Vous êtes sur la page 1sur 22

The Journal of Adhesion, 90:230251, 2014

Copyright # Taylor & Francis Group, LLC


ISSN: 0021-8464 print=1545-5823 online
DOI: 10.1080/00218464.2013.790773

Cohesive Zone Modeling of Propellant and


Insulation Interface Debonding

QING-CHUN ZHOU, YU-TAO JU, ZHEN WEI, BO HAN, and


CHANG-SHENG ZHOU
Department of Aeronautics and Astronautics, Nanjing University of Science and
Technology, Nanjing, China

The reliability of the bonding of propellant to insulation is a key


part of the analysis of rocket motor structural integrity. In this
study, the debonding of the propellant=insulation interface was
investigated by combining experiment and simulation. The impro-
ved exponential cohesive zone model and the bilinear cohesive
zone model were used to predict the fracture properties of the
adhesive interface. Double cantilever sandwich experiments and
uniaxial tensile tests were performed to determine the correspond-
ing model parameters. Furthermore, cohesive parameters were
calibrated by applying an inverse analysis based on Hooke-Jeeves
optimization algorithm. Good agreement was observed between
the numerical simulation of double cantilever sandwich beam tests
and the experimental curves. These results demonstrate that cohes-
ive zone models can simulate the crack initiation and propagation
of propellant=insulator interface in mode I. The bilinear law was
shown to be more suitable for simulating fracture of the propel-
lant=insulation interface in a strict sense than the exponential
law. The numerical load-displacement curve was found to be
sensitive to all cohesive parameters.

KEYWORDS Delamination; Fracture mechanics; Interfaces;


Numerical analysis; Rubbers

Received 27 December 2012; in final form 26 March 2013.


Address correspondence to Yu-Tao Ju, School of Mechanical Engineering, Nanjing
University of Science and Technology, 200 Xiaolingwei Street, Nanjing 210094, China. E-mail:
juyutao@mail.njust.edu.cn

230
Debonding Simulation of the Propellant and Insulation Interface 231

1. INTRODUCTION

Case-bonded solid propellant rocket motors are used throughout the


aerospace and military fields, and the structural integrity of the grain is key
to the effective motor function. Debonding along the propellant=insulation
interface is one of the major ways in which structural integrity can fail.
During production, storage, and transportation, defects such as microcracks
and voids form at the interface due to the concentration of stress and to
material damage. Then the elevated temperature and pressure produced
by the activity of the motor cause those defects to evolve into debonding,
which may affect motor ballistics or even cause the motor to explode. In this
way, the mechanical study of the propellant=insulation interface plays an
important role in structural integrity research.
Propellant and insulation are linked to each other by a liner that acts as
an adhesive. A considerable amount of research has focused on the improve-
ment of the liner, adhesive strength, and plasticizer migration. Kakade et al.
[1] investigated the influence of particle size of ammonium perchlorate on the
interface properties. Navale et al. [2] studied the effect of additives on liner
properties. Grythe et al. [3] found that plasma treatment may change the peel
strength between insulation materials and polyurethane polymers. In addi-
tion, Schloss [4] characterized the general adhesion properties of different
liners and compared the adhesive strength of carboxy-terminated polybuta-
diene (CTPB) propellant to various substrate materials encountered in solid
propellant rocket motors. Recently, Gottlieb and Bar [5] analyzed the plasti-
cizer migration between the propellant and insulation interface. However, to
the best of our knowledge, no studies on crack nucleation and propagation
of propellant=insulation interface have yet been made available.
There are two principal methods of fracture analysis: linear elastic frac-
ture mechanics (LEFM) and cohesive zone model (CZM). The stress intensity
factor in LEFM changes with the geometry of the specimen, so it cannot be
used as an independent parameter for the characterization of the interface.
This factor does not remain consistent through crack nucleation and propa-
gation [6]. CZM can overcome these shortcomings. In addition, LEFM can
only be used to solve fracture problems with a crack-like notch and flaw.
The nonlinear zone ahead of the crack tip is negligible in size. In contrast,
the CZM can be used to predict the behavior of uncracked structures, includ-
ing those with blunt notches, in addition to the behavior of cracked bodies
[7]. Because of this, the CZM has been used in numerous simulations of
various fracture behaviors, and it has performed well.
Ever since CZMs were first proposed by Dugdale [8] and Barenblatt [9],
failure investigations that have successfully used them have mostly been
about rigid quasi-brittle materials, such as concrete [10] and metal [11]. How-
ever, little attention has been paid to the application of this model on pliant,
232 Q.-C. Zhou et al.

elastic materials. It remains to be verified whether CZM is suitable for those


cases, especially those that involve bonding interfaces. CZM defined a consti-
tutive relationship between the traction and opening displacement in the
fracture process zone, and this relationship was named cohesive law. In
order to determine the whole cohesive law using experiments, some direct
methods may be suitable. Proposed methods include the direct tension test
[12], J integral [13], and equilibrium of energetic forces [14]. However, direct
tension requires a uniform damage, which is difficult to achieve, and the
other two methods can only be used on linear-elastic materials. Although
measured cohesive laws have more fundamental meaning in physics, most
studies assume a pre-defined shape. Then the parameters of CZM can be
obtained through experimentation [15] or extracted through inverse analysis
[16]. However, it is usually difficult to find precise values experimentally, and
without proper starting values, inverse analysis involves profound computa-
tional costs. An efficient approach capable of accurately determining the
values of the parameters is needed. The influence of the shape of cohesive
laws should also be considered.
In this article, we set out to investigate the debonding simulation of the
propellant=insulation interface with cohesive zone models. All essential
parameters of CZM were achieved by combining experiments and numeral
studies. First, two corresponding tests were carried out to determine the esti-
mated values of the parameters. Then those values were used as input for the
inverse procedure based on Hooke-Jeeves optimization algorithm. In this
way, optimal model parameters were acquired efficiently and accurately.
Two distinct cohesive laws were used to address the effects of the shape
of cohesive laws on the simulations. The simulation results of mode I fracture
were consistent with the experimental data. The results of this work may
facilitate numerical analysis of the grain structural integrity of a case-bonded
solid propellant rocket motor.
This paper is arranged as follows. In Section 2, we briefly describe two
kinds of CZMs used in this work. In Section 3, we establish the material
properties and present the experiments by which we determined the model
parameters. In Section 4, we provide the numerical implementation of inter-
face debonding and introduce an inverse analysis to calibrate the model
parameters. In Section 5, numerical results are compared to experimental
data. The sensitivity of CZM parameters was used to investigate the influence
of adhesive properties on the fracture behavior. Finally, we provide a
summary and conclusions.

2. COHESIVE ZONE MODEL

Cohesive laws can be classified into the following groups based on the
shapes of their curves: bilinear laws, piecewise laws, polynomial laws, and
Debonding Simulation of the Propellant and Insulation Interface 233

exponential laws. Numerous previous studies [17,18] have suggested that


only model parameters are critical, while the shape of the law is subordinate.
This rule has been recently challenged by other reports [1921] showing the
shape of the cohesive law can significantly affect results of the fracture analy-
sis. Therefore, it becomes essential to select an appropriate cohesive law in a
particular case. In keeping with this, we considered the two most widely
used cohesive laws, bilinear law and exponential law, to complete the frac-
ture simulation of the propellant=insulation interface. The bilinear law and
exponential law were selected because they are each one of the representa-
tive cohesive laws underlying the modeling of brittle fractures and ductile
fractures, respectively. The details of the two CZM are briefly described as
follows.
The bilinear law was originally proposed by Geubelle and Baylor [22]
for simulation of the impact-induced delamination of composites. By
taking the evolution of damage in a mixed mode into account, the normal
and tangential traction-separation of the bilinear law can be expressed
analytically [23]:
( rmax
dn Dn Dn < 0
Tn 1
1  D rdmax
n
Dn Dn  0)

smax
Tt 1  D Dt 2
dt

Here, the subscripts n and t represent normal and tangential behaviors. T


denotes the traction. rmax and smax are the normal and tangential cohesive
strength, respectively. d and D are the characteristic length parameter and
separation, respectively. The scalar variable D is the stiffness degradation,
which represents the overall cumulative damage to the material. It is defined
as follows:
(  
0 dm  dom
D dfm dm dom   3
f o dm > dom
dm dm dm

q
Here, the effective displacement dm is expressed as dm hDn i2 D2t . Here,
hi is the McCauley bracket and dom and dfm are the effective displacements at
the initiation of damage and complete failure, respectively.
The cohesive energy / is equal to the areas under the traction-
separation curve. Thus, the normal and tangential cohesive energy take the
following form:

/n rmax Dcn =2; /t smax Dct =2 4


234 Q.-C. Zhou et al.

Here, Dcn and Dct are the normal and tangential critical separation, respect-
ively. The pure mode I and mode II of the bilinear law are shown graphically
in Figs. 1a and 1b.
Like the bilinear law, the exponential law introduced by Xu and
Needleman [24] is also very frequently used. The exponential law is a
coupled law, which means normal and tangential traction depend not only
on the normal but also the tangential opening displacement. Unfortunately,
it only realistically describes the coupling interaction in certain specific cases.
In order to overcome this limitation, an improved exponential law was
proposed by van den Bosch et al. [25]:

  !
/n Dn Dn D2t
Tn exp  exp  2 5
dn dn dn dt

    !
/t Dt Dn Dn D2t
Tt 2 1 exp  exp  2 : 6
dt dt dn dn dt

The relationships between cohesive energy and cohesive strength in


normal and tangential are as follows:

p
/n rmax dn exp1; /t smax dt exp1=2: 7

FIGURE 1 The traction-separation law of bilinear and exponential cohesive zone model.
Debonding Simulation of the Propellant and Insulation Interface 235

The normal traction-separation law of pure mode I fracture and a


tangential law of pure mode II fracture are depicted in Figs. 1c and 1d,
respectively.

3. EXPERIMENT

In this section, the material properties of the propellant and insulation are
described in Section 3.1. The preparation of the propellant=insulation
adhesive joints and the bulk adhesive specimens are presented in Section
3.2. The determination of cohesive parameters using corresponding experi-
ments is given in Section 3.3.

3.1. Materials
The propellant (Huian1, Xian, China) was an HTPB-based composite vis-
coelastic material highly filled with solid particles. The formulation contained
69.5% oxidizer (ammonium perchlorate), 18.5% metallic fuel (aluminum),
8% binder (HTPB), and other additives. A generalized Maxwell model was
used to characterize its viscoelastic behavior. The viscoelastic material
properties, bulk relaxation modulus, and shear relaxation modulus can be
expressed in terms of the relaxation modulus E(t). E(t) can be expanded in
a Prony series:

X
n  t

Et E0  Ei 1  e si 8
i1

The Prony Series coefficients were experimentally determined using stress


relaxation tests at 20 C [26] and are listed in Table 1. The instantaneous
modulus E0 was 15 MPa, and Poissons ratio n was assumed to be 0.499 [27].
The insulation material (Huian) consisted of ethylene propylene diene
monomer (EPDM) rubber and cross-linked agent, flame-retardant agent,
and fibers. It is well accepted that strain energy potential U was commonly
used to relate stresses to strains in hyperplasic materials. Several different
strain energy potentials are available, including the polynomial model, the

TABLE 1 The Relaxation Modulus Data of HTPB Propellant

i Ei [MPa] si [s]

1 8.37 0.592
2 1.80 1.148
3 1.56 12.081
4 0.734 55.579
5 0.431 217.588
236 Q.-C. Zhou et al.

FIGURE 2 Hyperelastic constitutive model fitting of the insulation.

Ogden model, and the Marlow model. The material coefficients of the differ-
ent hyperelastic models can be calibrated using ABAQUS1(Simulia, Provi-
dence, U.S.) from experimental data. We can also determine the optimal
strain energy potential by comparing the numerical curves to the experi-
mental data.
The uniaxial tension test with a rectangular specimen was suggested.
The tension specimen was 140 mm in length, 4 mm in height, and 10 mm
in width, and the gauge length was 100 mm. The incomplete results of data
fitting are shown in Fig. 2.
From these results, it is evident that the Marlow model and Polynomial
model (N 1) agree more closely with the experimental data than the other
models do. In view of the large deformation of hyperelastic materials in prac-
tice, the polynomial model (N 1, i.e., Mooney-Rivlin model) was recom-
mended. The form of the Mooney-Rivlin strain energy potential is as follows:
1 el
U C10 I 1  3 C01 I 2  3 J  12 : 9
D1

Here, I 1 and I 2 are the first and second deviatoric strain invariants, respect-
ively, and Jel is the elastic volume ratio. The material coefficients C10, C01, and
D1 were calculated and reported in Table 2.

TABLE 2 The Material Coefficients of the Hyperelastic


Models of the Insulation

C10 [MPa] C01 [MPa] D1 [M=Pa]

0.07966 0.06078 0.14338


Debonding Simulation of the Propellant and Insulation Interface 237

3.2. Specimen Preparation


The propellant and insulation joints used in the following tests were pro-
duced as follows. First, coupons of propellant and insulation were cut from
plates. Each coupon was nominally 120 mm long, 10 mm wide, and 5 mm
(propellant) or 4 mm (insulation) thick. Then the adhesive, i.e., the liner,
was formed by blending hydroxyl-terminated polybutadiene (HTPB, Huian)
and isophorone diisocyanate (IPDI, aladdin1, Shanghai, China) proportion-
ately in a vertical planetary mixer in vacuum, in a water bath at 40 C for
30 min. After that, it was cured at 70 C for 2 h in a vacuum and used to bond
the propellant and insulation. Thickness was controlled at 0.2 mm. A Teflon
film was inserted into the interface to produce one 30-mm pre-crack. After
that, the entire component was co-cured in an oven at 70 C for 7 days.
The bulk adhesive specimens for measuring its tension properties were
prepared as follows. When the adhesive was processed and ready for bond-
ing, some of it was cast in a mold painted with a release agent and then
co-cured with the propellant and insulation joints. Then it was peeled off
and cut into sheets 100 mm in length, 10 mm in width, and 5 mm thick.

3.3. Determination of Cohesive Parameters


The cohesive law has the following parameters: cohesive strength, cohesive
energy, and characteristic length. For the exponential law, only two of them
are independent. Once estimated values of cohesive strength and cohesive
energy are given by the experiments, the remaining value can be calculated
using Eq. (7). However, the characteristic length of the bilinear law must also
be determined. Because it is equal to the ratio of cohesive strength to initial
stiffness, most studies have considered initial stiffness to be the third para-
meter. In this section, cohesive energy, cohesive strength, and initial stiffness
(for bilinear law) were determined using the following tests.

3.3.1. COHESIVE ENERGY


The cohesive energy is theoretically equal to the critical strain energy release
rate GIc of materials. As mentioned earlier, the propellant and insulation are
fairly pliant and have a small ratio between self-weight and stiffness. Hence,
most of those tests, such as three-point bend test and single-edge-notched
tensile test, are not suitable for the determination of GIc. The more applicable
double cantilever beam (DCB) test was selected for this project for this rea-
son. In addition, the propellant is a particulate filler material that is sensitive
to impact. If the loading pin is directly fixed on the propellant surface, the
areas of the propellant in contact with the loading pin can become damaged
or break during loading. In order to increase the entire stiffness and protect
the propellant surface, two aluminum plates are added to the DCB specimen.
238 Q.-C. Zhou et al.

We designed the double cantilever sandwich beam (DCSB) specimen shown


in Fig. 3 according to the ASTM D3433 [28].
The specimen is 150 mm in length, 19 mm in height, and 10 mm in
width. The aluminum plates are 5 mm thick and were thoroughly cleaned
with ethanol and dried at 60 C for 2 h, then attached to the prepared propel-
lant and insulation joints using an acrylic modified epoxy adhesive (Kangda1
WD101, Shanghai, China). This adhesive has been tested and shown to be
stronger than the liner. It is tough enough that deformation has a negligible
effect on the accuracy of the experiments. All specimens were solidified at
room temperature for 24 h before testing.
An electrical material testing system (Qingji1 WJ-211, Shanghai, China)
was used to test the specimens under normal environmental conditions (42%
relative humidity at 25 C), with a constant displacement rate of 1 mm=min.
One side of the specimen had a scale attached to it, so that the tip of the
crack could be located and the cohesive zone quantified with the help of
an optical microscope coupled with a CCD camera (CQIC1 ZSA302,
Chongqing, China). The load vs. load-point displacement curve was
recorded by a computerized data acquisition system.
Figure 4 presents the typical load-displacement curve characterizing the
macroscopic response of interfacial fracture obtained from experiments. The
P-d curve consists of the linear ascending component, softening ascending
component, and descending component. In stage I, the load increased line-
arly along with displacement. This can be explained by assuming that the
specimen acts as a pure bending aluminum beam because it is stiffer than
the propellant and insulation. Along with the loading, the propellant and
insulation began to show their nonlinear properties. Despite this, the cumu-
lative damage around the pre-crack tip caused the slope of the curve to
decline gradually so that it almost seems to gently incline. The experimental
observation showed that the initial debonding of the interface takes place
during this period. It was very difficult to capture and record the points accu-
rately. During stage III, the load reached the peak value and decreased with
increasing displacement. The interface crack experienced a rapid propa-
gation followed by a stable propagation.

FIGURE 3 Double cantilever sandwich beam specimen.


Debonding Simulation of the Propellant and Insulation Interface 239

FIGURE 4 Typical experimental load-displacement curve.

Images of the interface were taken using an optical microscope coupled


with a CCD camera, as depicted in Fig. 5. The propellant=insulation interface
turned out to be more complicated than the fracture of the homogeneous
materials or other bonded joints, which have unquestionable cracks. One sig-
nificant elastic deformation of the adhesive occurred in the crack tip of the
interface, producing continuous triangular areas, as shown in Fig. 5a. With
increasing loading, the edge of the triangular area evolved into discrete fila-
ments, as indicated in Fig. 5b. Once the adhesive was stretched to several
dozen times its original thickness, the filaments failed, which was consistent
with at least one previous report [29]. The CZM was first applied to brittle
materials and then extended to quasi-brittle materials with non-negligible
plastic deformation zones ahead of the crack tip. Though the failure of pro-
pellant=insulation interface is out of that range, we may also assume that the
triangular area is the same as the fracture process zone. In other words, it sat-
isfied the cohesive law. The numerical verification described below provided
more support for this hypothesis.

FIGURE 5 Optical micrographs of bonding interface. (color figure available online).


240 Q.-C. Zhou et al.

For linear-elastic material behavior, the strain energy release rate GI can
be obtained using IwrinKies equation [30]:

1 2 @C
GI P : 10
2B @a

Here, B is the width of the specimen and P is the applied load. C and a are
the compliance and crack length, respectively.
The relationship between compliance and crack length can be obtained
using corrected beam theory [31]. Upon insertion into Eq. (10), the critical
strain energy release rate GIc of the DCSB specimen is given as follows [32]:

3Pc dc F
GIc  : 11
2Ba  D N

Here, Pc and dc are the critical load and the corresponding displacement,
respectively. D, F, and N are the crack length correction, large-displacement
correction, and load block correction, respectively. For details of calculations
of those corrections, refer to IS0 15024:2001(E) [32]. Nine specimens were
tested, and five produced valid results. The GIc calculated using Eq. (11) is
listed in Table 3. Because crack blunting was obvious and the macroscopic
response of interface failure was not perfectly linear, this data reduction
based on the LEFM only gives an estimated value, but this value may be very
close to the real cohesive energy.

3.3.2. COHESIVE STRENGTH


The cohesive strength is approximately equal to the adhesive strength of the
interface, and it can be measured using tension tests. We also assessed
the uniaxial tensile specimens following ASTM D 2095 [33], as outlined in
Fig. 6. The aluminum plates were added to the specimen to carry the load.
The propellant and insulation joints were in the same batch of the DCSB
specimen in order to make sure they had identical interface properties.

TABLE 3 Experimental Values of Critical Strain Energy Release Rate, Bonding


Strength, and Initial Stiffness

Specimens GIc [kJ=m2] rc [MPa] K [N=mm3]

1 0.61972 0.13870 1.09452


2 0.81970 0.15136 1.03739
3 0.75830 0.13403 0.99401
4 0.61947 0.15090 1.059335
5 0.60989 0.16383 1.10072
Average 0.68541 0.14776 1.05657
Standard deviation 0.09710 0.01174 0.04512
Debonding Simulation of the Propellant and Insulation Interface 241

FIGURE 6 Uniaxial tensile specimen.

The uniaxial tensile specimens have a cross-section of 25  10 mm2. The


aluminum plates were cleaned with ethanol, dried at 60 C for 2 h, and
attached to the prepared propellant and insulation joints using the same
acrylic modified epoxy adhesive as in the DCSB tests. After the specimens
had been allowed to solidify at room temperature for 24 h, the tests were per-
formed using the same electrical material testing system under displacement
control at a rate of 1 mm=min.
Adhesive strength of uniaxial tensile tests has the following expression:

Pc
rc : 12
A

Here, P is the critical load and A is the adhesive area.


Seven uniaxial tensile specimens were tested. The adhesive strength rc
of five specimens, calculated using Eq. (12), is reported in Table 3. The scat-
tering of the experimental data may have been caused by the small differ-
ences in the thickness of the liner. All the specimens showed cohesive
failure of the adhesive layer.

3.3.3. INITIAL STIFFNESS


Many different methods have been used to determine the initial stiffness.
These include calculating it by dividing the relevant elastic modulus by the
adhesive thickness [23,34] and setting an artificially high value, from
106 N=mm3[35] to 108 N=mm3 [36], to avoid interpenetration of the crack face.
In the present study, the former technique was used in relation to the mate-
rial properties of the adhesive.
The uniaxial tension tests of bulk adhesive specimens were conducted
on the same electrical material testing system with a constant displacement
rate of 1 mm=min. Five specimens were examined, and their initial stiffness
was calculated with the resulting elastic modulus and the adhesive thick-
ness of 0.2 mm. The thickness of the adhesive was measured in the
propellant=insulation joints. These results are shown in Table 3.
242 Q.-C. Zhou et al.

4. NUMERAL IMPLEMENTATION

In this section, the validity of the CZM was examined and a methodology for
obtaining the accurate parameters of the model is discussed. A user-supplied
subroutine (UMAT) can now be used in conjunction with the finite element
code ABAQUS1 to predict crack growth using the exponential CZM. Bilinear
law has been integrated into ABAQUS. First, the debonding of the propel-
lant=insulation interface is simulated with a DCSB specimen performed in
the experimental tests. Next, an inverse analysis procedure was proposed
to calibrate the cohesive parameters.

4.1. Simulation of Double Cantilever Sandwich Beam Specimen


In order to verify the applicability of cohesive zone models, a simulation of
DCSB specimen was performed. Based on the geometry and loading con-
ditions of the DCSB specimen, plane strain conditions were assumed to sim-
plify the analysis. The material properties used in the numerical simulations
are given in Table 1 (propellant), Table 2 (insulation), and Table 3 (liner).
This is in addition to the aluminum alloy, which has a Youngs modulus E
of 70,000 MPa and Poissons ratio n of 0.33. The propellant sheet and Al sheet
were modeled with four-node quadrilateral solid finite elements (CPE4R
from ABAQUS), and the insulation was modeled using the hybrid four-node
quadrilateral solid finite elements (CPE4RH) to show fully incompressible
behavior. The liner was modeled using four-node cohesive elements
(COH2D4). They were built as different parts, and both the propellant and
insulation parts are connected to the Al by a tie constraint.
Preliminary work was performed to determine the sizes of the elements
by taking both computational efficiency and the accuracy of the results
into account. The sizes of the elements in the different parts of the
model, including propellant, Al, and insulation, could be 1  1 mm. The
liner was discretized with a singer layer of cohesive elements through
the thickness, which have a length of 1 mm in the horizontal direction.
The DCSB specimen was loaded using an imposed displacement at the
rightmost vertex on the top edge, as in the experiment. The maximum dis-
placement was about 10 mm according to the load-displacements curves
obtained from the tests. Contact conditions were imposed on the propel-
lant=insulation interface to prevent interpenetration. The final deformed
shape of the DCSB specimen and the applied boundary and loading con-
ditions are shown in Fig. 7.
CZM models sometimes present convergence problems because the
cohesive stress decreases rapidly in reverse as soon as it reaches the peak
strength. Gao and Bower [37] proposed an artificial damping method with
additional energy dissipation to improve the convergence performance.
However, inappropriate damping coefficients may affect the accuracy of
Debonding Simulation of the Propellant and Insulation Interface 243

FIGURE 7 Final deformed shape of DCSB specimen with the applied boundary and loading
conditions. (color figure available online).

the simulations. In this study, we set the damping coefficient at 1E-4 based
the results of a large number of simulations.

4.2. Parameter Calibration


Figure 8 shows the simulations using the cohesive parameters obtained from
experimentation compared with the experimental results. As shown, there is
a large discrepancy between the numerical curves of the two cohesive laws
and the experimental curves. This may be because the parameters obtained
from the experiments were estimated values. To determine more precise
values, an inverse analysis procedure was performed using the experimental
values as starting values. Inverse analysis, which involves evaluating certain
parameters by comparing numerical predictions from CZM simulations to the
load-displacement curve of experimental tests, was here found to satisfy the

FIGURE 8 Experimental and numerical load-displacement curves.


244 Q.-C. Zhou et al.

FIGURE 9 Flow chart of the inverse procedure.

following object function [38]:

N  2
1X exp
min R PDsim / n ; r max  P D : 13
N i1 i i

Here, PDsimi
and PDexp
i
are the load in numerical and average experimental
load-displacement curves, respectively, when the displacement is Di. N is
the total number of collection points.
The Hooke-Jeeves algorithm [39] was used to solve the optimization
problem. This algorithm can be described as exploratory searches and heu-
ristic patterns executed in turn. An exploratory search, which involves sear-
ching along the two directions of the cohesive energy and cohesive strength
sequentially, was here used to determine the decrease direction of the objec-
tive function R. Then a pattern move was executed in the direction estab-
lished by the exploratory search. This allowed the function value to drop
very quickly. The flow chart of the inverse analysis procedures is given in
Fig. 9.

5. RESULTS AND DISCUSSION

In this section, the numerical results of the bilinear law and exponential
law were compared to the experimental results. Then, the sensitivity of the
Debonding Simulation of the Propellant and Insulation Interface 245

parameters capable of influencing the fracture response of the model was


addressed.

5.1. Comparison of Different Cohesive Laws to the Experimental


Results
The comparison between the experimentally determined load-displacement
curves and the numerical curves determined through inverse analysis are
shown in Fig. 10. Both the bilinear law and exponential laws are consistent
with the experimental data to a reasonable degree of accuracy. They show
the same initial slope, peak load, and trends. Although numerical curves do
not perfectly overlap the average experimental curve, they were still included
within the scope of experimental results (gray shaded area in Fig. 10). This
indicated that the CZM can simulate the crack nucleation and propagation
of propellant=insulator interface.
As shown in Fig. 10, the numerical curves are always slightly higher than
the experimental curve in the softening ascending stage, and the bilinear law
provides a better prediction than the exponential law. This is because the
exponential law introduces artificial compliance [40], while the bilinear law
eliminates it by allowing adjustment of the initial slope of the cohesive
law. The displacement when the load reaches its peak value of simulations
is smaller than that of the experiment results. In the subsequent descending
stage of the load-displacement curves, the bilinear law gives predictions that
lie closer to the experimental curves than those provided by exponential law.
The discrepancy between the numerical results and experimental data
decreases with the displacement.

FIGURE 10 Comparison between experimental and numerical results using calibrated


parameters.
246 Q.-C. Zhou et al.

TABLE 4 Model Parameters Obtained by Inversion

/n [kJ=m2] rmax [MPa] dn [mm] K [N=mm3]

Exponential 0.44 0.26 0.62


Bilinear 0.41 0.31 0.48 0.65

In views of the computational cost, the total numbers of increments and


iterations in the analysis of the bilinear law are 111 and 259, respectively. The
exponential law has a same number of increments as the bilinear law, 111.
However, it requires only 114 iterations. The exponential law turned out to
be more effective than the bilinear law.
Based on this discussion, we can conclude that the bilinear law is more
suitable for simulating fractures in the propellant=insulation interface in the
strict sense.
The cohesive parameters of the bilinear and the exponential laws are
reported in Table 4. The cohesive energy indicated by inverse analysis was
65% of the experimental value, which is as expected because the data
reduction method based on LEFM was roughly used. The cohesive strength
indicated by inverse analysis was twice as large as that observed in the
experimental results. The corresponding parameters of two laws are not
identical. The bilinear law has smaller cohesive energy and characteristic
length but a higher cohesive strength than the exponential law. The initial
stiffness K was less than the experimental value, which may have been
caused by the difference between the interface properties and the bulk
adhesive properties.
The two resulting layers are shown in Fig. 11. The ascending stage of the
bilinear law is similar to that of the exponential law. While in the evolution of
damage, they also have a similar trend when the separation is no more than

FIGURE 11 Comparison between bilinear law and exponential law.


Debonding Simulation of the Propellant and Insulation Interface 247

2 mm. After that, the traction of bilinear law decreased linearly, as before, but
the traction of exponential law gradually dropped to zero. The critical separ-
ation of the bilinear law is 2.645 mm. The traction of the exponential law is
less than 0.01 MPa at the separation of 4 mm, which means that cohesive ele-
ments fail almost completely. These values are not significantly different from
the 3.3 mm value observed in the experiments.

5.2. Sensitivity Study


In the previous section, numerical simulation indicated that the parameters of
CZMs can significantly affect the results of fracture analysis. Given the corre-
lation between cohesive parameters and material properties, the parameter
sensitivity was examined for evaluation of the influence of interfacial mech-
anical properties on the progression of debonding, for optimization of the
interface design. As discussed earlier, simulations of the bilinear law can give
better results than the exponential law. For this reason, the bilinear law was
applied to sensitivity analysis with respect to the cohesive energy, cohesive
strength, and initial stiffness.
The effect of cohesive energy on the loaddisplacement curve was stud-
ied and the results are shown in Fig. 12. Three different cohesive energies,
0.8/n, /n, and 1.2/n, are employed with a constant value of cohesive
strength (rmax) and initial stiffness (K). As the cohesive energy is increased,
the initial slope of the load-displacement remains unchanged, but the
maximum load was increased, producing a greater displacement relating to
the maximum load. In the experiments, we observed that the interfacial
debonding occurred in the softening ascending stage of the curve. For this
reason, debonding occurred less frequently when cohesive energy was large.

FIGURE 12 Effects of cohesive energy on macroscopic response to debonding.


248 Q.-C. Zhou et al.

The descending part of the curve is sensitive to cohesive energy. Increasing


cohesive energy allowed the interface to withstand higher load levels.
The effects of the cohesive strength on the load-displacement curve is
also discussed and the results are illustrated in Fig. 13. The cohesive strength
was found to be 0.8 rmax, rmax, and 1.2 rmax. Cohesive energy and initial
stiffness were kept at a constant value of /n and K, respectively. As the
cohesive strength increased, the initial slope of the load-displacement curve
remained the same and the maximum load increased considerably. In con-
trast to the cohesive energy, the cohesive strength showed a slight effect
on the displacement relating to the maximum load. The influence seemed
to be more significant in the vicinity of the maximum load.
Finally, the effect of initial stiffness on the loaddisplacement curve was
studied and the results are shown in Fig. 14. Four different initial stiffnesses,
0.8K, K, 1.2K, and a high value of 10,000, were used with a constant value of
cohesive strength (rmax) and cohesive energy (/n). The initial slope of the
curve was found to increase with the initial stiffness. As discussed earlier,
the cohesive energy and cohesive strength showed no effects on the initial
slope. Therefore, for these materials and combination, the initial stiffness
was the only parameter that determined the initial portion of the load
displacement curve. As shown in Fig. 14, a high penalty stiffness could cause
the numerical results to fail to match experimental results because it may
introduce a fictitious compliance before crack propagation.
In summary, interfaces with greater cohesive energy and cohesive
strength are less prone to debonding. The results indicated that the resistance
to cracks in the interface is proportional to the rate of release of critical strain
energy and adhesive strength. If either of these is invariable, smaller ratios
between cohesive energy and cohesive strength will produce a steeper
curve. In addition, the load-displacement curve is sensitive to the cohesive

FIGURE 13 Effects of cohesive strength on macroscopic response to debonding.


Debonding Simulation of the Propellant and Insulation Interface 249

FIGURE 14 Effects of initial stiffness on macroscopic response to debonding.

energy and the cohesive strength. This is consistent with the results of other
reports [41,42]. In this case, the initial stiffness must be carefully chosen rather
than set as a high penalty value.
The uniqueness of the set of parameters obtained by the inverse analysis
is worth noting. In the present study, the parameters were calculated by com-
paring the whole loaddisplacement curve of numerical simulations and
experiments. Different sets of cohesive parameters were found to yield
distinct curves. Those curves each have a deviation R relative to the average
experimental curve. In this way, the inverse analysis based on the Hooke-
Jeeves optimization algorithm was used. This algorithm can search within
a range and produce the best set of parameters corresponding to the lowest
R if it converges. In this way, the uniqueness of the parameters is guaranteed
by the algorithm itself.

6. CONCLUSION

In this paper, we present an initial attempt to realize the debonding process


of the propellant and insulation interface. The widely adopted CZMs were
used to characterize the adhesive interface between HTPB propellant and
EPDM insulation. The DCSB tests and uniaxial tensile tests were carried
out to determine the model parameters. From that, the ensuing calibration
of cohesive parameters was accomplished using the inverse analysis based
on the Hooke-Jeeves algorithm.
The finite simulation of DCSB specimen was performed to examine
the validity of the CZM implemented as a UMAT within ABAQUS. The
250 Q.-C. Zhou et al.

load-displacement curves predicted using CZMs with calibrated parameters


matched the experimental results closely. The results demonstrated that the
CZMs could accurately simulate the crack initiation and propagation of pro-
pellant=insulator interface in mode I. Although a particular damage evolution
in crack growth was found in those tests, the CZM appeared to be well suited
for this case.
The improved exponential law and bilinear law were used in this simu-
lation to take into account the shape of the cohesive laws. It was concluded
that the bilinear law was more appropriate than the improved exponential
law for simulating interface debonding in the strict sense. However, the
bilinear law was required much larger computational cost.
A sensitivity analysis of cohesive parameters, specifically cohesive
strength, cohesive energy, and initial stiffness, was performed using the
bilinear law. It was concluded that the initial slope of the load-displacement
curve was uniquely defined by the initial stiffness, and the curve was sensi-
tive to both cohesive energy and cohesive strength.

ACKNOWLEDGMENTS

The authors wish to thank Chang Wu-Jun of Nanjing University of Science


and Technology for general help with the simulations.

REFERENCES

[1] Kakade, S. D., Navale, S. B., and Narsimhan, V. L., J. Energ. Mater. 21, 7385
(2003).
[2] Navale, S. B., Sriraman, S., and Wani, V. S., Defence Sci. J. 54, 353359 (2004).
[3] Grythe, K. F., Hansen, F. K., and Olsen, T., J. Adhesion 83, 223254 (2007).
[4] Schloss, H. R., J. Adhesion 4, 333351 (1972).
[5] Gottlieb, L. and Bar, S., Propellant, Explos. Pyrotech. 28, 1217 (2003).
[6] Yan, Y. B. and Shang, F. L., Sci. China, Ser. G 39, 10071017 (2009).
[7] Elices, M., Guinea, G. V., Gomez, J., and Planas, J., Eng. Fract. Mech. 69,
137163 (2002).
[8] Dugdale, D. S., J. Mech. Phy. Solids 8, 100104 (1960).
[9] Barenblatt, G. I., Adv. Appl. Mech. 7, 55125 (1962).
[10] Song, S. H., Paulino, G. H., and Buttlar, W. G., J. Eng. Mech. 132, 12151223
(2006).
[11] Roy, Y. A. and Dodds, Jr., R. H., Int. J. Fract. 110, 2145 (2001).
[12] Ivankovic, A., Pandya, K. C., and Williams, J. G., Eng. Fract. Mech. 71, 657668
(2004).
[13] Srensen, B. F. and Jacobsen, T. K., Eng. Fract. Mech. 70, 18411858 (2003).
[14] Carlberger, T. and Stigh, U., J. Adhesion 86, 816835 (2010).
[15] Banea, M. D., da Silva, L. F. M., and Campilho, R. D. S. G., J. Adhesion 88,
534551 (2012).
Debonding Simulation of the Propellant and Insulation Interface 251

[16] Wang, J., Kang, Y. L., Qin, Q. H., Fu, D. H., and Li, X. Q., Comp. Mater. Sci. 43,
11601164 (2008).
[17] Needleman, X., J. Mech. Phys. Solids 38, 289324 (1990).
[18] Tvergaarsd, V. and Hutchinson, J. W., J. Mech. Phys. Solids 40, 13771397
(1992).
[19] Chandra, N., Li, H., Shet, C., and Ghonem, H., Internat. J. Solids Struct. 39,
28272855 (2002).
[20] Volokh, K. Y., Commun. Numer. Meth. Eng. 20, 845856 (2004).
[21] Alfano, G., Compos. Sci. Technol. 66, 723730 (2006).
[22] Geubelle, P. H. and Baylor, J. S., Compos. Part B-Eng. 29, 589602 (1998).
[23] Davila, C. G., Camando, P. P., and Moura, M. F., Mixed-mode decohesion ele-
ments for analyses with progressive delamination, 42nd AIAA=ASME=ASCE=
AHS=ASC Structures, Structural Dynamics and Materials Conferences, Seattle,
(2001).
[24] Xu, X. P. and Needleman, A., Modelling Simul. Mater. Sci. Eng. 1, 111132
(1993).
[25] van den Bosch, M. J., Schreurs, P. J. G., and Geers, M. G. D., Eng. Fract. Mech.
73, 12201234 (2006).
[26] Xu, J. S., Ju, Y. T., Han, B., Zhou, C. S., and Zheng, J., Mech. Time-Depent. Mat.
(DOI 10.1007=s11043-012-9203-z).
[27] Han, B., Ju, Y. T., Xu, J. S., and Zhou, C. S., J. Ballistics 24, 6368 (2012).
[28] ASTM D 3433. Standard test method for fracture strength in cleavage of
adhesives in bonded metal joints, (ASTM, Philadelphia, PA, USA, 2005).
[29] Pandya, K. C. and Williams, J. G., Polym. Eng. Sci. 40, 17651776 (2000).
[30] Irwin, G. R. and Kies, J. A., Weld. J. 33, 193198 (1954).
[31] Hashemi, S., Kinloch, A. J., and Williams, J. G., Compos. Sci. Technol. 37,
429462 (1990).
[32] IS0 15024:2001(E). Fibre-reinforced plastic compositesDetermination of mode
I interlaminar fracture toughness, GIc, for unidirectionally reinforced materials,
(ISO, London, 2001).
[33] ASTM D 2095. Tensile strength of adhesive by means of bar and rod specimens,
(ASTM, Philadelphia, PA, USA, 2005).
[34] Daudeville, L., Allix, O., and Ladeveze, P., Compos. Eng. 5, 1724 (1995).
[35] Camanho, P. P., Davila, C. G., and de Moura, M. F., J. Compos. Mater. 37,
14151438 (2003).
[36] Schellekens, J. C. J. and de Borst, R., Numerical simulation of free edge delami-
nation in graphite-epoxy laminates under uniaixal tension, Proceedings of the
6th International Conference on Composite Structures, (1991), pp. 647657.
[37] Gao, Y. F. and Bower, A. F., Modelling Simul. Mater. Sci. Eng. 12, 453463
(2004).
[38] Han, B., Ju, Y. T., and Zhou, C. S., Eng. Fail. Anal. 26, 306317 (2012).
[39] Hooke, R. and Jeeves, T. A., J. ACM 8, 212229 (1961).
[40] Espinosa, H. D. and Zavattieri, P. D., Mech. Mater. 35, 333364 (2003).
[41] Banea, M. D., da Silva, L. F. M., and Campilho, R. D. S. G., Int. J. Adhes. Adhes.
31, 273279 (2011).
[42] Sun, C., Thouless, M. D., Wass, A. M., Schroeder, J. A., and Zavattieri, P. D., Int.
J. Solids Struct. 45, 47254738 (2008).

Vous aimerez peut-être aussi