Anda di halaman 1dari 11

Composite Structures 94 (2012) 33363346

Contents lists available at SciVerse ScienceDirect

Composite Structures
journal homepage: www.elsevier.com/locate/compstruct

Predictions of crack propagation using a variational multiscale approach


and its application to fracture in laminated ber reinforced composites
Shiva Rudraraju a, Amit Salvi b, Krishna Garikipati a, Anthony M. Waas a,b,
a
b

University of Michigan, Department of Mechanical Engineering, 2350 Hayward Street, Ann Arbor, MI 48109, USA
University of Michigan, Department of Aerospace Engineering, FXB Building, 1320 Beal Avenue, Ann Arbor, MI 48109, USA

a r t i c l e

i n f o

Article history:
Available online 11 April 2012
Keywords:
Variational multiscale
VMCM
Cohesive crack propagation
Curved crack
Mixed mode
Laminated ber composite

a b s t r a c t
Predictions of crack propagation is a valuable resource for ensuring structural integrity and damage tolerance of aerospace structures. Towards that end, a variational multiscale approach to predict mixed
mode in-plane cohesive crack propagation is presented here. To demonstrate applicability and to provide
validation of the nite element based predictive methodology, a comparative study of the numerical
results with the corresponding experimental observations of crack propagation in laminated ber reinforced composite panels is presented.
2012 Elsevier Ltd. All rights reserved.

1. Introduction
The aerospace industry relies on extensive testing to ensure the
structural integrity and damage tolerance of vehicle structures, thus
leading to a substantial increase in the total design cost of many
aerospace vehicles. These costs can be reduced by developing high
delity computational models which can provide valuable information regarding the performance of a structure up to and including
failure, provided the modeling is based on material parameters that
can be measured, and is validated using laboratory tests that are
designed to be discriminatory. The nite element method is a key
computational tool in solid mechanics and has become the mainstay
of problems involving any of the broad phenomena of material
deformation elasticity, plasticity and damage. However, its utility
for problems of crack propagation has met with mixed success. The
distinguishing characteristic of crack problems, in general, is the formation and propagation of boundaries, which are not part of the original boundary value problem. These boundaries emerge from
within the domain of the problem conguration, under load, rendering a departure from a traditional continuum description for the
initially solid body, at least across the newly formed boundaries.
This is not an obstacle, if the resulting crack path is known a priori,
and the mesh is ensured to have elemental surfaces align along possible crack surfaces; but often, neither conditions are feasible. For all
but trivial crack propagation problems, the crack path is not known
Corresponding author at: University of Michigan, Department of Mechanical
Engineering, 2350 Hayward Street, Ann Arbor, MI 48109, USA. Tel.: +1 734 764
8227.
E-mail address: dcw@umich.edu (A.M. Waas).
0263-8223/$ - see front matter 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.compstruct.2012.03.035

beforehand and has to be determined as part of the solution process,


and in structural level problems adaptive mesh generation/realignment is prohibitively costly. The traditional Galerkin FEM implementation is not suitable for problems involving strain localization
and cracks, as it leads to mesh subjective schemes, and the related
limitations have been well documented in the context of spurious
mesh related length scales [3,6,1] and requirements of mesh alignment relative to the localization band [13,20]. Further in the context
of cracks, the global non-linearity of the load response which depends on the microstructure of the material, also requires new
constitutive relations which can span across different length scales.
These additional cohesive relations between the crack face opening
and its internal tractions, referred to as tractionseparation relations, lead to the more challenging class of cohesive cracks and
bridging cracks, where the crack surface may be a diffused zone of
damage rather than a sharp boundary.
In this context, cohesive zone models, which embed fracture
process zone mechanics through non-linear tractionseparation
relationships across the crack faces become an important tool for
analysis [19,29,27,24,33,5,32,31]. However the numerical implementation of cohesive zone modeling is often through surface
elements which are placed along the expected crack path, limiting
crack growth studies to cases, where the crack path is known a priori.
This major limitation can be overcome by developing continuum
elements that can simultaneously be used for both continuum and
non-continuum (severed by a crack or cracks, leading to a discontinuity in the displacement eld) modeling. Development of such continuum elements with discontinuities (enriched nite elements)
have gained increasing interest in recent years. Depending on the
support of the enriching discontinuous displacement modes, the

3337

S. Rudraraju et al. / Composite Structures 94 (2012) 33363346

enhanced nite elements are popularly classied as element enrichment methods [2,9,12,4,17,16,10] and nodal enrichment methods
like XFEM [15,14,30]. Interested readers are referred to Oliver
et al. [18], for detailed discussion and comparison of these methods.
The primary task of this paper is presenting a new multiscale framework, referred to as the Variational Multiscale Cohesive Method
(VMCM), to embed cohesive models into continuum elements
through discontinuous enrichment of the shape functions and demonstrating its effectiveness by simulating mixed-mode failure in ber-reinforced composites.
The paper is organized as follows: Section 2 provides the motivation for a multiscale approach to capture displacement discontinuities involved in crack propagation. Then, Section 3 presents the
variational multiscale formulation for modeling cohesive crack
propagation, followed by details of the nite dimensional implementation in Section 4. Simulations and experimental validation
of the multiscale approach are provided in Section 5 and the concluding remarks follow in Section 6.
2. Crack propagation as a ne scale problem
Physically, crack propagation is a process of congurational
change by which new surfaces are created. The creation of new surfaces is governed by surface laws, different from the constitutive
laws of the continuum. Classically, this process of surface creation
is handled by affecting changes in the numerical discretization,
involving incremental grid renement and remeshing. However,
changing the grid to reect the evolving domain boundaries is computationally very expensive. Instead, an alternative view of cracks
as displacement discontinuities in the continuum domain is considered here. The concept of discontinuous displacement elds and the
resulting singular strains nds its mathematical treatment in the
work of Temam and Strang [28] on BD(X), the space of bounded
deformations for which all components of the strain are bounded
measures. This idea was used to develop a numerical framework
for the problem of strong discontinuities due to strain localization
by Simo et al. [26], Simo and Oliver [25] and Armero and Garikipati
[1]. The physical process of strain localization involves localized
changes in the continuum constitutive response and no new
boundaries and surface laws appear, but its numerical treatment
introduced the use of the distributional framework and discontinuous basis functions, which was adopted in Garikipati [8] for embedding micromechanical surface laws into a macroscopic continuum
formulation, albeit in a multiscale setting. The presentation in this
work follows and extends these multiscale arguments specically
for numerical representation and evolution of cohesive cracks.
As shown in Fig. 1, a crack opening can be represented by a discontinuous displacement eld over an uncracked body. The use of
zero volume elements (interface elements, standard cohesive zone
elements, etc.), to capture such a discontinuous displacement eld,

is widely known to render the scheme subjective to the numerical


discretization. Instead the variational multiscale method, introduced by Hughes [11], is adopted here in the context of crack propagation, and the cracks are represented by a ne scale
discontinuous eld superposed on a coarse scale deformation eld.
3. Multiscale formulation of discontinuous displacement
The weak formulation of the quasi-static elasticity is the point
of departure for the multiscale development. Also, the scope of
the presentation is limited to the innitesimal strain theory of elasticity. Starting with the weak form: For S  BDX and V  H1 X,
nd u 2 S fv jv g on Cg}, such that 8 w 2 V fv jv 0 on Cg},

rw : r dV

w f dV

w T dS

Ch

where f is the body force, g and T are the prescribed boundary displacement and surface traction, respectively. r is the (Cauchy)
stress tensor given by r = C:sym(ru), where C is the fourth-order
elasticity tensor.
Now, scale decompositions of u and w are introduced. The
decompositions are qualied by requiring that the ne scales, u0
and w0 , vanish outside the neighborhood of the crack path, which
is contained in X0 (Fig. 2), referred to as the microstructural or
ne-scale subdomain


u
|{z}

|{z}
u0

2a


w
|{z}

|{z}
w

2b

coarse scale

coarse scale

fine scale
0
fine scale

 2 S fv j
u

2c

 2 V fv j
w

v g on Cg g
v 0 on Cg g
u0 2 S 0 fv j v 0 on X n intX0 g
w0 2 V 0 fv j v 0 on X n intX0 g

2d
2e
2f

where S S  S 0 and V V  V 0 . Further, V and V 0 are chosen to be


linearly independent.
Given the scale decomposition of u and w, we can be split Eq.
(1) into two separate weak forms

 : r dV
rw

X0

 f dV
w

rw0 : r dV

X0

 T dS
w

3a

Ch

w0 f dV

w0 T dS

C0h

3b

Now consider a crack surface, Cc, in the ne-scale subdomain


(Fig. 2). Assuming no body force in the ne-scale subdomain, using
integration by parts and standard variational arguments, Eq. (3b)
can be reduced to,

Z
Cc

Fig. 1. Representation of crack as a displacement discontinuity. sut is the


magnitude of the displacement discontinuity which physically represents the
magnitude of the crack opening and Cc is the crack surface.

w0 r  n dS

Z
Cc

w0 Tc dS

W0

Fig. 2. The microstructural domain, X0 , and the crack surface, Cc. Shown in the inset
are the crack orientation vectors and the crack surface traction.

3338

S. Rudraraju et al. / Composite Structures 94 (2012) 33363346

Fig. 3. Construction of the discontinuous multiscale shape function in 1D.

(a)

(b)
Fig. 4. Possible constructions of the discontinuous multiscale shape function in 2D. n is the normal to the crack path in the direction of the desired jump in displacement.

Fig. 5. The crack path is chosen along the direction of minimum average shear traction along a nite length ahead of the crack tip. Stress, strain and/or strain energy metrics
are needed in the neighborhood of the crack tip for mesh insensitive path prediction.

where Tc is the external traction on the crack faces. In the subsequent sections, Equations W and (W0 ) are referred to as the coarse
scale and ne scale weak forms, respectively.
3.1. Fine-scale eld and micromechanics embedding
Equation (W0 ) allows us to embed any traction based cohesive
surface-law, Tc, into the continuum formulation. Writing the trac-

tion on Cc in terms of the components T cn and T cm along n and m


respectively

Tc T cn n T cm m

The ne scale eld, u0 , for crack problems is composed of a displacement discontinuity, sut, which can be expressed in terms of
the components sunt and s umt along n and m respectively,

3339

S. Rudraraju et al. / Composite Structures 94 (2012) 33363346

(a)

(b)

(c)
Fig. 6. Mesh objectivity for straight crack propagation. (A) Symmetrically loaded SETB specimen, (B) displacement magnitude contours for different mesh densities, and (C)
corresponding loaddisplacement response. The P and D values have been normalized with xed reference values. Mesh objectivity for curved crack propagation in
eccentrically loaded SETB specimen has been shown in Rudraraju et al. [23].

sut sun t n sum t m


|{z}
|{z}
opening

shear

sunt and sumt are referred to as the crack face opening displacement
and crack face shear displacement, respectively. Similarly, the crack
face opening mode is referred to as Mode-I and crack face shear
mode is referred to as Mode-II. Now consider a simple micromechanical surface traction relation given by:

T cn T cn0  Hn sun t

7a

T cm T cm0  Hm sum t

7b

where T cn0 and Hn are the Mode-I critical opening traction and
Mode-I softening modulus, and T cm0 and Hm are the Mode-II critical
shear traction and Mode-II softening modulus. Using Eqs. (5) and
(7), u0 (characterized by sut) can be eliminated from Equation
 . Once u
 is obtained it can be
W, which can then be solved for u
used to recover u0 , thereby determining the complete displacement

eld. This procedure is demonstrated in a nite element setting in


the following section.
4. Finite dimensional weak form and discretized equations
In the nite dimensional setting, the problem domain is divided
S
h
into non-overlapping elements such that X nel
1 Xe , where nel is
the number of elements. Restricting the presentation to two
dimensions and triangular elements, a specic re-parametrization
of the displacement eld (Eq. (2a)) in terms of local interpolation
functions is given by,

 he n; g
u

n
node
X

NA n; gde

8a

A1
c
u0h
e n; g M C n; gsute

8b
A

where (n, g) are the iso-parametric coordinates, de is the nodal value


of the coarse-scale displacement and sute is elemental value of the

3340

S. Rudraraju et al. / Composite Structures 94 (2012) 33363346

(a)

(b)

(c)
Fig. 7. Mesh objectivity for straight crack propagation. (A) CTS specimen, (B) displacement magnitude contours for different mesh densities, (C) corresponding load
displacement response. The P and D values have been normalized with xed reference values.

Table 1
Lamina and laminate properties of carbon ber/epoxy [45/0/+45/90]6s laminated
ber reinforced composite.
Laminate

Lamina

Exx: 51.5 GPa


Eyy: 51.5 GPa
Gxy: 19.4 GPa
mxy: 0.32

E11: 141 GPa


E22: 6.7 GPa
G12: 3.2 GPa
m12: 0.33

Ou0 OMCc sut


where

"
sut

OM Cc
ne-scale displacement discontinuity. N is the usual linear shape
function for triangle elements, and MCc is a multiscale shape function given by,

M Cc N  HCc

where HCc is a Heaviside function which has its discontinuity on Cc.


Thus, MCc is a composite shape function constructed by superposing
a Heaviside function on a linear shape function, ensuring that
MCc 0 on X n intX0 . This construction is depicted in Figs. 3 and 4
for 1D and 2D, respectively. In the numerical implementation,
OMCc enters the system of equations through the expression for
Ou0 , which in matrix form is given by

10

sutx

suty
2

3
2
3
nix 0
nx 0
16
7
6
7
i 4 0 niy 5 dCc 4 0 ny 5
h
i
i
ny nx
ny nx
|{z}
|{z}
G

G and H are the matrix representation of ni and n, respectively.


Now the expressions for strain and stress in the nite dimensional setting are given by,

e Bd G  dCc H sut
r C : Bd G sut

11a
11b

where B is the standard matrix form of the shape function gradients. Substituting the above expression for stress in Eqs. (3a) and
(4), the respective nite dimensional coarse scale and ne scale
weak forms are given by,

S. Rudraraju et al. / Composite Structures 94 (2012) 33363346

K u0 u H T CB
T

K u0 u0 H CG Hn n  n Hm m  m

3341

15c
15d

5. Simulations and comparison with experiments

Fig. 8. Cross section of the [0/45/90/45]6s specimen layup.

Table 2
Scaling observed in SETB specimen experiments. The P, D, and GIc values were
normalized with xed reference values.
Size

Geometry
scaling (Fig. 9)

Peak
load P/P

Load point
displacement D/D

Fracture
toughness
GIc =GIc

1
2
3
4
5

1
1.5
2
3
4

0.27
0.4
0.6
0.81
1.0

0.1
0.15
0.2
0.28
0.37

1.08
1.23
1.84
2.46
2.58

BT C : Bd G sut dV

Nf dV

N T dS

12a

Ch

H T C : Bd G sut T c

12b
c

where the ne-scale weak form is reduced to r  n = T , as for linear


triangles both r and sut (and hence Tc) are constant over the element. Also, it should be noted that only the regular portion of the
strain contributes to the stress.1 To suit an iterative solution procedure, the above equations are expressed as coarse-scale and nescale residuals,

r

BT C : Bd G sut dV 

Z
X

r0 H T C : Bd G sut  T c

Nf dV 

N T dS

13a

Ch

13b

Linearizing the above residuals about d and sut and rearranging


terms, we obtain the following system of equations in (dd, dsut)

K u u

K u u0

K u0 u

K u0 u0



dd
dsut

r
r 0


14

where

K u u
K u u0

Z
ZX

BT CB dV

15a

BT CG dV

15b

1
We choose to represent cracks as displacement discontinuities, which implies
u R C0. This results in the strain being a singular distribution which has a bounded
measure, since u 2 BD(X). However the stress should not be a singular distribution as
required by the classical jump condition on the traction (sr  nt = 0). This requirement
on the stress eld is enforced by the material constitutive response which mollies
the singular strains to yield regular stresses [26].

This section presents numerical simulations to demonstrate


mesh objectivity of the multiscale formulation and comparisons
of numerical and experimental observation of crack propagation
in laminated ber reinforced composite panels. The inputs to the
numerical model are the elastic and cohesive constitutive models.
The cohesive models are essentially the tractionseparation laws
for each of the fracture modes, which in 2D are the standard
Mode-I and Mode-II. The method easily supports mode-mixity
conditions relating Mode-I and Mode-II evolution, but given the
complex micromechanics and limited experimental data, only the
simplied case of locally Mode-I evolution has been considered
in the subsequent simulations. This implies that the crack always
grows along the direction of zero shear stress and the local
Mode-II fracture toughness is assumed to be zero. So once the
crack opens in Mode-I, it is assumed that there will be no shear
resistance in its wake.
The crack direction criterion used to advance the crack path is
very critical in determining the resulting global crack path and
the load response. The incremental crack advancement can be
along the direction of maximum strain energy dissipation, or along
the direction of maximum normal tractions or minimum shear
tractions. In this work, the crack path is chosen along the direction
of minimum shear traction. Further, to minimize any local aberrations due to insufciently resolved stress elds ahead of the crack
tip, the average shear tractions acting along a nite length ahead of
the crack tip in several directions are computed and the path with
the minimum average shear traction is selected as the instantaneous crack direction (Fig. 5). The nite length over which the
average shear traction is computed should be such that it encompasses sufcient number of elements ahead of the crack tip. From a
code implementation point, this requires that one has access to the
stress elds of all the elements in the vicinity of the crack path.
Thus, the decision making process implemented in the present
work is non-local. Experiments show that crack path in ber composites is mostly along some of the ber directions. This physical
observation can be simulated by restricting the numerical crack
path along a ber direction which is closest to the direction of
the average minimum shear traction.
It is to be noted out that the apparent distortion of the elements
representing the crack path in the simulations may be seen as contradicting the innitesimal strain assumption of linear elasticity.
This is not the case as the regular part of the strain in the cracked
elements is consistent with the global strain eld, and the singular
components which lead to this observed element distortion do not
contribute to the stressstrain constitutive relation. All the simulations were carried out using an in-house, C++ based, nite element
code. A standard NewtonRaphson scheme was used for solving
the system of non-linear equations, based on a direct solution procedure using the SuperLU library [Demmel et al. [7]]. Also, for a
better visualization of the crack path, the crack path elements
are removed from the simulation images, except in Section 5.1.
5.1. Mesh objectivity demonstration
Eliminating pathological mesh dependence of crack propagation simulations was one of the primary motivations for the development of the multiscale framework, and this section seeks to
demonstrate the mesh objectivity of this implementation. The results presented here focus on the dependence of the global load

3342

S. Rudraraju et al. / Composite Structures 94 (2012) 33363346

Fig. 9. SETB specimen congurations used in the crack propagation experiments. Based on the load point location, the congurations are classied as symmetric or eccentric.
For symmetric specimens, size-1 has the dimensions shown in the gure and other sizes are scaled versions of this base size. For the eccentric specimens, size-1 has the
dimensions shown in the gure and size-2 is scaled up by a factor of two. All specimens have a nominal thickness of 6.35 mm.

Size 1
Size 2
Size 3
Size 4
Size 5

0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0

0.1

0.2

0.3

0.4

0.5

0.6

Fig. 10. Experimental loaddisplacement curves obtained for various sizes of SETB specimens subjected to symmetric loading conditions. Multiple specimens of each size
were tested to capture the envelope of the failure response. The P and D values have been normalized with xed reference values.

0.3

0.4

0.2
0.2

0.1
0

0.05

0.1

0.15 0.2 0.25 0.3 0.35

0.4 0.45

0.6

0.8

0.4

0.6

0.1

0.2

0.3

0.4

0.5

0.4

0.2
0

0.2
0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

1
0.5
0

0.2

0.4

0.6

0.8

Fig. 11. Loaddisplacement response obtained from simulations of symmetrically loaded Size 1-5 SETB specimens. For a particular specimen size, GL and GH are the least and
highest values of fracture toughness obtained from the multiple experimental loaddisplacement curves. GL corresponds to the curve exhibiting least toughness and GH
corresponds to the curve exhibiting the highest toughness. The P and D values have been normalized with xed reference values.

S. Rudraraju et al. / Composite Structures 94 (2012) 33363346

3343

3
2.5
2
1.5
1
0.5
0

6
5
4
3
2
1
0

10

Fig. 12. Loaddisplacement response obtained from simulations of eccentrically loaded Size 1,2 SETB specimens. For a particular specimen size, GL and GH are the least and
highest values of fracture toughness obtained from the multiple experimental loaddisplacement curves. GL corresponds to the curve exhibiting least toughness and GH
corresponds to the curve exhibiting the highest toughness. The P and D values have been normalized with xed reference values.

(a)

(b)

Fig. 13. Eccentric compact tension specimens used in the crack propagation experiments. (a) Eccentricity: 10 mm and (b) Eccentricity: 20 mm. Both specimens have a
nominal thickness of 6.35 mm.

displacement response and the crack path, the two most important
metrics from a structural viewpoint, on the mesh density.
Consider the problem of symmetrically loaded Single Edge
Notch Three-Point Bend (SETB) specimen as shown in Fig. 6. Shown
are the problem schematic, resulting crack paths for meshes whose
density varies over an order of magnitude, and the corresponding
global loaddisplacement response. Similarly, Fig. 7 shows the
problem of a standard Compact Tension Specimen (CTS). It should
be sufciently clear from these results that the traditional mesh
dependence is absent for the case of straight crack path. However,
at rst glance, the small variation in the loaddisplacement re-

sponse and crack path may suggest mesh sensitivity. This is


expected, as even in the absence of cracks, the resolution of the
high stress gradients does depend to a small degree on the element
dimension, and this naturally effects the crack evolution and
consequently the loaddisplacement response. Thus, these small
variations are not pathological, but an artifact of the numerical discretization. These two simulations demonstrate that the multiscale
method has no pathological mesh dependence, and that it objectively yields the global loaddisplacement response and the crack
path. The loaddisplacement responses (Figs. 6and 7C) are physically relevant, as they indicate that the strain energy release rate

3344

S. Rudraraju et al. / Composite Structures 94 (2012) 33363346

1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

0.7

0.8

0.9

0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1

0.1

0.2

0.3

0.4

0.5

0.6

Fig. 14. Loaddisplacement response obtained from simulations of eccentrically loaded CTS specimens. The P and D values have been normalized with xed reference values.

is mesh independent, because the area under the curve is equal to


the energy dissipated due to surface creation. However, these
problems do not involve crack turning, so the sensitivity of the
crack path to mesh discretization is not manifested here. For
curved crack propagation, mesh objectivity of the crack path has
been demonstrated in eccentrically loaded SETB specimen simulations reported in Rudraraju et al. [23].
5.2. Curved crack propagation in laminated ber composites
A 48-ply carbon ber polymer matrix composite with a quasiisotropic stacking sequence of [45/0/+45/90]6s, and which has a
ber volume fraction of 0.55 is used for all the subsequent experiments. The experimental setup and related observations are described in Rudraraju et al. [23], Rudraraju [21]. The lamina and
laminate properties are given in Table 1. The nominal thickness
of the panels studied is 6.35 mm. The cross section of the specimen
layup is shown in Fig. 8.
Curved crack paths, in general, involve mixed-mode conditions,
where the crack face is subjected to both normal and tangential
tractions. This mixed-mode evolution is driven by the traction
separation relations of the respective modes, and potentially a
mode-mixity condition. But as stated earlier, all simulations presented here consider only locally Mode-I conditions. The Mode-I
cohesive strength was obtained from double edge notched tension
(DENT) specimen experiments, and this value was xed for all
specimen sizes and geometries simulated in this section. The fracture toughness was obtained from standard compact tension specimen (CTS) experiments. The Mode-I fracture toughness in this
class of materials is both size and geometry dependent, so the frac-

ture toughness obtained from CTS experiments could be used directly in the eccentric CTS simulations, but for the symmetric
and eccentric SETB simulations, the fracture toughness was computed by normalizing the area under their respective experimental
loaddisplacement curves by the total crack area, as summarized
in Table 2. For more information on the DENT, CTS experiments
and the determination of Mode-I fracture toughness, readers are
referred to Rudraraju et al. [22]. In each of the simulations below,
the meshes contain about ten to twenty thousand elements. However, unlike the interface methods like standard cohesive zone
models, the crack is free to traverse an element along any direction,
thus removing any mesh based restriction on the crack path.
Consider the rst case-study of the symmetrically loaded SETB
specimens whose conguration is shown in Fig. 9. Five specimen
sizes with geometrically scaled planar geometry and xed thickness were considered. Multiple specimens of each size were tested
to signicantly capture the failure response envelope. The load
load point displacement responses of these specimens are shown
in Fig. 10. Table 2 summarizes the observed scaling in the loaddisplacement response, and hence in the value of fracture toughness,
calculated for each size by normalizing the area under their respective experimental loaddisplacement curves by the total crack
cross section area. P, D and GIc are xed reference values. Using
the fracture toughness values listed in Table 2, we obtain the
numerical loaddisplacement curves shown in Fig. 11. Across the
ve sizes, the numerical simulations faithfully reproduce the
experimental loaddisplacement response. Usually, the crack initiates before the peak load, and at the peak load the full bridging
zone will be formed. Further crack growth leads to a drop in the
load bearing ability of the panels due to the failure of the bers

S. Rudraraju et al. / Composite Structures 94 (2012) 33363346

3345

Fig. 15. Comparison of experimental and numerical crack paths for eccentrically loaded CTS specimens.

in the crack wake leading to movement of the active bridging zone.


The details of the effect of bridging zone formation and movement
on the load bearing ability of the specimens has been explained in
detail in Rudraraju et al. [22].
Similarly, for the more complex case of eccentric SETB specimens whose conguration is also shown in Fig. 9, the comparison
of the global loaddisplacement response and its comparison with
experimental results is given by Fig. 12. For this class of materials,
experimental loaddisplacement responses show sharp drops in
the post-peak load regime, which can be attributed to the heterogeneity in the cohesive response of the bridging zone in the crack
wake. Since numerically the material is modeled as a homogeneous medium (with the effective elastic and cohesive properties),
the numerical model has no spatial or angular distribution of these
properties. Thus, the numerical model only predicts a smooth load
displacement response in the post-peak regime. However, as seen
from the comparison, the mutiscale method captures the postpeak response fairly well, albeit without the sharp drops.
In the case of eccentric SETB specimens, eccentricity was due to
the asymmetry in the loading location. Now consider eccentrically
loaded CTS specimens, where the eccentricity is in the geometry as
shown in Fig. 13. The notches in the center of the CTS specimens
were moved by 10 mm for rst set of tests and 20 mm for the second test. Fig. 14 shows the comparison of the global loaddisplacement response. Again the global response is very similar. The
signicant difference in the slope of the experimental and numerical curves in the linear regime of the curve is due to some amount
of crushing under the loading rollers in the experiments. Further,
Fig. 15 shows a comparison of the experimental and numerical
crack paths, which match signicantly.

From a structural viewpoint, the crack path and the effective


load bearing ability of the cracked panel are of primary importance. As seen from the above simulation results, where the crack
path and loaddisplacement response have been signicantly captured by the simulations, it should be clear that the multiscale
methodology is a potential tool in simulating progressive damage
involving cohesive crack propagation.

6. Conclusions
In this paper, a multiscale approach to crack propagation in
ber reinforced panels, homogenized through the panel thickness,
was introduced which treated the discontinuous displacement
eld in crack propagation as a ne scale problem. The necessary
weak formulation, nite element framework, system of equations
and code implementation details were presented. The objectivity
of the resulting computational implementation was demonstrated
through some benchmark simulations and its applicability was
shown through the comparison of numerical and experimental
loaddisplacement responses and crack paths of laminated ber
reinforced composite specimens.
Acknowledgments
This work was supported by a NASA NRA Grant under the
ARMD IVHM Project. The interest and constant encouragement of
Dr. Steven M. Arnold and Dr. Brett Bednarcyk of the MACE Center
at NASA Glenn Research Center, and Craig Collier and Phil Yarrington of Collier Research Corporation is gratefully acknowledged.

3346

S. Rudraraju et al. / Composite Structures 94 (2012) 33363346

References
[1] Armero F, Garikipati K. An analysis of strong discontinuities in
multiplicative nite strain plasticity and their relation with the numerical
simulation of strain localization in solids. Int J Solids Struct 1996;
33(2022):286385.
[2] Armero F, Garikipati K. An analysis of strong discontinuities in multiplicative
nite strain plasticity and their relation with the numerical simulation of
strain localization in solids. Int J Solids Struct 1996;33:286385.
[3] Bazant ZP. Mechanics of distributed cracking. Appl Mech Rev
1986;39:675705.
[4] Borja RL, Regueiro RA. A nite element model for strain localization analysis of
strongly discontinuous elds based on standard galerkin approximation.
Comput Methods Appl Mech Eng 2000;190:152949.
[5] Camacho GT, Ortiz M. Computational modeling of impact damage in brittle
materials. Int J Solids Struct 1996;33:2899938.
[6] Criseld MA, Wills J. Solution strategies and softening materials. Comput
Methods Appl Mech Eng 1988;66(3):26789.
[7] Demmel JW, Eisenstat SC, Gilbert JR, Li XS, Liu JWH. A supernodal
approach to sparse partial pivoting. SIAM J Matrix Anal Appl 1999;20(3):
72055.
[8] Garikipati K. A variational multiscale method to embed micromechanical
surface laws in the macromechanical continuum formulation. Comput Model
Eng Sci 2002;3(2):17584.
[9] Garikipati K, Hughes TJR. A study of strain-localization in a multiple scale
framework. The one dimensional problem. Comput Methods Appl Mech Eng
1998;159:193222.
[10] Gasser TC, Holzapfel GA. Geometrically non-linear and consistently linearized
embedded strong discontinuity models for 3d problems with an application to
the dissection analysis of soft biological tissues. Comput Methods Appl Mech
Eng 2003;192:505998.
[11] Hughes TJR. Multiscale phenomena: Greens functions, the dirichlet-toneumann formulation, subgrid scale models, bubbles and the origins of
stabilized methods. Comput Methods Appl Mech Eng 1995;127(1
4):387401.
[12] Jirasek M. Comparative study on nite elements with embedded
discontinuities. Comput Methods Appl Mech Eng 2000;188:30730.
[13] Larsson R, Runesson K, Ottosen NS. Discontinuous displacement
approximation for capturing plastic localization. Int J Numer Methods Eng
1993;36:2087105.
[14] Moes N, Belytschko T. Extended nite element method for cohesive crack
growth. Eng Fract Mech 2002;69(7):81333.
[15] Moes N, Dolbow J, Belytschko T. A nite element method for crack growth
without remeshing. Int J Numer Methods Eng 1999;46:13150.

[16] Mosler J, Meschke G. Embedded crack vs. smeared crack models: a comparison
of elementwise discontinuous crack path approaches with emphasis on mesh
bias. Comput Methods Appl Mech Eng 2004;193:335175.
[17] Oliver J, Huespe AE. Continuum approach to material failure in strong
discontinuity settings. Comput Methods Appl Mech Eng 2004;193:3195220.
[18] Oliver J, Huespe AE, Sanchez PJ. A comparative study on nite elements for
capturing strong discontinuities: E-fem vs x-fem. Comput Methods Appl Mech
Eng 2006;195:473252.
[19] Pietruszczak ST, Mroz Z. Finite element analysis of deformation of strain
softening materials. Int J Numer Methods Eng 1981;17:32734.
[20] Ramakrishnan N, Okada H, Atluri SN. On shear band formation: II. Simulation
using nite element method. Int J Plas 1994;10(5):52134.
[21] Rudraraju SS. On the Theory and numerical simulation of cohesive crack
propagation with application to ber-reinforced composites. Ph.D. Thesis,
University of Michigan Ann Arbor; 2011.
[22] Rudraraju SS, Salvi A, Garikipati K, Waas AM. In-plane fracture of laminated
ber reinforced composites with varying fracture resistance: experimental
observations and numerical crack propagation simulations. Int J Solids Struct
2010;47(78):90111.
[23] Rudraraju SS, Salvi A, Garikipati K, Waas AM. Experimental observations and
numerical simulations of curved crack propagation in laminated ber
composites. J Compos Sci Technol 2012;72:106474.
[24] Schellekens JCJ, DeBorst R. On the numerical integration of interface elements.
Int J Numer Methods Eng 1993;36:4366.
[25] Simo JC, Oliver J. A new approach to the analysis and simulation of strain
softening in solids. Fract Damage Quasibrittle Struct 1994.
[26] Simo JC, Oliver J, Armero F. An analysis of strong discontinuities induced by
strain-softening in rate-independent inelastic solids. Comput Mech
1993;12:27796. http://dx.doi.org/10.1007/BF00372173.
[27] Song S, Waas AM. A nonlinear elastic foundation model for interlaminar
fracture of laminated composites. Compos Eng 1993;3(10):94559.
[28] Temam R, Strang G. Functions of bounded deformation. Arch Ration Mech Anal
1980;75:721. http://dx.doi.org/10.1007/BF00284617.
[29] Ungsuwarungsri T, Knauss WG. The role of damage-softened material behavior
in the fracture of composites and adhesives. Int J Fract 1987;35:22141.
[30] Wells GN, Sluys LJ. A new method for modelling cohesive cracks using nite
elements. Int J Numer Methods Eng 2001;50:266782.
[31] Xie D, Salvi A, Sun C, Waas AM, Caliskan A. Discrete cohesive zone model to
simulate static fracture in 2-d triaxially braided carbon ber composites. J
Compos Mater 2006;40:122.
[32] Xie D, Waas AM. Discrete cohesive zone model for mixed-mode fracture using
nite element analysis. Eng Fract Mech 2006;73:178396.
[33] Xu XP, Needleman A. Numerical simulation of fast crack growth in brittle
solids. J Mech Phys Solids 1994;42:1397434.

Anda mungkin juga menyukai