Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Damage Evolution Law

Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Damage Evolution Law

Related terms:

Energy Engineering, Cohesive Zone Model, Continuum Damage Mechanic, Co-


hesive Zone, Fracture Energy, Damage Model, Damage Variable, Failure Criterion,
Intralaminar

View all Topics

FAILURES OF MATERIALS
JEAN LEMAITRE, RODRIGUE DESMORAT, in Handbook of Materials Behavior
Models, 2001

6.14.3.1 ISOTROPIC CASE


The damage evolution law is given by

(13)

where is the plastic multiplier and S and s are two material- and temperature-depen-
dent parameters. The normality rule also defines the evolution of the state variable
r associated with R as

(14)

Then = is identified as the accumulated plastic strain rate multiplied by (1 - D).

Finally, the istropic damage law of evolution is

(15)

> Read full chapter


Predicting the fatigue life of adhe-
sively-bonded composite joints under
mixed-mode fracture conditions
P. Naghipour, in Fatigue and Fracture of Adhesively-Bonded Composite Joints, 2015

15.3.3 Maiti and Guebelle (2005)


Maiti and Geubelle (2005) developed a model that relies on the combination of a
bilinear cohesive failure law used for fracture simulations under monotonic loading
and a damage evolution law relating the cyclic degradation of the cohesive stiffness
with the rate of crack opening displacement and the number of cycles since the onset
of failure. The fatigue component of the cohesive model involves two parameters
that can be readily calibrated based on the classical log–log Paris crack growth
curve between the crack advance per cycle and the amplitude of the stress intensity
factor applied. The cohesive model, leading to similar unloading and reloading
paths in the traction–separation curve, prevents crack growth under subcritical cyclic
loading because of the progressive degradation of the cohesive properties in the
failure zone. This limitation suggests the need for an evolution law to describe
the changes incurred by the cohesive strength under fatigue. A phenomenological
model of such processes involves the progressive degradation of the cohesive zone's
strength during reloading events. Hence, under cyclic loading, the evolution law of
the instantaneous cohesive stiffness Kc, that is, the ratio of the cohesive traction, Tn,
to the displacement jump, Δu, during reloading is expressed as shown in Eqn (15.5).
The cohesive strength decays exponentially, and the rate of decay is controlled by the
parameter . Maiti and Geubelle define a power law relation as a function of number
of cycles, N, for , with mathematical constants and describing the degradation
of the cohesive failure properties.

(15.5)

The proposed evolution law for the cohesive model can also be expressed in terms
of the rate of change of Kc and then discretized in time steps, as follows:

(15.6)

The superscripts i and i + 1 stand for loading steps i and i + 1, respectively. As shown
in the above-mentioned relations, during the reloading phase, the cohesive stiffness
at each material point along the cohesive zone gradually decreases in proportion to
the increment in displacement of the crack opening. This proportionality factor
evolves with the number of cycles N and thus gives a measure of the total damage
accumulated during the degradation process.
> Read full chapter

Materials Modeling and Characteriza-


tion
M.V. Donadon, S.F.M. de Almeida, in Comprehensive Materials Processing, 2014

2.07.4.2.1 Mode I Delamination


The interfacial behavior for mode I opening ( I > 0) is given by

[126]

where the damage evolution law dI(w) is defined in terms of normal relative displace-
ments. For a bilinear constitutive law, the expression for dI(w) is given by

[127]

where and is the through-thickness interfacial strength in mode I. Kww is the


through-thickness interfacial stiffness given in terms of the adhesive through-thick-
ness Young's modulus, that is, Kww = Ezz/h0, where h0 is the initial thickness of the
element associated with the undeformed configuration.

The strain energy release rate associated with mode I delamination is defined by the
area underneath the stress–relative displacement, which in turn is defined by the
bilinear constitutive law. That is,

[128]

where wf is the relative displacement in which the interfacial stress in mode I is equal
to zero (complete decohesion). From eqn [128], wf can be written in terms of the
strain energy release rate as follows:

[129]

For a linear-polynomial constitutive law, the expression for damage evolution law
dI(w) is written as

[130]

with

[131]
The derivation of the damage law for a linear-polynomial constitutive law enforces
the areas under the stress–relative displacement defined for both bilinear and linear
polynomials to be the same.

The material behavior in compression ( I ≤ 0) is assumed to be linear elastic in order


to avoid element interpenetration:

[132]

> Read full chapter

Advances in Applied Mechanics


S.A. Silling, R.B. Lehoucq, in Advances in Applied Mechanics, 2010

8.3 Bond Breakage


Recall the definition of the scalar extension state ,

(8.11)

where is the identity state defined in (3.1). A useful example of a damage evolution
law is given by the following model:

(8.12)

where f (e, ) is a nondecreasing function of e, , and the maximum is taken over all
times up to t. Assume for simplicity that for a given motion, is a nondecreasing
function of time. Observing from (8.11) that

(8.13)

differentiating (8.12) with respect to time yields an equivalent damage evolution law
in terms of the rate:

(8.14)

where denotes the partial derivative of f with respect to e. It is helpful to introduce


a vector state defined by

(8.15)

so that (8.14) can be written as

(8.16)
provided is nondecreasing. A specific case of such a damage model is bond breakage
in tension, in which

(8.17)

where H is the Heaviside step function and is the prescribed bond breakage extension
for the bond . In this case,

and from (8.14),

Alternatively, the same damage evolution law can be defined in terms of the rate
through (8.16) with

Bond breakage in compression can be treated in a similar way.

> Read full chapter

MULTIPHYSICS BEHAVIORS
PIERRE LADEVÈZE, in Handbook of Materials Behavior Models, 2001

10.6.3.2.1 Remarks
• Two damage variables are used to describe the damage associated with matrix
microcracking and fiber-matrix debonding. They seem to account for all the
proposed damage kinematics, including that starting from an analysis of the
microcracks. Many works have established, experimentally or theoretically, a
relation between the microcrack density and our damage variable d, which can
be very useful for the identification of a damage fatigue model.
• What we call the single layer is the assemblage of adjacent, usual elementary,
plies of the same direction. The damage forces, being integral values through
the thickness of the single layer, can be interpreted as energy release rates. It
follows that the damage evolution law of the single layer is thickness-depen-
dent. For single layers which are not too thick, such damage evolution laws
include results coming from shear lag analyses. Consequently, the size effects
– observed, for example, in tension – are produced by both the single-layer
model and the interface model through a structure problem. This theory,
which is very simple, works very well for most engineering laminates; however,
it cannot be satisfactory for rather thick layers. A first solution is to modify the
damage evolution law, using the thickness as a parameter.
• The damage variables are active for [0°, 90°]n laminates even if the apparent
modulus does not change. The model predicts this hidden damage [15].
• For fatigue loadings, we introduce:(6)where ds and d's are the quasi-static
part of the damage defined by Eq. 4 or 5. dF and d'F denote the fatigue part
characterized by the following fatigue evolution laws:(7)where a, a' are two
material functions and [.] denotes the maximum value over the cycle.

> Read full chapter

An overview of continuum damage


models used to simulate intralaminar
failure mechanisms in advanced com-
posite materials
A. Forghani, ... R. Vaziri, in Numerical Modelling of Failure in Advanced Composite
Materials, 2015

6.5.2 Damage evolution laws


The second challenge in the continuum damage approach is writing the evolution
law for the multiple damage parameters. Evolution laws that govern the condition for
initiation and growth of damage parameters are usually written in terms of the strain
state in the layer. In composites, owing to the existence of various damage modes
and their potential interactions, defining the damage evolution laws for continuum
damage models becomes challenging.

A popular approach to formulate the evolution law is based on the hypothesis of


maximum damage dissipation that is analogous with the associated flow rule of
plasticity. In this approach, the evolution of each damage parameter is governed by
its work conjugate, the so-called energy release rate density parameters. Ladeveze
and co-workers have employed this hypothesis to derive the evolution law for in-
tralaminar damage (Ladeveze et al., 2000; Ladeveze, 1995). Although the hypothesis
of maximum damage dissipation offers a straightforward framework for formulating
the evolution law, it is a pure mathematical hypothesis that does not necessarily
represent the underlying physics involved in damage evolution. For instance, models
based on maximum damage dissipation leads to path-independent energy dissipa-
tion. In other words, according to this hypothesis, the energy dissipated during the
course of a damage process becomes independent of the load path taken to reach a
particular damage state.

A more general approach motivated by the nonassociated flow rule of plasticity


defines a dissipation potential function (which is different from the free energy
function) and the damage evolution law is written in terms of the derivatives of
that potential function with respect to each damage parameter. The choice of the
dissipation potential function depends on the available test data and the purpose
of the damage model (see Krajcinovic, 1985; Lemaitre et al., 2000; Lemaitre and
Desmorat, 2005, for instance). The hypothesis of maximum damage dissipation can
be reckoned as a special case where the Helmholtz free energy is taken to be the
dissipation potential function.

> Read full chapter

Constitutive Models of Polycarbonate


Weihong Zhang, Yingjie Xu, in Mechanical Properties of Polycarbonate, 2019

2.7 Conclusion
In this chapter, constitutive models of polycarbonate were presented. A dam-
age-based elastic–viscoplastic constitutive model was developed for polycarbonate
within the framework of irreversible thermodynamics and CDM. It was able to char-
acterize the mechanical behavior and damage evolution of the material over a wide
range of strain rates and temperatures. The constitutive equations, specifically the
viscoplastic flow rule and damage evolution law, were derived from the Helmholtz
free energy and the damage-coupled plastic flow potential. To realize the numerical
implementation of the proposed model, a coupled elastic-damage predictor and
viscoplastic corrector algorithmic scheme was introduced and further simplified
into a single-scalar Newton–Raphson scheme. The model was implemented into
LS-DYNA by using the UMAT subroutine. Finally, numerical examples, including
both the single-element and the cylinder specimen, were given. The good correlation
between the numerical predictions and the experimental data has demonstrated
the capabilities of the proposed model in capturing the mechanical behavior of
polycarbonate over a wide range of strain rates and temperatures.

> Read full chapter

Multi-scale modeling of high-temper-


ature polymer matrix composites for
aerospace applications
S. Roy, in Structural Integrity and Durability of Advanced Composites, 2015
6.3.2 Damage evolution law for micromechanical representa-
tive volume element
Within the cohesive layer RVE, damage initiation (i.e., initiation of voids and/or
polymer fibrils) is assumed to occur if the applied principal stretch along the fibril
exceeds the critical value of the principal stretch, that is,  >  cr. Because the change
of fibril diameter as a function of time is proportional to the applied principal stretch
along the polymer fibril, a phenomenological power law-based damage evolution
law is adapted (Upadhyaya, Singh, & Roy, 2011), given by

(6.5)

where is a principal stretch measure within the RVE (Upadhyaya et al., 2011), and 0
and m are material constants that are assumed to be dependent on environmental
conditions but independent of the applied strain rate. These material constants are
actually evaluated in this chapter by performing fracture experiments using DCB
specimens in conjunction with DIC, as described in Section 6.5.

> Read full chapter

Multiaxial fatigue of a unidirectional


ply
I. Koch, M. Gude, in Fatigue of Textile Composites, 2015

6.3.1.3 Parameter identification


As already mentioned above, for textile reinforcements with a significant amount
of z-reinforcement a separated single constitutive layer is not available by nature,
and therefore engineering constants, strength values and degradation parameters
of the constitutive idealised unidirectional layers cannot directly be elaborated. Here,
a set of fatigue experiments and an appropriate parameter identification method is
provided.

All layer-wise calculations presented in this paper are based on a modified classical
laminate theory with the structural law written as

(6.15)

By setting the coupling matrix to zero, the specimen geometry and the unsym-
metrical layup are considered (Koch, 2010). The differences in calculated stress
distribution in the inner and outer layer for tube specimens in tension loading
comparing three different mathematical approaches is displayed as an example in
Figure 6.9. The stress distribution in the undisturbed area is taken to be comparably
within the context of this paper, and the fastest method (CLT mod.) is used.

Figure 6.9. Comparison of the calculated stress distribution in cross-ply tube speci-


mens under tension loading (3D layer theory (Kroll, 1992); modified laminate theory
(CLT), FE model).

The engineering constants E1, E2, G12, 12 and strength values R|| , R|| , R , R
and R|| of the undamaged i-UD layer (Table 6.2) have been derived from the inverse
laminate theory and quasi-static tests until damage initiation and total failure (Böhm
et al., 2011). The remaining values are adapted from typical values from literature
and comparable in-house material data.

Table 6.2. Engineering constants and strength values, i-UD layer GF-MLG/EP

E1 (MPa) E2 (MPa) E3 (MPa) G12 G13 G23 12 (–) 13 (–) 23 (–)


(MPa) (MPa) (MPa)
24,000 5800 5800 2700 1900 2700 0.27 0.27 0.45

R|| (- R|| (- R (- R (- R|| μ (–)


MPa) MPa) MPa) MPa) (MPa)
482 800 32 75 25 0

The identification of the remaining parameters for damage growth ci, coupling qi
and tension–compression asymmetry hi are derived iteratively from the elaborated
stiffness degradation curves of the cross-ply-reinforced tube specimens (type S) only.
In a first step the stiffness degradation curve for loading path 1 (tension–tension
fatigue) and here the characteristic constant stiffness drop is focussed. In the frame-
work of the pursued top-down approach, this stage of stiffness degradation can be
dedicated to damage in the 0° layer only. Here damage due to inter-fibre failure in
the 90° layer reached saturation already. Hence the damage evolution law Eqn (6.5)
can be idealised to the damage growth in fibre direction

(6.16)
Hence the parameters of the growth function || can be elaborated by analysing
the constant stiffness drop in pulsating tension loading. It is proposed to analyse
the constant stiffness drop of several tension fatigue experiments on different load
levels. By fitting the chosen damage growth function to the experimentally derived
dependence of damage increment and material effort for all results, the parameters
are elaborated more precisely.

For the parameter identification of the first stage of stiffness degradation in pulsat-
ing tension fatigue is focused. Following the top-down approach in this stage, two
failure modes are active and the damage evolution law can be rewritten to

(6.17)

Hence with the given damage growth function || the first-stage stiffness degra-
dation can be used for the identification of the parameters of . The final failure of
the material under pulsating tension loading is driven by damage localisation and
instable crack growth. The beginning of damage localisation is modelled with the
help of a critical damage value Dc1 where the calculation is stopped.

A similar approach is used for the identification of . Assuming identical perpen-


dicular layers, the damage evolution law for plane shear fatigue loading can be
formulated as:

(6.18)

Here the first and second stage of stiffness degradation of pulsating torsion tests is
used. The results of often recommended ±45° specimens under pulsating tension
loading cannot be used because of parallel-acting damage modes such as transverse
tension and compression. In Table 6.3 the identified parameters for the damage
growth of GF-MLG/EP are displayed.

Table 6.3. Parameters for the damage growth of GF-MLG/EP

Failure mode Equation Model parameter Value


IFF1 (6.12) C1 7E-5

Ds 0.4

IFF2 (6.11) C2 8E-12

C3 12

Dc2 0.4

FF1 (6.11) C4 6E-10

C5 18.46

0.11
Dc1

The remaining parameters of coupling, , are adjusted to model the differences in


stiffness degradation under combined tension–shear to pure tension and pure shear
fatigue loading. The experimental results suggest that the micromechanically sub-
stantiated coupling of fibre damage to shear stiffness is not measurable. Therefore
is assumed. The achieved set of remaining parameters is given in Table 6.4.

Table 6.4. Parameters of coupling and passive damage

Model parameter Value


1.8
0.3
0.5
h1 = h2 0.5

For the identification of the parameters h1 and h2, combined compression–shear


fatigue experiments have been performed. As described above, a significant differ-
ence in the degradation under pure compression to combined compression–shear
loading has been observed. The parameters h1 and h2 are set equal and adjusted
to model the difference. In Figure 6.10 the normalised stiffness degradation curves
for tension, shear and combined tension–shear and compression–shear loading are
displayed together with the theoretical results after identifying all model parame-
ters as described above. For better visualisation of scatter, the elaborated stiffness
degradation is given in the form of a range rather than as single curves.
Figure 6.10. Comparison of experimental and analytical stiffness degradation used
for parameter identification.

> Read full chapter


Modelling the structure and behaviour
of 2D and 3D woven composites used in
aerospace applications
D.S. Ivanov, S.V. Lomov, in Polymer Composites in the Aerospace Industry, 2015

2.4.3 Local stiffness degradation and damage evolution law


We consider damage as a measure of property degradation. The complexity of
textile architectures does not allow making a flawless choice of degradation scheme.
Analytical and numerical models developed for the flat laminates clearly show that at
a given crack density, ply degradation depends on the stiffness of constraining plies,
the thickness of a cracked ply and the ply position in a laminate. The situation with
the textile composite is more complicated. The degradation of a particular segment
may also depend on inter-ply nesting configuration, segment location and crack
length. The implementation of a sophisticated multiparametric damage model does
not make sense unless all these peculiarities are taken into account. Currently, the
error introduced by the simplifications of actual reinforcement geometry exceeds all
possible benefits of employing a complex damage model.

To capture the stiffness evolution, it seems reasonable to follow the physically found-
ed and very simple idea of Zinoviev [101], who suggested that the effective transverse
and shear stresses in the damaged ply do not exceed the level at which cracks
initiates. The model of Zinoviev proved to be one of the five most successful models
of the World-Wide Failure Exercise (WWFE) I [102]. Another successful degradation
model of WWFE, the model of Puck [103], employs a similar post-cracking assump-
tion. It requires reducing the elastic constants in a way that keeps the failure initiation
index at a constant level. Upon active loading, both the models exhibit a stress–strain
response resembling the curve of an ideal plasticity. Both the damage models do
not require any material constants other than those needed for the failure initiation
prediction.

Zinoviev's model is formulated as an elaborated conditional algorithm distinguish-


ing various possible situations in loading and unloading. To overcome the difficulti-
es of implementing this algorithm in multi-segment textile architectures, we marry
this concept with damage evolution laws offered by continuous damage mechanics.
A single damage evolution law can then be obtained.

Following the considerations of Murakami and Ohno [104], damage is introduced as


a second rank tensor. For a case of an open crack (transverse stress is tensile), energy
equivalence of cracked and homogenous damaged media imposes the following
reduction of the elastic constants:
(2.6)

where E and G are Young's and shear moduli, and d2, d12 degradation of the elastic
constants, index ‘0’ corresponds to intact material, coordinate ‘1’ denotes the fibre
direction, 1–3 is the plane of the crack defined by means of the Mohr-Hashin-Puck
failure initiation criteria [105]. Without contradicting available experimental data, it
is postulated that all the cracks in a segment have the same orientation [106]. All the
other elastic constants remain unaffected by cracking. The symmetry of the stress
tensor in the damaged media leads, as described by Zako et al. [46], to the following
relation between degradation of the Young's and shear moduli:

(2.7)

The damage factors d2 and d12 are nearly proportional at small values, and the
transverse degradation is slightly higher at the advanced damage levels. Following
the continuum damage models [107], we employ the energy release rate Y12 as a
factor driving damage accumulation. It is introduced as the derivative of the elastic
energy of the damaged material with respect to the degradation parameter:

(2.8)

where operator denotes the maximum over the time period t and reflects the fact
that the damage state is non-healing. The square root is used for convenience, to
make it proportional to strain. The elastic energy may be written as a function of
the damage parameters d2, d12, the elastic stiffness matrix of intact material C0,
and the strains averaged over the segment volume . This allows differentiating the
expression explicitly, which results in the following definition:

(2.9)

In the case of a closed crack (transverse compression), the normal stress traction
is assumed to be continuous across the crack face. Hence, the degradation factor
d2 = 0 and . Zinoviev's and Puck's assumptions are in good correspondence with the
following damage evolution law:

(2.10)

where indexi reflects the value of a variable at the moment of failure onset and .
Equation (2.10) matches the shear diagram assumed in [101] – Figure 2.8(a), as it is
reduced to in pure shear. The stress–strain responses in other loading situations are
obtained from the solution of Eqns (2.8) and (2.10) relative to d12. Figure 2.8(b) shows
predicted response in uniaxial transverse tension. Right after the damage onset, the
stress drops and before settling down at a value of about 80–90% of transverse
strength.
Figure 2.8. Stress–strain diagrams of an elementary unidirectional fibre bundle in
(a) pure shear (input), (b) pure transverse tension

Such a law has an analogy with classical fracture mechanics, where energy released
through crack propagation is linked to the crack length. The meaning of Eqn (2.10)
is that the stiffness reduction is proportional to the energy, which can be released
through damage accumulation.

As far as the numerical implementation is concerned, it is required to (1) calculate


the damage parameters at a given strain state, (2) redefine the elastic constants
according to Eqn (2.1). An iteration procedure is needed to ensure convergence and
satisfaction of Eqns (2.8) and (2.10) at the same time. The procedure is found to be
insensitive to load step size and defragmentation of yarns into segments.

> Read full chapter

ScienceDirect is Elsevier’s leading information solution for researchers.


Copyright © 2018 Elsevier B.V. or its licensors or contributors. ScienceDirect ® is a registered trademark of Elsevier B.V. Terms and conditions apply.

You might also like