Sie sind auf Seite 1von 9

International Journal of Adhesion & Adhesives 59 (2015) 53–61

Contents lists available at ScienceDirect

International Journal of Adhesion & Adhesives


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

Progressive damage modeling of adhesively bonded lap joints


Iordanis T. Masmanidis, Theodore P. Philippidis n
Department of Mechanical Engineering & Aeronautics, University of Patras, P.O. Box 1401, GR 26504 Panepistimioupolis Rion, Greece

art ic l e i nf o a b s t r a c t

Article history: A continuum damage model for simulating damage propagation of bonded joints is presented, introducing a
Accepted 1 February 2015 linear softening damage process for the adhesive agent. Material models simulating anisotropic non-linear
Available online 8 February 2015 elastic behavior and distributed damage accumulation were used for the composite adherends as well. The
Keywords: proposed modeling procedure was applied to a series of lap joints accounting for adhesion either by means
B. Composites of secondary bonding or co-bonding. Stress analysis was performed using plane strain elements of a
D. Cohesive zone model commercial finite element code allowing implementation of user defined constitutive equations. Numerical
E. Joggle lap joint results for the different overlap lengths under investigation were in good agreement with experimental data
Progressive damage in terms of joint strength and overall structural behavior.
& 2015 Elsevier Ltd. All rights reserved.

1. Introduction fracture characterization experiments must be performed to specify


the cohesive law parameters.
The increasing size of structures built exclusively from light- In the present work a modeling procedure for simulating adhe-
weight composite materials, such as wind turbine rotor blades, has sive joint behavior is presented. Degradation models, simulating
raised the need for more advanced design tools to optimize the damage propagation, without predefining the failure path, are intro-
joints between the various components. Moreover, the fact that duced for both composite adherends and polymeric adhesives.
maintenance and repair of such structures is becoming a major Material nonlinearity of the adherends is also implemented, while
issue, due to the high replacement cost, increases the demand of the presented softening procedure accounts for energy dissipation
more efficient joint design techniques. during debonding. A series of secondary bonded and co-bonded lap
Numerous studies on the analysis of bonded joints with compo- joints, of varying overlap length, were analyzed by means of the
site adherends, using the Finite Element method, have been pub- Finite Elements method to verify the predictive capabilities of the
lished, see the recent reviews [1,2]. These studies can be categorized proposed model. The mesh refinement issue is addressed by corre-
by their approach for predicting the strength of the adhesive joints. lating element size with the softening law parameters so that the FE
The continuum mechanics approach, that assumes perfect bonding results are independent from the mesh density. Validation of the FE
between the adhesive and the adherends, suffers from the bi- modeling procedure was performed by comparing predictions with
material singularities inherent in a bonded joint and as a result experimental results and numerical predictions using the CZM
maximum stress and strain for such a model will vary greatly with approach.
mesh refinement. The fracture mechanics approach addresses the
singularity issue but still there are limitations such as the difficulty
of the finite element modeling procedure to calculate the stress state 2. Material nonlinearity and progressive damage model
at the crack tip and the need for measuring the fracture properties of
the materials. Finally, there is the Cohesive Zone Modeling (CZM) 2.1. Composite nonlinear behavior
approach which simulates the macroscopic damage along a pre-
defined crack path by specification of a traction-separation response Mechanical properties of both, the glass fiber composite adher-
between initially coincident nodes on either side of the path. The ends and adhesive paste, used for coupon manufacturing, were
great advantage of the cohesive models is their ability to simulate determined experimentally in a comprehensive material character-
onset and non-self-similar growth of adhesive damage. However, ization campaign. The epoxy resin system used was HUNTSMAN
except from the downside of predefining the damage path in CZMs, Araldites LY 3505 / Hardener Aradurs 3405 and laminate unidirec-
tional reinforcement was AHLSTROM E-glass fibers of an areal
weight equal to 700 g/m2. Slight material non-linearity was found
n
Corresponding author. Tel.: þ 30 2610 969450, 997235; fax: þ30 2610 969417. parallel to the fibers, whereas a more pronounced one was mea-
E-mail address: philippidis@mech.upatras.gr (T.P. Philippidis). sured transverse to the fibers, different in tension and compression.

http://dx.doi.org/10.1016/j.ijadhadh.2015.02.001
0143-7496/& 2015 Elsevier Ltd. All rights reserved.
54 I.T. Masmanidis, T.P. Philippidis / International Journal of Adhesion & Adhesives 59 (2015) 53–61

As expected, a highly non-linear behavior under in-plane shear was the experimental stress–strain curve deviates from linearity close to
observed as well. coupon failure. Properties of the two resins are listed in Table 3.
To account for material non-linearity, incremental stress–strain
relations were implemented, retaining the validity of the general- 2.3. Progressive damage model
ized Hooke law for each individual interval as described in [3,4]:
E1t ν12 E2t 2.3.1. Composite adherends
dσ 1 ¼ dε1 þ dε2 Besides non-linear mechanical response, progressive damage
1  ðE2t =E1t Þν212 1  ðE2t =E1t Þν212
mechanics were also implemented in the FE modeling procedure.
ν12 E2t E2t
dσ 2 ¼ dε1 þ dε2 To account for the composite adherends progressive failure, the
1  ðE2t =E1t Þν212 1  ðE2t =E1t Þν212 Puck criterion [6] with the associated property degradation strategy
dσ 6 ¼ G12t dε6 ð1Þ is used. Details of the failure mode dependent stiffness degrada-
tion were described in [3] and are summarized for completeness
The tangential elastic moduli in the principal coordinate system of the
in Table 4. According to Puck theory, there are 5 ply damage modes,
orthotropic material, E1t (parallel to the fiber), E2t (transversely), G12t
two associated with either tensile or compressive fiber failure (FF)
(in-plane shear) were derived as follows by adopting the nonlinear
and three describing matrix cracking or inter-fiber failure (IFF);
constitutive model introduced by Richard and Blacklock [5]:
IFFA, -B, -C resulting mainly from a combination of transverse to the
Eoi εi fiber normal stress and in-plane shear.
σi ¼ h  ni ið1=ni Þ ; i ¼ 1; 2; 6 ð2Þ
1 þ Eoi εi =σ oi Index (k) in the above relations refers to an arbitrary load step
after failure has been detected. The degradation factor, η r 1, multi-
By differentiating Eq. (2) one has: plying the engineering elastic constants to account for damage
  ni ðð1=ni Þ þ 1Þ growth in the ply is given by [6]:
dσ σi
Eit ¼ i ¼ Eoi 1  i ¼ 1; 2; 1  ηr
dεi σ oi ηðk  1Þ ¼ þ ηr ð4Þ
1 þ cðf EðIFFÞ  1Þξ
ðk  1Þ
  n6 ðð1=n6 Þ þ 1Þ
dσ 6 σ6
G12t ¼ ¼ Go12 1  ð3Þ
dε6 σ o6 where fE(IFF) is the failure effort as calculated by Puck's matrix failure
criterion while c¼5, ξ ¼3 and ηr ¼ 1  10  6 are the values of the
A summary of the numerical values for all constants in elasticity parameters of Eq. (4).
expressions can be found in Table 1; they were derived through
non-linear regression on the experimental data. Mean values for 2.3.2. Polymer resin
tensile and compressive strength properties in the fiber direction To account for the adhesive paste progressive damage (micro-
(XT, Xc), transversely to the fibers (YT, Yc) and in shear (S) for the cracking), since a brittle isotropic adhesive material is assumed, the
composite tested are given in Table 2. The relatively low elastic paraboloid failure surface criterion by Stassi D' Alia [7], adapted for
properties of the adherend are due to the wet hand-layup manu- generalized plane strain, is implemented:
facturing technique, characteristic of the in-situ patching procedure      
of the industrial partner; typically this results in a fiber weight σ x  σ y 2 þ σ y  σ z 2 þ ðσ z  σ x Þ2 þ 6τ2xy þ 2σ u ðR  1Þ σ x þ σ y þ σ z  2Rσ 2u r 0
fraction of ca. 51%. Mean values were deduced from 5 tests for each ð5Þ
specimen type while engineering elastic constants were derived as
where σ u represents the adhesive tensile strength and R, expressing
suggested by relevant standards.
the strength differential effect, is the ratio of compressive to tensile
failure stress. Here R is calculated in terms of the measured values
2.2. Polymer matrix and adhesive resin properties
Table 3
The epoxy resin, used as adhesive for the secondary bonded Engineering elastic constants and failure stresses for the polymer systems.
specimens, is HUNTSMAN XD 4734 with XD 4741-S hardener cured
E [GPa] G [GPa] σu [MPa] τu [MPa]
at 80 1C for 1 h. The response of both the, previously described,
polymer matrix of the adherends and adhesive resin was found to be Araldite LY3505/Aradur 3405 3.98 1.48 56.94 51.64
slightly non-linear especially under shear stressing. In this work the XD 4734/XD 4741-S 4.01 1.39 35.29 39.94
epoxy resins are assumed to have linear behavior until failure since

Table 1
Elasticity constants for the non-linear model, Eq. (3), of UD Glass/Epoxy composite. Table 4
Progressive stiffness degradation model for the
ν12 ¼ 0:26 composite adherends.

Eoi ½MPa σ oi ½MPa ni Failure mode

E1t 26,870.00 9,016.00 1.00 FF(T) or FF(C) EðkÞ  10


1 ¼ 10  E1
EðTÞ 9,478.00 49.00 3.04
EðkÞ  10
2t 2 ¼ 10  E2
EðCÞ 10,473.00 178.00 2.54
GðkÞ  10
12 ¼ 10  G12
2t
G12t 2,760.00 44.00 1.87

IFF(A) EðkÞ
2 ¼η
ðk  1Þ
 E2
GðkÞ
12 ¼ η
ðk  1Þ
 G12
Table 2
Failure stresses for the Composite material (in MPa). IFF(B) GðkÞ
12 ¼ η
ðk  1Þ
 G12
IFF(C) EðkÞ
2 ¼ 10
 10
 E2
XT XC YT YC S
GðkÞ
12 ¼ 10  10
 G12
558.60 411.12 40.00 128.14 38.42
I.T. Masmanidis, T.P. Philippidis / International Journal of Adhesion & Adhesives 59 (2015) 53–61 55

of shear strength τu and σ u of the adhesive. As it can be readily A continuum damage approach has been also introduced in [8],
derived from Eq. (5): where a stress/strain softening relationship is used as well. There
 2 the area under the stress/strain curve was related to the critical
τu energy release rate by introducing a characteristic length in order
R¼3 ð6Þ
σu to transform displacements to strains.
When failure is detected a softening process is imposed to the
adhesive to account for damage evolution. For the kth step, after
failure has been detected, the degraded moduli are calculated by: 3. Coupon description and FE implementation

ðk  1Þ
EðkÞ ¼ 1  d E The lap joint geometry presented here, see Fig. 2, is usually
referred to as joggle lap joint (JLJ) and is often encountered in joints
GðkÞ ¼ EðkÞ =2ð1 þ νÞ ð7Þ where a smooth final surface is required, such as wind turbine rotor
blade restoration procedures. The presence of the joggle increases
The damage index, d, is calculated by considering the linear
the complexity of the problem due to the curvature; a study in
softening process shown in Fig. 1 for the equivalent stress and
which the JLJ configuration was also considered can be found in [9].
strain:
Total length and width of the coupons are equal to l¼500 mm and
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1    w¼25 mm while the curvature length equals to a¼15 mm. In both
σ eq ¼ σ x  σ y 2 þ σ y  σ z 2 þ ðσ z  σ x Þ2 þ 6τ2xy ð8Þ secondary bonded and co-bonded joint types the overlap length
2
(2c) varied from 50 to 200 mm in steps of 50mm, see Fig. 3.
An alternative approach concerning coupon geometry could be
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

1    3 to keep constant the ratio of overlap to total specimen length;
εeq ¼ pffiffiffi εx  εy 2 þ εy  εz 2 þ ðεz  εx Þ2 þ γ 2xy ð9Þ another, to keep constant the distance of the joint edges from the
2ð1 þ νÞ 2
clamp area, both resulting in specimens of varying overall length.
As it can be readily proved from the above equations, σ eq ¼ Eεeq Nevertheless, a preliminary numerical investigation has revealed
and thus the scalar parameter d of Eq. (7) expressing the damage no significant differences in the ultimate load capacity and thus
accumulation for the current load step (k) is given by: results concerning joint behavior from the comparison between
 the various overlap lengths presented in Section 4 are believed
ðkÞ
εequ εðkÞ
eq  εeqo unaffected of said geometry variations.
d ¼ ðkÞ   ð10Þ
εeq εequ  εeqo All JLJ coupon adherends consist of 2 layers of the Glass/Epoxy
UD previously mentioned, fibers directed along the coupon long
where εðeqkÞ is the equivalent strain at that load step, εeqo ¼ εðeqnÞ is the axis, with a nominal total thickness equal to tc ¼1.744 mm.
equivalent strain of the material at the (nth) load step when the
failure criterion was satisfied and the softening process began and
3.1. Secondary bonded coupons
finally εequ is its maximum value.
Therefore, by updating the stiffness matrix, the stress for this
In the coupons manufactured with secondary bonding the
step is calculated by:
adhesive thickness was measured with the use of a digital caliper
σ ðkÞ ðkÞ ðkÞ
i ¼ C ij εj with i; j ¼ 1; …; 6 ð11Þ and by analyzing coupon free edge pictures with Mathworks
Matlab Image Processing toolbox, resulting in an average thickness
The slope of line AB in Fig. 1 defines the new stiffness of the of ta ¼1.30 mm for all coupons. For each overlap length, the curved
degraded material for a load step during the softening process. The adherend plate was manufactured by the industrial partner using
value σ eqo is of the equivalent stress when the failure criterion is a mold and was bonded to the already cured flat adherend plate
satisfied (nth load step) and therefore, it can have different values using spacers to ensure even distribution of adhesive thickness.
at the various elements depending on the current stress combina-
tion when failure is detected.
On the other hand, the value of the maximum equivalent
strain,εequ , for which complete failure occurs is assumed dependent
only on joint geometry and the adhesive material; it is derived by
adapting numerical simulations on the observed experimental
behavior of an arbitrary overlap length specimen from each joint
type. See Section 3.3 for details.

Fig. 2. Geometry of the JLJ repair coupons.

Fig. 1. Softening stress-strain curve after failure onset of the adhesive. Fig. 3. Adhesively bonded JLJ coupons of varying overlap length.
56 I.T. Masmanidis, T.P. Philippidis / International Journal of Adhesion & Adhesives 59 (2015) 53–61

Specimens were then cut using water jet. The adhesive paste spew, determined by:
present in all coupons, was included in the FE models, see Fig. 4,
since it was found to drastically influence the predicted joint t c Af tr
tr ¼  and t ec ¼ t c  ð12Þ
behavior. 2 ρf 2

where tr and tf are calculated as the thicknesses of the polymer


and fiber layer respectively of a single UD ply as shown in Fig. 6. In
3.2. Co-bonded coupon modeling the above relations Af is the areal mass of the UD glass fabric, equal
to 0.7 kg/m2 and ρf the density of the E-glass fibers, 2560 kg/m3.
While the previously described progressive damage modeling can The fiber elastic properties were back calculated using the ply and
be implemented straightforwardly for secondary bonded joints, the resin effective properties Eoi of Tables 1 and 3 and the micromecha-
same does not apply for the co-bonded ones since there is no distinct nics equations (38a-d) of VDI 2014 Part 3 [11]. Then, by means of the
adhesive layer present. Nevertheless, since adhesion is mainly same equations the elastic constants for the higher fiber volume
achieved by the interaction between the matrix resin located at the fraction equivalent composite layer, with tec ¼1.445 mm, were
joined faces of the composite adherends, a model is introduced that derived and are shown in Table 5. The nonlinear trend, defined by
replaces the ply of the composite adherend by a two-layer effective the other parameters of Table 1, and failure stresses of the equivalent
material, consisting of a modified composite and a distinct polymer composite layer were assumed to be equal to the ones of the original
layer, see Fig. 5. Motivation was provided by an early concept by composite. Finally, the resin layer of thickness tr /2¼0.299 mm has
Puppo and Evensen [10]. the properties of the polymer matrix in the UD composite ply,
More specifically, the total thickness, tc, of the composite Araldite LY 3505, already described in Table 3.
adherend is assumed to be equal to the sum of tec, thickness of
an equivalent composite layer and tr /2, that of the polymer resin

3.3. FE mesh optimization

Non-linear material behavior and progressive damage models


were implemented in ANSYS commercial FE code using the user
programmable features of the PLANE182 element. In order to
reduce computational effort, generalized plane strain analysis was
chosen since relatively small strains are expected in the width
direction of the model. Geometric nonlinearity was also included
in the analysis by taking into account large strain effects.
Traditional finite elements have difficulties in resolving the
Fig. 4. Close-up view of the 50 mm coupon model showing mesh density. stress state at bi-material wedges due to the existence of singula-
rities and their results vary with mesh refinement. The current
work addresses this issue by correlating mesh density with the
maximum equivalent strain, εequ of Eq. (10). Square elements are
used around the overlap region, as it can be seen in Fig. 4, and by
changing the mesh density the value of εequ , for which satisfactory
agreement for joint strength was reached between numerical
predictions and test results, was calculated. The study was
performed with element size ranging from ℓ ¼ 0:1 mm to
ℓ ¼ 0:4 mm in steps of 0.05 mm. The relationship between εequ
and ℓ was derived by non-linear regression; for the epoxy resin of
the co-bonded joints it was found thatεequ ¼ 1:28  10  5 ℓ  1:174
while for the adhesive paste of the secondary bonded joints
εequ ¼ 4:65  10  5 ℓ  1:031 ,ℓ in [m]. Calculations were performed
for the coupon geometry with the largest overlap, i.e. 200 mm, for
each joint type and have proven to be valid for all overlap lengths
resolving the mesh convergence issue. For the results presented
Fig. 5. Co-bonded JLJ coupon model. later on this work an element size of ℓ ¼ 0:3 mm was used.

t r /2

tf
tc t ec
t r /2

Fig. 6. Composite adherend modeled as (a) two homogeneous UD plies (b) a set of resin and fiber layers and (c) as an equivalent composite layer and a resin layer at the
interface.
I.T. Masmanidis, T.P. Philippidis / International Journal of Adhesion & Adhesives 59 (2015) 53–61 57

Table 5 The progressive damage models described earlier were used to


Elasticity constants for the adherend equivalent composite. predict ultimate loads at failure for all the series of coupons. The
term failure at the numerical procedure denotes the complete
Eo1 [MPa] EðTÞ
o2 [MPa] EðCÞ Go12 [MPa] ν12
o2 [MPa]
separation of the adherends. Even though localized failure at areas
31,606.00 11,981.00 13,932.00 3,097.00 0.24
with high stress concentration and along the bondlines was
observed for the adherend elements, the stress developed until
coupon failure was low for extensive composite damage. This
agrees with the experimental results where the extent of damage
at the composite increased with overlap length but only reached a
failure mode of moderate fiber tear out. More specifically the failure
mode for the 50 mm overlap is pure adhesive failure; for the
100 mm overlap, while mostly adhesive, some spots of cohesive
Fig. 7. Loading and boundary conditions. failure are also observed. The dominant failure modes observed by
further increasing overlap length to 150 and 200 mm, were light
3.4. Boundary conditions and moderate fiber tear-out respectively, as it can be seen in Fig. 8.
Even though average stress at the adherends is much lower than
The FE models boundary conditions were selected to properly the respective failure stress, localized damage in the composite plies
simulate the testing procedure. For all coupon models stepwise is observed at the overlap regions due to high stress concentrations
loading is introduced by applying axial displacement increments on induced by the joint geometry and dissimilar material interfaces.
the nodes at the right end of the coupons. To avoid the dependence While increasing overlap length, a plateau in joint strength was
of load step convergence to mesh density, every time an element reached since the majority of the coupons with 200 mm overlap
failure is detected the load is not increased for the following solution failed at almost the same loads with the 150 mm ones, see Table 6.
steps, until element failure stops propagating to adjacent elements Similar observations for single lap joints with a brittle adhesive were
and then the next increment is applied. After a thorough load step reported in [15]; the increase of strength restoration was limited to a
convergence study, including a range of displacement increments of certain overlap length, accompanied with more extensive adherend
0.06 mm to 0.005 mm, it was found out that for a step equal to damage as well.
0.01 mm, all models have fully converged. The above loading strategy A comparison between the numerically predicted load-dis-
resulted in the calculation of an average of more than 1000 steps placement curves and the experimental data is presented
until joint final failure. All other nodal displacements were con- in Fig. 9; the FE results compare well with the test data even though
strained for a distance of d¼50 mm at either end of the coupon the strength of the coupons with smaller overlap is overestimated.
model to account for the gripping of the test machine, Fig. 7. The trend for the coupons to reach their maximum strength for an
overlap length of 150 mm, observed in the tests, is fairly corrobo-
3.5. Cohesive zone models rated by the FE results. This behavior is driven by the geometry and
material non-linearity combined with stress concentration at the
In order to further validate the modeling procedure introduced edges of the overlap during debonding progression, leading to the
in the previous section of this manuscript, the co-bonded coupon variation in failure mode as observed in Fig. 8. A systematic and
behavior has been also simulated by modeling the progressive
debonding using ANSYS contact elements and its integrated Table 6
bilinear Cohesive Zone Material model [12]. JLJ coupons failure loads.
Values for the critical strain energy release rate for the mixed mode
Overlap length Secondary bonding Co-bonding Reference
CZM model are taken equal to GΙ c ¼ 1160 N=m and GΙΙ c ¼ 2030 N=m
(mm)
[13,14], while the relative displacement for damage initiation is equal FEA Test COV FEA Test COV Test COV
to δο ¼ 1  10  6 m. All FE runs using the CZM model have fully (kN) (kN) (%) (kN) (kN) (%) (kN) (%)
converged regarding mesh density and number of load increments.
– 24.88 3.10
50 10.27 6.64 22.10 15.60 11.61 9.70
100 13.62 10.51 8.80 16.36 16.31 9.23
4. Results and discussion 150 14.10 14.83 4.67 16.74 16.82 7.63
200 14.11 14.65 4.20 16.60 16.33 3.28
4.1. Test procedure

Seven samples from each joint configuration were tested on a


100 kN MAYES DH 100 S test rig equipped with a 407 MTS
50 mm overlap
controller with its crosshead speed set to 5.8 mm/min. All coupons
were inserted in the grips for 50 mm at each side leaving a gauge
length of 400 mm. Load and displacement were measured by
acquiring the test machine load cell and LVDT signals respectively. 100 mm
Strain monitoring on several key positions was also performed in a
number of JLJ specimens in order to validate the predictions of the
numerical models. 150 mm

4.2. JLJ coupons joined by secondary bonding

The average failure loads and coefficient of variation (COV) from


7 coupons for each JLJ configuration are summarized in Table 6
along with the ultimate load from 7 continuous 2 layer UD coupons,
of same geometry with the JLJ specimens, as a reference. Fig. 8. Typical fracture surfaces of the JLJ secondary bonded coupons.
58 I.T. Masmanidis, T.P. Philippidis / International Journal of Adhesion & Adhesives 59 (2015) 53–61

Fig. 9. Comparison of numerical predictions and test results for the adhesively bonded JLJs.

Fig. 10. Location of strain gauges overlaid on the 2c¼ 150 mm overlap FE model.

Fig. 11. Measured vs. FEA calculated strains at the edge of lower adherend. Fig. 12. Measured vs. FEM calculated strains at the middle of the overlap.

thorough study of this mechanism should be undertaken in the slope change in the FEA curve indicates that debonding onset is
future. predicted at somewhat lower loads by the numerical model; to
The sudden load drop observed in the experimental curves at a further verify and validate its predictions, strain measurements were
magnitude around 4 kN marks the onset of debonding. The observed performed in some specimens of 150 mm overlap.
I.T. Masmanidis, T.P. Philippidis / International Journal of Adhesion & Adhesives 59 (2015) 53–61 59

Fig. 13. Comparison of numerically predicted and experimental damage patterns. (For interpretation of the references to color in this figure legend, the reader is referred to
the web version of this article.)

Fig. 14. Progressive debonding of the JLJ coupon at 20, 50, 75, 95 and 100% of the ultimate load.

Strain gauges 5 mm long were placed on the upper adherend in according to Puck criteria; FFT and FFC denote Tensile and
the middle of the overlap length, i.e. c¼75 mm, and on the lower Compressive Fiber Failure respectively while IFFA, IFFB and IFFC
adherend at a distance of s¼5 mm from the edge as shown in Fig. 10. refer to the 3 matrix failure modes as described in Section 2.3.1. It
Their measurements were compared with the strains calculated should be noted that even the trend for the adhesive spew to
from FEA at the corresponding nodes of the JLJ model. The strain remain attached to the upper adherend after complete debonding,
gauge placed at the lower adherend edge measures zero strain when see Fig. 8, was reproduced for all joints.
total debonding of the first 10 mm of the joint has occurred. A more detailed sequence of damage progression in the JLJ
As it is seen in Fig. 11, complete debonding of the overlap edge coupon of 150 mm overlap is also displayed in Fig. 14 showing
area was indeed predicted by the FE model. However the softening consistency with the observed fracture surfaces of the repair
procedure, eventually leading to the complete debonding, begins coupons indicating light fiber tear out, see Fig. 8.
at lower load levels than the experimentally observed abrupt
failure of the area. The strain gauge placed on the upper adherend 4.3. JLJ coupons joined by co-bonding
was strained up to 14 kN, see Fig. 12, indicating that the middle of
the overlap is close to the region to be debonded just before Comparison of experimental results in the form of load–
complete joint failure. This was verified by the FE model strain displacement curves from tests of JLJ coupons produced by co-
calculations and also by damage progression patterns, as those bonding and numerical predictions are presented in Fig. 15, where
shown in Fig. 13 where the debonding pattern is very similar to FEA curves compare well with experimental ones. The predictions
the experimental observation. The red colored elements in the FE of the Cohesive Zone Model (CZM), described in Section 3.5, are
model correspond to failed adhesive material while other colors also plotted along with the ones of the Continuum Damage Model
correspond to the various failure modes of the adherends (CDM), introduced in this work.
60 I.T. Masmanidis, T.P. Philippidis / International Journal of Adhesion & Adhesives 59 (2015) 53–61

Fig. 15. Numerical against test results for JLJs manufactured with co-bonding.

The trend to reach maximum strength restoration at 100 mm at the strength recovery that can be achieved using JLJ joint config-
was closely predicted as well by the FE results of both the proposed uration. In the current study experiments showed no increase in joint
model and the use of cohesive elements. It is of great significance to strength for overlap length greater than 150 mm or even 100 mm for
note that the time needed by the solver for the fully converged CZM the co-bonded coupons and this behavior was closely predicted by the
analysis approach, in some cases was even 7 times greater than the FE analysis.
one required for the proposed procedure.

Acknowledgments
5. Conclusions
Financial support by Compblades Ltd. (www.compblades.com)
A bilinear softening model combined with a failure criterion through contract D357-2011 with the Research Committee of the
suitable for brittle polymers has been introduced to account for University of Patras is gratefully acknowledged. The research work
damage progression and accumulation phenomena associated with was also partially funded by EYDE-ETAK of the Greek Ministry of
lap joints failure. Interaction of all stress components was taken into Development, in the frame of SYNERGASIA 2011 under contract
account using equivalent stress–strain response at the crack tip to ΣYN11_7_1000 (REWIND).
drive the softening procedure.
In addition to the proposed model, implementation of a progres- References
sive damage strategy for the composite adherends along with material
and geometrical nonlinearity resulted in load–displacement curves [1] Banea MD, Da Silva LFM. Adhesively bonded joints in composite materials: an
that showed good agreement with the experimental response. Esti- overview. Proc IMechE Part L: J Mater.: Des. Appl 2009:1–18.
mating for an arbitrary overlap length the only model parameter that [2] Xiaocong He. A review of finite element analysis of adhesively bonded joints.
Int J Adhes Adhes 2011;31:248–64.
is not related to the adhesive effective properties, resulted in satisfac- [3] Antoniou AE, Kensche C, Philippidis TP. Mechanical behavior of glass/epoxy
tory predictions for the other cases as well. Comparison of numerical tubes under combined static loading. Part II: Validation of FEA progressive
predictions from cohesive zone models and the current approach was damage model. Compos Sci Technol 2009;69:2248–55.
[4] Philippidis TP, Antoniou AE. A progressive damage FEA model for glass/epoxy
also favorable. Furthermore, the lack of need for a predefined crack
shell structures. J Compos Mater 2013;47(5):623–37.
path and use of generalized plane strain with a relatively coarse mesh [5] Richard RM, Blacklock JR. Finite element analysis of inelastic structures. AIAA
make the proposed FEA procedure adaptable for potentially modeling 1969;7:432–8.
large scale repair patches of complex geometries. [6] Puck A, Shürmann H. Failure analysis of FPF laminates by means of physically
based phenomenological models. Compos Sci Technol 2002;62:1633–62.
Numerical simulation provided satisfactory results for ultimate [7] Stassi F-D.' Alia. Teoria Della Plasticita e Sue Applicazioni. G. Denaro Ed.
loads in most cases. Test results demonstrate the existence of a limit Palermo 1958.
I.T. Masmanidis, T.P. Philippidis / International Journal of Adhesion & Adhesives 59 (2015) 53–61 61

[8] de Moura MFSF, Chousal JAG. Cohesive and continuum damage models applied to [13] Tsouvalis NG, Anyfantis KN. Determination of the fracture process zone under
fracture characterization of bonded joints. Int J Mech Sci 2006;48:493–503. Mode I fracture in glass fiber composites. J Compos Mater 2011;46(1):27–41.
[9] Taib AA, Boukhili R, Achiou S, Gordon S, Boukehili. H. Bonded joints with [14] Anyfantis KN, Tsouvalis NG. Experimental and numerical investigation
composite adherends. Part II. Finite element analysis of joggle lap joints. Int of Mode II fracture in fibrous reinforced composites. J Reinf Plast Compos
J Adhes Adhes 2006;26:237–48. 2011;30(6):473–87.
[10] Puppo AH, Evensen HA. Interlaminar shear in laminated composites under [15] Neto JABP, Campilho RDSG, da Silva LFM. Parametric study of adhesive joints
generalized plane stress. J Compos Mater 1970;4:204–20. with composites. Int J Adhes Adhes 2012;37:96–101.
[11] VDI 2014 Part 3. Development of FRP components Analysis. Verein Deutscer
Ingenieure e.V., 2006.
[12] Alfano G, Crisfeld MA. Finite element interface models for the delamination
analysis of laminated composites: mechanical and computational issues. Int
J Numer Methods Eng 2001;50:1701–36.

Das könnte Ihnen auch gefallen