Indeed, Figure 8 shows the pressure distribution calculated from the finite element

contact model compared to the sinusoidal distribution and it can be seen that the

calculated pressure is significantly different from a sine function. This will lead to

local differences in the stress and strain fields that will affect the strength prediction

model, as already mentionned by Murthy et al. [16], for instance. Another interesting

8
feature is the variation of the contact angle as a function of the load for the two

clearance values. This is reported in Figure 9 obtained by recording the contact

angle at each step of the iterative solving process. As can be seen, an increase in the

clearance results in a smaller contact area, which explains the lower failure stress for

the 2 mm configuration, as expected.

Finally, when considering Figure 7, it can be seen that there is still a difference

between the model and the experimental response of the joint. This is certainly due

to the fact that the displacement measured experimentally is that of the machine

cross-head and the displacement calculated from the finite element model is the

relative displacement of the pin with respect to the clamped end of the composite

plate. Since the fixture is not infinitely stiff, some deformation of the fixture is

expected. This has been checked and measured using a dial gauge fixed on the pin.

Since on the test machine used for this study, it is the bottom part of the fixture

that moves (see Figure 2), i.e the one with the grip, the displacement of the pin

relatively to the machine should be zero. Any non-zero displacement of the pin will

indicate deformation of the upper part of the fixture, including the pin itself which

can deform.

The experimental curve from Figure 7 has therefore been corrected to take this

fixture deformation into account. This was done by substracting the pin displacement

to the bottom grip displacement. This is represented in Figure 10. It can be seen that

the model fits the bold line very well at the beginning of the curve but that there is

still a significant difference towards the end. Since any extra deformation of the upper

part of the fixture is now accounted for by the use of the dial gauge, the explanation

has been sought on the lower part of the fixture. Using a second dial gauge measuring

the relative displacement of the top of the grip and the specimen, it was found that

some sliding in the grips was present. It was sufficient to account for the few tenths

of millimeters difference between the model and the corrected experimental results

of Figure 10. In order to fully check the model, it would be necessary to use a device

that would measure the relative displacement between the pin and a fixed reference

9
on the specimen near the grips. However, the validation performed here was thought

sufficient to give some initial confidence in the finite element modelling. Since what is

of interest to predict the failure load is the local stress and strain distributions around

the hole, it was decided to carry on the validation by using local strain measurements.

4.2 Local validation

The local validation of the finite element model has been performed by two different

means.

4.2.1 Strain gauges

First, two specimens - one for each clearance value - have been equipped with two

back-to-back Vishay Micromeasurement WA-13-060WR-120 rosettes each and tested.

The center of the rosettes was positioned 2.5 mm away from the hole edge and the

gauge length was 1.5 mm. Again, it should be noted that the gauges are here much

closer to the hole edge than in [18] where they are positioned at least one hole

diameter away from the hole edge, that is, away from the failure initiation zone. The

specimen with the 0.1 mm clearance has first been tested twice up to 9 kN while being

taken out of the rig, turned around and fitted back into the fixture between the two

tests. This was performed to check whether significant and reproduceable differences

existed between the strains on the two specimen faces. The results are reported in

Figure 11 for the longitudinal strain. It can be seen that more than 100% strain

differences are recorded between the two faces. Moreover, it is the same specimen

face that exhibits the lower strain. This is consistent with other observations on

Iosipescu shear specimens where it was shown that the cause of this phenomenon is

in-plane Saint-Venant effects due to the local load heterogeneity that results from

uneven pin/composite contact [19, 20], as sketched in Figure 12. Therefore, proper

interpretation of the results must be ensured by averaging out the strains over the two

faces, as demonstrated in [19]. This is the procedure adopted to produce to following

results. It should be noted that such a procedure is rarely applied in practice in

other published studies ([18], where only one specimen face is strain gauged) and

10
discrepancies between model and experiment could well arise from this point. To

the present authors opinion, this is one of the significant contributions of the present

paper.

In order to compare the computed results to the experimental ones, the finite

element strains have been calculated over a certain number of elements to simulate

the strain gauge. Also, in order to check whether friction had any effect on the local

strain distributions, models with different friction coefficients have been developed for

both pin-joint configurations. The nominal friction coefficient input in the model has

been evaluated by sliding a metallique bar on a composite plate and by measuring the

sliding angle. This experiment gave a friction coefficient of about 0.3. The results are

reported in Figures 13 and 14. The first observation that arises from these results is

that the influence of friction is very dependent on the clearance value. Indeed, for the

0.1 mm clearance specimen, the difference between a 0 and 0.3 friction coefficient is

very important (nearly a 100% strain difference for a 15 kN load). On the other hand,

when the clearance value increases, this influence decreases and for a 2 mm clearance

value, there is only a very small difference between the responses obtained with and

without friction. This result is important since although the influence of friction has

been numerically pointed out by several authors [21, 22, 23], the mixed experimental

and numerical studies very often neglect the influence of friction, even for clearance-fit

configurations [3, 2, 5]. In the case of the work by Hung and Chang [5], a validation

using strain gauges has also been performed and good agreement between model and

experiment was found but the closest gauge was positioned one hole radius away from

the hole edge (details in [18]). It is interesting to note (Figure 17 of reference [5])

that the strains from the first two gauges, positioned 3 and 5 radius away from the

hole edge were exactly predicted by the model but that for the last gauge, positioned

one radius away from the hole edge, the experimental strains were lower than that

given by the model, exactly as in Figure 13 when friction is not taken into account.

Though this difference remained small in [5] (about 20%), it seems that if the strain

measurement location was moved up further towards the hole edge, where the failure

11
is actually initiated, results such as that of Figure 13 would have been found (in the
1
present work, the strain gauge is positioned 8 th of the radius away from the hole

edge). In the other reference cited above ([3]), no experimental validation of the local

strain distribution obtained from the finite element model is provided.

Finally, it can be said that the results of Figures 13 and 14 are very satisfactory,

even more so since the strain gauges have been positioned very close to the hole edge

where high strain gradients are present. It is important to note that if only one gauge

had been used or if friction had been neglected, such a thorough validation would

have been impossible.

4.2.2 Whole-field measurements

The previous validation has been very successful but it is only a very punctual verifi-

cation. In order to spread this validation over a larger field, a whole-field displacement

measurement technique has been used. The technique is based on the deformation

of a unidirectional grid followed by a CCD camera [24]. The phase of the grid is

measured before and after specimen loading and the difference between these two

phase fields is related to the inverse displacement field, u−1 (r), which, for small dis-

placements, is equal to −u(r). The image acquisition and processing is ensured by

the Frangyne software developed in-house and based on phase-stepping [25]. The

grid used in the present study is a Mécanorma Normatex 3121 grid of 610 µm pitch.

The grid is bonded on the composite using the VPAC-1 glue from Vishay Micromea-

surement but before bonding, a thin layer of white paint is spread onto the composite

specimen to ensure good contrast of the grid. Details of the procedure can be found

in [9, 26, 27, 28]. In order to be able to see the grid above the pin-head, the fixture

was modified by machining the flanges in this area.

The direct results from the experiments are the longitudinal displacements for the

two clearance specimens. These results are presented in Figure 15. It can be seen that

there is an excellent correlation between the calculated and measured displacement

fields, and also that the shapes of the contour lines are somewhat different between

the 0.1 mm and the 2 mm clearances. From these displacement maps, it is possible

12
to obtain the longitudinal strain by differentiating the data numerically with respect

to the x2 coordinate. However, this procedure requires some spatial smoothing to

obtain reasonable results. In the present case, the best least-square linear fit over 5

pixels has been used to derive the strains. The resulting strain maps are represented

in Figure 16. What can be said from these results is that there is good correlation

with the calculated fields, though significant noise is present on the experimental

strains because of the differentiation process. The information given by these results

is complementary to that from the strain gauges. Indeed, in the present case and

contrariwise to the strain gauges procedure, the local strain values close to the hole

edge are not precisely measured, though this would be possible by using a finer grid

(this work in on-going at Ecole des Mines). Nevertheless, the present procedure

enables the measurement of the whole longitudinal displacement and strain fields

around the hole, which strain gauges cannot perform. Therefore, both the validation

results enable to put good confidence in the finite element model and particularly in

the difficult process of modelling the boundary conditions. This step towards the final

goal of strength prediction appears essential to the present authors, all the more since

such validation is often absent or incomplete in other studies from the literature. It

is the authors opinion that there is no need in developing and implementing complex

damage models if the computations are not thoroughly validated by experimental

work.

5 STRENGTH PREDICTION

Once the finite element model has been validated, the last step is to use the calculated

strain or stress fields together with some damage/failure model to predict the failure

load of the joint, since this is what is of interest to the designers.

In order to reach the above objective, additional experimental work has been

carried out to investigate the failure of the specimens. A classical force displace-

ment curve is represented in Figure 17 together with the visible specimen damage

associated with the first and the final load drops. The first load drop (failure 1) is

13
associated with the appearance of a swollen V-shaped zone under the contact point.

This phenomenon, also described by Wang et al. [18], is caused by the dramatic

propagation of interlaminar cracks. After this event, the joint bears further load un-

til the final failure (failure 2) occurs as a complete crushing of contact zone. Clearly,

alhough the joint carries on bearing load after failure 1, the designer must ensure

that the joint never reaches failure 1 since the joint cannot recover its properties

after that point (irreversible damage). Therefore, the present study will concentrate

on the part of the curve leading to failure 1.

Although failure 1 is due to interlaminar cracks, it is certainly the accumulation

damage due to in-plane stresses that is the cause of the interlaminar cracks, as already

suggested by Wang et al. [18]. In order to investigate this point, four specimens for

each clearance value have been tested up to failure. These joints have been equipped

with an acoustic emission sensor (type EPA-Dunegan S9220) connected to an EPA-

Dunegan 3000 acquisition board (100 channels). The system has been calibrated

according to the ASTM E1118-89 standard [29]. The results are reported in Table 2

and typical curves are given in Figures 18 and 19. The first thing that can be

seen in Table 2 is that there is not much scatter on the failure forces for the four

tested specimens. This justifies the fact that no additional data have been sought.

Concerning Figures 18 and 19, the main observation is that the acoustic emission

is much more important for the 0.1 mm clearance specimen that for the 2 mm one.

Also, for the latter, the start of the events is very close to the failure point. This

phenomenon may well be caused by the fact that there is much more sliding between

the pin and the hole for the 0.1 mm specimen (see also the influence of friction in the

previous section). Therefore, it might be misleading to consider the emitted noise

as representative of damage. In any case, these acoustic emission results show an

important increase of noise near the first load drop. This suggests that significant

in-plane damage is to be sought near this first load drops.

In order to check for significant in-plane damage before the ‘macroscopic’ failure

point (first load drop), and to identify what type of damage occurs (i.e. caused by

14
which stress component), two sets of three specimens per clearance value have been

tested. They were loaded up to a certain fraction of the average failure load of Table 2

and then, microscopic observation was carried out to see whether in-plane damage

was visible in the mid-plane of the specimens. The results are reported in Table 3.

For both configurations, significant damage is visible at around 90% of the failure load

and not detectable below 70%. The aspect of the visible damage is representative

of in-plane shear damage (whitening of the fibre tows), as can be seen in Figure 20.

Similar observation had been made on the 45◦ off-axis tensile specimens, confirming

that in-plane shear seems responsible for the specimen failure. It is important to

note that the photograph of Figure 20 was taken from a failed specimen. The extend

of the damage observed at 90% of the failure load is much lower.

As a summary of the above experimental study, it can be said that the damage

leading to the failure of the joint is suspected to be caused by in-plane shear and that

significant damage has been found only around 90% of the failure load. The next

step is now to compare these results to the finite element calculations.

The full finite element model (including friction contact and material nonlinearity)

was used to interpret the above results. The stress fields around the hole for the two

clearance values are given in Figures 21 and 22 in the orthotropy axes, respectively

for 25 and 17 kN, which corresponds to the values of Table 3 (i.e. 90% of the failure

load). The main result from these stress maps is that only the in-plane shear stress in

close to its critical value over a significant area (see Table 1). This confirms the fact

that the failure of the joint is caused by in-plane shear. Looking more closely at the

in-plane shear maps (Figures 23 and 24), it can be seen that for the 2 mm clearance,

the negative shear (under the contact zone) is significantly higher than the positive

one as for the 0.1 mm clearance, they are of the same order of magnitude. Clearly,

two phenomena are in competition here, compression load at the contact point due

to the local pressure of the pin but also tension at 90◦ from the contact point due

to the ovalization of the hole, both causing shear in the material axes because of

the ±45◦ fibre orientation. In the case of the 2 mm clearance specimen, the fact

15
that the contact load is more concentrated than for the 0.1 mm one results in higher

shear stress under the contact point, but for the 0.1 mm specimen, the positive and

negative shear stress values are similar. This is surprising since no trace of damage

has been found in the positive shear zone in the failed specimens, indicating that only

the negative shear zone is important for failure. There are two possible explanations

the present authors can think of. First, the finite element model does not well

represent what happens at 90◦ from the contact point. Since no precise validation

using strain gauges has been performed in this area, it is possible that the finite

element modelling is not accurate in that area, though the full field measurements

were quite reassuring. The second possible cause is related to the local Saint-Venant

effects described in the previous section. Indeed, it was shown that important strain

differences existed between the two specimen faces because of local contact defects,

particularly for the 0.1 mm configuration. However, such differences are not to be

expected on the shear strain in the positive shear zone since this area is far from the

contact point. Therefore, it is probable that the damage is prematurely initiated on

the higher strain side of the specimen, at higher stress values than that of the 2D

map in Figure 23. This should be confirmed by localized microscopic observation.

Nevertheless, this feature is an important obstacle to the correct prediction of the

failure load of such joints.

Finally, the previous results have been used to predict the failure load of the two

types of joints. As a first approach, a very simple procedure is used. The maximum

stress criterion [30] is selected as the failure model. So, the maximum shear stresses

from the finite element results are compared to the in-plane shear strength, estimated

from the 45◦ off-axis tensile test (Table 1). When these values meet, the computed

force is recorded and compared to the experimental measurements. The results are

reported in Table 4. The predicted failure loads are slightly overestimated. This

again may be due to the Saint-Venant effect, as explained above. A 10% difference is

compatible with the recorded strain differences at failure, as can be seen in Figure 25.

Also, using the maximum stress value from the finite element maps does not take the

16
influence of the gradient into account (see [6]). Nevertheless, considering the very

basic failure model used here, the results are very satisfactory. This may be a hint

that instead of developing complex advanced failure models that industrial design

departments may not be able to use (cumulative damage etc. . . , see [5, 3]), it may

be worthwhile to put more effort into the correct mechanical modelling (boundary

conditions etc.) and the experimental validation and use basic failure criteria such

as that largely spread among the industrial designers (maximum stress, maximum

strain or Tsai-Wu, for instance).

6 CONCLUSION

The following conclusions can be drawn from the present study:

• The clearance between the pin and the hole has an important influence on the

failure load of pin-joints. Minimum clearance should be ensured according to

the assembly requirements.

• Local Saint-Venant effects due to pin-hole contact defects cause significant strain

differences between the two specimen faces. Strain gauges should be fixed on

the two specimen faces and the strains averaged to correctly validate the finite

element local strain calculations.

• The minimum level of finite element modelling required to describe the be-

haviour of the pin-joint considered here requires the use of friction contact and

shear nonlinearity.

• Thorough experimental validation of the finite element model is required to see

the influence of certain parameters of the model (friction, in particular).

• The failure of the present ±45◦ joints is caused by in-plane shear.

• Using the maximum stress criterion applied to the maximum stress value from

the finite element computations, the failure loads of the two types of joints

(clearances of 0.1 and 2 mm) has been predicted within 10% of the experimental

values.

17
• The present approach has been limited to a certain type of material and layups

and should be extended to other configurations for additional validation. Also,

the influence of the tightening of the bolt has been discarded. However, the pro-

cedure is compatible with design departments requirements where more complex

failure models cannot be implemented because of the very costly identification

procedures, particularly in industrial sectors where the added value is much

lower than that of the aerospace industry. The present study is a step towards

simplified design procedures for composite bolted joints.

18
References

[1] K. Hollmann. Failure analysis of bolted composite joints exhibiting in-plane

failure modes. Journal of Composite Materials, 30(3):358–387, 1996.

[2] C. L. Hung and F. K. Chang. Strength envelope of bolted composite joints under

bypass loads. Journal of Composite Materials, 30(13):1402–1435, 1996.

[3] H. Hamada, Z.-I. Maekawa, and K. Haruna. Strength prediction of mechanically

fastened quasi-isotropic carbon/epoxy joints. Journal of Composite Materials,

30:1596–1612, 1996.

[4] F. K. Chang, R. A. Scott, and G. S. Springer. Strength of mechanically fastened

composite joints. Journal of Composite Materials, 16:470, 1982.

[5] C. L. Hung and F. K. Chang. Bearing failure of bolted composite joints. Part II:

model and verification. Journal of Composite Materials, 30(12):1359–1399, 1996.

[6] J. M. Whitney and R. J. Nuismer. Stress fracture criteria for laminated com-

posite containing stress concentrations. Journal of Composite Materials, 8:253,

1974.

[7] R. Quatremer and J.P. Trottignon. Construction mécanique. Nathan, 1986. in

French.

[8] F. Cerisier, M. Grédiac, F. Pierron, and A. Vautrin. Design of a locomotive

transmission in composite materials. In Proceedings of European Symposium on

Design and Analysis, July 1996 in Montpellier, France, 1996.

[9] F. Cerisier. Conception d’une structure travaillante en composite et étude de ses

liaisons. PhD thesis, Université Jean Monnet de Saint-Etienne, France, 1998.

7th May 1998, in French.

[10] P.P. Camanho and F.L. Matthews. Stress analysis and strength prediction of

mechanically fastened joints in frp: a review. Composites Part A, 28A:529–547,

1997.

[11] ASTM D3039-76. Test method for tensile properties of fiber-resin composites,

1976. American Society for the Testing of Materials.

19
[12] F. Pierron and A. Vautrin. New ideas on the measurement of the in-plane

shear strength of unidirectional composites. Journal of Composite Materials,

31(9):889–895, 1997.

[13] ASTM D3410-87. Test method for compressive properties of unidirectional or

cross-ply fiber-resin composites, 1987. American Society for the Testing of Ma-

terials.

[14] J.C. Grandidier, G. Ferron, and M. Potier-Ferry. Microbuckling and strength in

long fiber composite: theory and experiments. International Journal of Solids

and Structures, 29:1753–1761, 1992.

[15] S. Drapier, J.C. Grandidier, and M. Potier-Ferry. Non linear numerical approach

of microbuckling. Composites Science and Technology, 1998. to appear.

[16] A.V. Murthy, B. Dattaguru, H.L.V. Narayana, and A.K. Rao. Stress and

strength analysis of pin-joints in laminated anisotropic plates. Composites Struc-

tures, 19:299–312, 1991.

[17] ABAQUS 5.5 user manual.

[18] H. S. Wang, C. L. Hung, and F. K. Chang. Bearing failure of bolted composite

joints. Part I: experimental characterisation. Journal of Composite Materials,

30(12):1285–1313, 1996.

[19] F. Pierron. Saint-Venant effects in the Iosipescu specimen. Journal of Composite

Materials, 32(22):1986–2015, 1998.

[20] F. Pierron. Experimental evidence of Saint-Venant effects in composite test-

ing. In European Conference on Composite Materials -Composite Testing and

Standardisation 4, pages 47–56, 1998. 31 August to 2 September in Lisbon,

Portugal.

[21] M.H. Hyer, E.C. Klang, and D.E. Cooper. The effect of pin elasticity, clear-

ance and friction on the stresses in a pin-loaded orthotropic plate. Journal of

Composite Materials, 21:190–206, 1987.

20
[22] L.J. Eriksonn. Contact stresses in bolted joints of composite materials. Com-

posite structures, 6:57–75, 1986.

[23] R.A. Naik and J.H. Crews. Stress analysis method for a clearance-fit bolt under

bearing loads. AIAA Journal, 24(8):1348–1353, 1985.

[24] Y. Surrel and B. Zhao. Moiré and grid methods: a ’signal processing’ approach.

In Photomechanics, Proceedings of Interferometry ’94, vol. SPIE 2342, pages

118–127, 1994. 16-20 May 1994 in Warsaw.

[25] Y. Surrel. Design of algorithms for phase measurements by the use of phase-

stepping. Applied Optics, 35(1):51–60, 1996.

[26] F. Cerisier, L. Dufort, M. Grédiac, F. Pierron, and Y. Surrel. Application d’une

méthode de grille à l’étude du contact boulon-trou dans une pièce composite. In

Photomécanique, pages 97–104, 1998. 14-16 April in Marne-la-Vallée, France.

[27] F. Pierron, E. Alloba, Y. Surrel, and A. Vautrin. Whole-field assessment of

the effects of boundary conditions on the strain field in off-axis tensile testing

of unidirectional composites. Composites Science and Technology, 58(12):1939–

1947, 1998.

[28] L. Dufort, M. Grédiac, Y. Surrel, and A. Vautrin. Applying the grid method

to the measurement of displacement and strain fields through the thickness of

a sandwich beam. In A. Vautrin, editor, Mechanics of sandwich structures,

Proceedings of Euromech 360 colloquium, 13-15 May in Saint-Etienne, France.

Kluwer Academic Publishers, the Netherlands, 1997.

[29] ASTM E118-89. Practice for acoustic emission examination of reinforced ther-

mosetting resin pipes. American Society for the Testing of Materials.

[30] S.W. Tsai. Theory of composites design. Think Composites, Dayton, Ohio,

USA, 1992.

21
List of Figures

1 Pin-joint testing configuration . . . . . . . . . . . . . . . . . . . . . . . 24

2 Experimental setup for pin-joint testing . . . . . . . . . . . . . . . . . 25

3 Woven glass-epoxy pin-joint response as a function of pin clearance . . 26

4 In-plane shear response of the glass fabric epoxy composite . . . . . . 27

5 ABAQUS mesh for half the model . . . . . . . . . . . . . . . . . . . . 28

6 Detailed schematic of the finite element mesh around the hole for the

two clearance values . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

7 Force-displacement curves obtained from finite element analysis and

experiment, for a clearance value of 0.1 mm . . . . . . . . . . . . . . . 30

8 Local force distribution on the hole obtained by finite element on the

0.1 mm clearance contact model compared to the classically assumed

sine distribution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

9 Contact angle as a function of the load for the two clearance values . . 32

10 Force-displacement curves obtained from finite element analysis and

experiment, corrected to account for fixture deformation, for a clear-

ance value of 0.1 mm . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

11 Force against longitudinal strain for two pin-joint tests on the same

specimen, clearance=0.1 mm . . . . . . . . . . . . . . . . . . . . . . . 34

12 Local contact defect causing strain differences between the specimen

faces (in-plane Saint-Venant effect) . . . . . . . . . . . . . . . . . . . . 35

13 Comparison between calculated and computed strains for the 0.1 mm

clearance specimen, with different friction coefficients . . . . . . . . . . 36

14 Comparison between calculated and computed strains for the 2 mm

clearance specimen, with different friction coefficients . . . . . . . . . . 37

15 Comparison of the u2 displacement fields for clearance values of 0.1

and 2 mm, friction coefficient 0.3 . . . . . . . . . . . . . . . . . . . . . 38

16 Comparison of the ǫ2 strain for clearance values of 0.1 mm and 2 mm,

friction coefficient 0.3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

22
17 Typical force displacement curve for a 0.1 mm clearance joint with

corresponding specimen damage . . . . . . . . . . . . . . . . . . . . . . 40

18 Typical force displacement curve together with acoustic emission events

for the 0.1 mm clearance joint . . . . . . . . . . . . . . . . . . . . . . . 41

19 Typical force displacement curve together with acoustic emission events

for the 2 mm clearance joint . . . . . . . . . . . . . . . . . . . . . . . . 42

20 Microscopic view of the in-plane damage on a failed 2 mm clearance

specimen (mid-plane) . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

21 Stress field in material axes for the 0.1 mm clearance specimen loaded

at 25 kN (90% of the failure load) . . . . . . . . . . . . . . . . . . . . 44

22 Stress field in material axes for the 2 mm clearance specimen loaded

at 17 kN (90% of the failure load) . . . . . . . . . . . . . . . . . . . . 45

23 In-plane shear stress map in material axes for the 0.1 mm clearance

specimen loaded at 25 kN (90% of the failure load) . . . . . . . . . . . 46

24 In-plane shear stress map in material axes for the 2 mm clearance

specimen loaded at 17 kN (90% of the failure load) . . . . . . . . . . . 47

25 Force displacement response of a 0.1 mm clearance specimen equipped

with back-to-back strain gauges . . . . . . . . . . . . . . . . . . . . . . 48

23
w = 60 mm

AAAA
AAAA
e = 60 mm

AAAA
AA
clearance

AAAA
AA
= 16 mm

l = 220 mm 45

Fibres

AAAAAAAA
AAAAAAAA
AAAAAAAA
F

Figure 1: Pin-joint testing configuration

24
support

washer

bolt
nut
composite specimen

grips

Figure 2: Experimental setup for pin-joint testing

25
30

25
0.5 mm
Force (kN)

20 1 mm 0.1 mm
1.5 mm
15 2 mm

10

0
0 0.5 1 1.5 2 2.5 3 3.5
Displacement (mm)
Figure 3: Woven glass-epoxy pin-joint response as a function of pin clearance

26
70

60

50
Shear stress (MPa)

40

30

20

10

0
0 1 2 3 4 5 6 7
Engineering shear strain (%)

Figure 4: In-plane shear response of the glass fabric epoxy composite

27
Figure 5: ABAQUS mesh for half the model

28
0.1 mm clearance 2 mm clearance

Figure 6: Detailed schematic of the finite element mesh around the hole for the two clear-
ance values

29
contact + material nonlinearity
30 linear elastic
contact

25

20 experimental
Force (kN)

15

10

5
Clearance: 0.1 mm
0
0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
Displacement (mm)

Figure 7: Force-displacement curves obtained from finite element analysis and experiment,
for a clearance value of 0.1 mm

30
350
sine distribution
300

250 finite element


Force (N)

200

150

100

50

0
0 10 20 30 40 50 60 70 80 90
Angle ( )
Figure 8: Local force distribution on the hole obtained by finite element on the 0.1 mm
clearance contact model compared to the classically assumed sine distribution

31
40
0.1 mm
35 2 mm

30
Force (kN)

25

20

15

10

0
0 20 40 60 80 100 120 140 160 180
Angle ( )
Figure 9: Contact angle as a function of the load for the two clearance values

32
30 finite element model

25 experimental
corrected
experimental
20
Force (kN)

15

10

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
Displacement (mm)
Figure 10: Force-displacement curves obtained from finite element analysis and experiment,
corrected to account for fixture deformation, for a clearance value of 0.1 mm

33
Figure 11: Force against longitudinal strain for two pin-joint tests on the same specimen,
clearance=0.1 mm

34
AAAAA
AAAAA
composite specimen
AAAAA
AAAAA
strain gauge
strain gauge
(lower strain)
AAAAA
AAAAA
AAAAA
(higher strain)

AAAAA
AAAAA
AAAAA
threaded rod
AAAAA
local contact
defect
(magnified)

F
Figure 12: Local contact defect causing strain differences between the specimen faces (in-
plane Saint-Venant effect)

35
Longitudinal strain (%)
Figure 13: Comparison between calculated and computed strains for the 0.1 mm clearance
specimen, with different friction coefficients

36
Longitudinal strain (%)
Figure 14: Comparison between calculated and computed strains for the 2 mm clearance
specimen, with different friction coefficients

37
Displacement in micrometer

Finite element calculation Experimental

Clearance: 0.1 mm - Force: 7900 N

Displacement in micrometer

Finite element calculation Experimental

Clearance: 2 mm - Force: 8100 N

Figure 15: Comparison of the u2 displacement fields for clearance values of 0.1 and 2 mm,
friction coefficient 0.3

38
Strain in microstrain

Finite element calculation Experimental


Clearance: 0.1 mm - Force: 7900 N

Strain in microstrain

Finite element calculation Experimental


Clearance: 2 mm - Force: 8100 N

Figure 16: Comparison of the ǫ2 strain for clearance values of 0.1 mm and 2 mm, friction
coefficient 0.3

39
40

35

30

25
Force (kN)

20

15

10

0
0 1 2 3 4 5 6 7 8 9 10
Displacement (mm)

Figure 17: Typical force displacement curve for a 0.1 mm clearance joint with corresponding
specimen damage

40
30 1000
Clearance: 0.1 mm
25

Cumulated number of events


800
Force (kN)

20
600
15
400
10

200
5

0 0
0 1 2 3
Displacement (mm)
Figure 18: Typical force displacement curve together with acoustic emission events for the
0.1 mm clearance joint

41
20 1000
Clearance: 2 mm

Cumulated number of events


16 800
Force (kN)

12 600

8 400

4 200

0 0
0 0.5 1 1.5
Displacement (mm)

Figure 19: Typical force displacement curve together with acoustic emission events for the
2 mm clearance joint

42
Figure 20: Microscopic view of the in-plane damage on a failed 2 mm clearance specimen
(mid-plane)

43
Stress in Pa Stress in Pa

σx stress σy stress

Stress in Pa

σs stress

Figure 21: Stress field in material axes for the 0.1 mm clearance specimen loaded at 25 kN
(90% of the failure load)

44
Stress in Pa Stress in Pa

σx stress σy stress

Stress in Pa

σs stress

Figure 22: Stress field in material axes for the 2 mm clearance specimen loaded at 17 kN
(90% of the failure load)

45
Stress in Pa

σs stress

Figure 23: In-plane shear stress map in material axes for the 0.1 mm clearance specimen
loaded at 25 kN (90% of the failure load)

46
Stress in Pa

σs stress

Figure 24: In-plane shear stress map in material axes for the 2 mm clearance specimen
loaded at 17 kN (90% of the failure load)

47
average face 1
25

face 2 20

Force (kN)
15

10

5
clearance: 0.1 mm
0
-3 -2.5 -2 -1.5 -1 -0.5 0
Longitudinal strain (%)

Figure 25: Force displacement response of a 0.1 mm clearance specimen equipped with
back-to-back strain gauges

48
List of Tables

1 Mechanical properties of the glass fabric epoxy composite . . . . . . . 50

2 Results of the joint tests . . . . . . . . . . . . . . . . . . . . . . . . . . 51

3 Damage investigation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

4 Predicted and measured failure forces for the two clearance values . . 53

49
Table 1: Mechanical properties of the glass fabric epoxy composite

Type Tensile Tensile failure Poisson’s


modulus stress (MPa) ratio
(GPa)
warp 25.4 540 0.13
weft 25.4 530 0.13

In-plane shear modulus In-plane shear failure


(GPa) stress (MPa)
5.2 91

Type Compressive failure stress (MPa)


warp 570 (Hexcel - PrEN2850)
weft 506 (Hexcel - PrEN2850)

50
Table 2: Results of the joint tests

Clearance(mm) / Specimen Failure load (kN)


number
0.1 / spec. 1 29.2
0.1 / spec. 2 28.2
0.1 / spec. 3 29.2
0.1 / spec. 4 29.4

Mean 29.0
2 / spec. 1 18.0
2 / spec. 2 19.3
2 / spec. 3 19.7
2 / spec. 4 18.3

Mean 18.8

51
Table 3: Damage investigation

Clearance(mm) Maximum Visible


/ Specimen load in kN damage
number (% of failure
load)
0.1 / spec. 1 15.9 (55%) NO
0.1 / spec. 2 20.1 (69%) NO
0.1 / spec. 3 25.0 (86%) YES
2 / spec. 1 10.0 (53%) NO
2 / spec. 2 13.5 (72%) NO
2 / spec. 3 17.0 (90%) YES

52
Table 4: Predicted and measured failure forces for the two clearance values
0.1 mm clearance 2 mm clearance
Experimental fail- 25 17
ure load (kN)
Predicted failure 26 19
load (kN)
Relative difference 4 12
(%)

53

View publication stats

You might also like