Vous êtes sur la page 1sur 19

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

net/publication/338126752

A phase-field model for fracture of unidirectional fiber-reinforced polymer


matrix composites

Preprint · December 2019

CITATIONS READS

0 234

4 authors:

Funda Aksu Denli Osman Gueltekin


Middle East Technical University Middle East Technical University
4 PUBLICATIONS   0 CITATIONS    11 PUBLICATIONS   93 CITATIONS   

SEE PROFILE SEE PROFILE

Hüsnü Dal Gerhard Holzapfel


Middle East Technical University Graz University of Technology
60 PUBLICATIONS   443 CITATIONS    362 PUBLICATIONS   19,036 CITATIONS   

SEE PROFILE SEE PROFILE

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

Multiphysics Phenomena in Li-Ion Batteries View project

Computational Cardiology View project

All content following this page was uploaded by Hüsnü Dal on 23 December 2019.

The user has requested enhancement of the downloaded file.


Noname manuscript No.
(will be inserted by the editor)

A phase–field model for fracture of unidirectional fiber–reinforced


polymer matrix composites
Funda Aksu Denli · Osman Gültekin · Gerhard A. Holzapfel · Hüsnü Dal

Received: date / Accepted: date

Abstract This study presents a crack phase–field approach 1 Introduction


for anisotropic continua to model, in particular, fracture of
fiber–reinforced matrix composites. Starting with the vari- Since the introduction of composite materials in the 1960’s
ational formulation of the multi–field problem of fracture such as glass–polyester and carbon–epoxy, great success has
in terms of the deformation and the crack phase fields, the been achieved in estimating the effective micromechanical
governing equations feature the evolution of the anisotropic properties of composites, the respective homogenized re-
crack phase–field and the balance of linear momentum, pre- sponse and the plate theories for laminates. However, the-
sented for finite and small strains. A recently proposed ories pertaining to the fracture of composite materials are
energy–based anisotropic failure criterion is incorporated not on par with the afore–mentioned theories in terms of
into the model with a constitutive threshold function reg- their applicability and accuracy. Although a lot of efforts
ulating the crack initiation in regard to the matrix and the have been made over five decades, the prediction of com-
fibers in a superposed framework. Representative numerical posite failure still remains largely unsolved with plenty of
examples are shown for the crack initiation and propaga- uncertainties, as reviewed by Talreja [47].
tion in unidirectional fiber–reinforced polymer composites Regarding fiber–reinforced polymers (FRPs), the failure
under Mode–I, Mode–II and mixed–mode bending. Model mechanisms are, in general, related to (i) the mechanical be-
parameters are obtained by fitting to sets of experimental havior of the individual lamina and laminate as a whole and
data. The associated finite element results are able to cap- (ii) the direction of the loading. Compared with steel and
ture anisotropic crack initiation and growth in unidirectional other more conventional materials, the failure mechanism
fiber-reinforced composite laminates. of FRP composites is much more complex and its predic-
tion presents a tremendous challenge for engineers and re-
searchers as they possess an inherent heterogeneity with dis-
Keywords fibrous biological tissues · anisotropy · hyper-
tinct interfaces in their structure. The model approaches for
elasticity · quasi-incompressibility · quasi-inextensibility ·
the crack initiation and progression in composite materials
mixed finite element formulation
can be divided into two categories: one is based on strength
criteria (i.e., failure at a material point) and the other one is
F. A. Denli · O. Gültekin · H. Dal based on energy criteria (i.e., surface formation), see, e.g.,
Department of Mechanical Engineering, Middle East Technical Uni- Talreja & Singh [46].
versity, Dumlupinar Bulvari 1, 06800, Çankaya,Ankara, Turkey. According to strength–based criteria, micro-cracks form
G. A. Holzapfel when the local stress (strain) state in a ply reaches a critical
Institute of Biomechanics, Graz University of Technology, Stremayr- level. To date, several strength–based failure criteria that are
gasse 16/II, 8010, Graz, Austria
Department of Structural Engineering, Norwegian University of Sci- rooted in metal fracture have been proposed for composite
ence and Technology (NTNU), 7491 Trondheim, Norway materials namely, Tsai–Hill [6,22], Tsai–Wu [48], Hashin
Corresponding author: H. Dal
[20] and Cuntze [13], just to mention a few. Expressed by
Tel.: +90(312) 210 2584 quadratic polynomials, they involve strength values as ma-
Fax: +90(312) 210 2536 terial constants that need to be determined from experimen-
E-mail: dal@metu.edu.tr tal data. In principle, failure manifests when the elastic re-
2 Funda Aksu Denli et al.

sponse in any combination of the stress components that ex- data and the strength–based estimations, see Talreja & Singh
ceed a threshold given by the respective criterion. Unlike [46].
the early maximum stress and strain criteria, the theories by Energy–based criteria originate from linear elastic frac-
Tsai–Hill, Tsai–Wu and Hashin take into account the pos- ture mechanics where the crack starts to grow when the en-
sible stress interactions at failure. However, their use also ergy release rate G reaches a critical value Gc expressed by
leads to other issues. The Tsai–Hill criterion is based on the the equilibrium G = Gc , as introduced by Griffith [16]. In a
Hill criterion, which is basically an anisotropic extension of multiple cracking of a composite laminate, the progression
the von–Mises [32] yield theory developed for isotropic ma- of a crack located in between the plies is arrested at the in-
terials such as metals. The predicted yield stress therein is terface and any further input of energy to the laminate leads
the same in tension and compression, an appropriate conclu- to the formation of more ply cracks occurring elsewhere. In
sion for orthotropic metal sheets; however, such an assump- such a case, the conventional fracture mechanics approach
tion is far–fetched in regard to unidirectional composites as requires modifications, e.g., the concept of finite fracture
the mechanisms characterizing the failure under tension and mechanics and variational stress analysis, which was among
compression are quite different from each other. In fact, a others presented by Hashin [21] and Nairn [33]. Aside from
cluster of failed fibers in a cross–section is involved under that, a strain energy–based failure criterion was suggested
tension and accompanied by some splits of fibers linking the by Wolfe & Butalia [51] for a wide variety of unidirectional
neighboring cross–sections, whereas a local kink band facil- and symmetric laminates under biaxial loading.
itated by the micro–buckling deformation modes results in
The cohesive zone modeling (CZM) appears to be one
the failure of the composite under compressive loads, see,
of the most prevalent approaches used to mimic the me-
e.g., Argon [5]. Tsai-Wu account for unequal strengths in
chanical failure of laminae which was presented by many
tension and compression that evokes the Bauschinger effect.
including Turon et al. [49], Yang & Cox [52], Naghipour et
This criterion assumes a scalar–valued function of stress
al. [34] and Zhao et al. [54] for uni– and multi–directional
components in regard to the failure surface characterized by
composite laminates in Mode–I and Mode–II fracture. Like-
an ellipsoid. The problem associated with this criterion is
wise, there have been several successful applications of the
how to determine the inclination of the ellipsoid as the biax-
extended finite element method (XFEM) by, e.g., Grogan
ial components of the strength tensor do not have a unique
et al. [17], Wang et al. [50] and Yazdani et al. [53] on the
value, but hinge on the stress state. To overcome this diffi-
delamination of composite materials. Recent revelations by
culty, Hashin suggested many piece-wise smooth surfaces,
Dal et al. [14], Reinoso et al. [40], Alessi & Freddi [2] and
each representing a distinct failure mode. In fact, the fiber
Arterio et al. [4] highlight an alternative approach, namely
failure in the criterion is decoupled from the matrix fail-
the crack phase–field modeling to predict damage and fail-
ure. Yet the problem of ascertaining the strength constants
ure of composite laminates. In contrast to CZM and XFEM,
for compressive modes render the theory rather impractical.
the crack phase–field approach utterly ignores the realiza-
Later, Puck & Schürmann [38] proposed a more justifiable
tion of discontinuities as the 2D crack surface smears out
failure theory in the sense of model constants.
in a volume domain in 3D, which is determined by a spe-
cific field equation alongside the balance of linear momen-
All of the afore–mentioned criteria regard the failure tum describing the elasticity of the solid. The well-known
as a single event and consider composites as homogeneous limitations of the classical fracture mechanics, e.g., curvilin-
solids. Besides, the failure plane is not explicitly influenced ear crack paths, crack kinking, branching angles, and mul-
by the existence of fibers, i.e. the crack does not cut across tiple cracking are surmounted through a variational princi-
the fibers. Nor do the distribution of fibers (uniform or pal of the minimum energy, see Francfort & Marigo [15].
nonuniform) and the nature of the matrix–fiber bond alter The thermodynamically consistent and algorithmically ro-
the critical tractions and the orientation of the crack surface. bust formulations of phase–field models were introduced in
Furthermore, they are created on the basis of the traditional the seminal works by Miehe et al. [27,28]. The approach
strength of the materials approach from the structural design is modular and can be applied to non–standard solids ex-
aspect, and can only impart knowledge about the critical de- hibiting complex cracking mechanisms under multiphysics
sign points where the failure may occur. They fall short of phenomena, see, e.g., Miehe et al. [29–31]. Ductile failure of
describing differences in the crack initiation (as a material elastoplastic materials is treated in Ambati et al. [3] and Bor-
point failure process) and the crack propagation (as a sur- den et al. [8]. Anisotropic crack phase-field evolution has
face growth process). As a consequence, the effect of ply recently been considered by, e.g., Li et al. [25] and Teicht-
thickness on the transverse cracking cannot be properly ac- meister et al. [45], which is based on the extended Cahn-
counted by such criteria. Other problems are the impossibil- Hilliard model, see Cahn and Hilliard [10], to account for
ity of analytical characterization of local stress states except the anisotropic surface energy emanating from the preferred
for a few cases and the conflict between the experimental directions in materials. Apart from that Clayton and Knap
A phase–field model for fracture of unidirectional fiber–reinforced polymer matrix composites 3

f0

ϕ(X, t)
f

∂Bt ∂St
X x
F

B S
∂Bϕ ∂Sϕ
t0 t

Fig. 1: Nonlinear deformation of a solid. The reference configuration B ∈ R3 and the spatial configuration S ∈ R3 ; ϕ : B × R 7→ R3 is the
nonlinear deformation map which maps the material point position X ∈ B onto the spatial position x = ϕ(X, t) ∈ S, at time t ∈ Rt . The
deformation gradient F maps a Lagrangian line element dX onto its Eulerian counterpart dx = F dX. The anisotropic micro-structure of the
material point X is rendered by unidirectional fibers with the unit vector f 0 . Likewise, the anisotropic micro-structure of the spatial point x is
described by f , as the spatial counterpart of f 0 .

[11] and Nguyen et al. [37] proposed anisotropic phase-field 2 Basics of the Multi-field Problem of Fracture
models for polycrystals. The approach of Schreiber et al.
[41] uses an anisoptropic geometric resistance to failure in This section lays bare the primary field variables, namely
the sense of Gültekin [18]. the crack phase–field d and the deformation map ϕ govern-
ing the diffusive crack evolution and the balance of linear
For materials such as soft biological tissues and compos- momentum in a coupled manner. The framework is provided
ite laminates, the anisotropic fracture is not only a geometri- for the finite and small–strain settings which cover both the
cal phenomenon but also a mechanical event arisen from the mechanical and phase–field problems.
fibrous content embedded in an otherwise isotropic matrix
material, necessitating the use of an anisotropic crack driv-
ing force apart from directional geometric resistance. Hence, 2.1 The primary field variables
the current study follows the footsteps of the previous con-
Let us consider a continuum body at time t0 ∈ T ⊂ R,
tributions by Gültekin et al. [18,19] in which the anisotropic
which we refer to as the reference configuration, as desig-
crack phase–field at finite strains was introduced. Incorpo-
nated by B ⊂ R3 , with the material point X ∈ B. Similarly,
rated was also a novel energy–based failure criterion based
the deformed body at current time t ∈ T ⊂ R, which we
on the distinction of fibrous and matrix contributions to the
refer to as the spatial configuration, is denoted by S ⊂ R3
elastic mechanical response. The current study, however, ex-
with the spatial point x ∈ S mapped via the deformation
amines the fracture of the FRPs composed of a polymer ma-
field ϕ, see Fig. 1. Thus,
trix reinforced with fibers. (
B × T → S,
The article is organized as follows. Section 2 outlines the ϕt (X) : (1)
(X, t) 7→ x = ϕ(X, t).
primary field variables with the corresponding finite strain
kinematics and the diffusive features of the anisotropic crack Along with the deformation field given in (1), the basic ge-
phase–field. Section 3 is concerned with the variational for- ometric mapping for the crack phase–field d is expressed by
mulation of the multi–field problem of fracture resulting (
B × T → [0, 1],
in the coupled balance equations. Section 4 focuses on the d: (2)
anisotropic hyperelastic constitutive response reflecting the (X, t) 7→ d(X, t),
elastic mechanical behavior of the composite in the Eulerian which interpolates between the intact (d = 0) and the rup-
framework. A brief account of the energy–based anisotropic tured (d = 1) state of the material.
failure criterion is also provided. In section 5, the representa-
tive numerical examples exhibit the capabilities of the model
with regard to a standard problem of a transversely isotropic 2.2 Kinematics of the mechanical problem
single edge–notched specimen under Mode–I and Mode–II
loading scenarios and a realistic test case for a unidirectional We start with the description of the deformation gradient,
laminate withstanding mixed–mode bending. Finally, Sec- i.e.
tion 6 summarizes the article. F = ∇ϕt (X), (3)
4 Funda Aksu Denli et al.

P ·N = T L∇d · N = 0
∂Bt

N l N
X∈B X∈B

ϕ d
Γl (d)
ϕ = ϕ̄ ∂Bϕ ∂Bd

(a) Deformation field (b) Crack phase–field

Fig. 2: Multi-field problem: (a) mechanical problem of deformation along with Dirichlet and Neumann-type boundary conditions, that is ϕ = ϕ
and P · N = T , respectively; (b) evolution of the crack phase–field problem with the Neumann-type boundary condition L∇d · N = 0.

which maps the unit Lagrangian line element dX onto its 2.3 Kinematics of the phase–field problem
Eulerian counterpart dx = F dX. The gradient operators
∇(•) and ∇x (•) denote the gradient operator with respect to The deformable domain for the concerning problem is as-
the reference X and the spatial x coordinates, respectively. sociated with he deformation field as given in Fig. 2(a). For
The determinant of F , i.e. the Jacobian J := detF > 0, a non-deformable domain, the gradient operator can simply
characterizes the map of an infinitesimal reference volume be taken as ∇x (•) = ∇X (•) = ∇(•). A sharp crack sur-
2
element onto the associated spatial volume element. The R ⊂ R in
face topology at a frozen time t is defined by Γ (d)
right and left Cauchy–Green tensors read the solid B through a surface integral Γ (d) = Γ dA. The
hallmark of the crack phase–field approach is that it circum-
C = F T gF , b = F G−1 F T , (4) vents the cumbersome task of tracking such discontinuities
which measure the deformation in the Lagrangian and Eule- and it approximates the surface integral by a volume inte-
rian configurations, respectively. Furthermore, we adopt the gral, thereby creating a regularized crack surface Γl (d), see
formalism in the sense of Marsden & Hughes [26] and equip also Fig. 2(b) such that
the two manifolds, namely B and S with the covariant ref- Z
erence and spatial metric tensors G = δIJ E I ⊗ E J and Γl (d) = γ(d, ∇d)dV where
B
g = δij ei ⊗ ej , respectively, where δIJ and δij are simply (9)
1 2 2
evaluated as the Kronecker deltas in the Cartesian coordi- γ(d, ∇d) = (d + l ∇d · ∇d) ,
2l
nate system. The energy stored in a hyperelastic isotropic
material is characterized by the three invariants designates the isotropic crack surface density function,
which satisfies the condition γ(d, Q∇d) = γ(d, ∇d), ∀Q ∈
1 2 
I1 = trb, I2 = I1 − tr(b2 ) , I3 = det b. (5) O(3). The tensor variable Q denotes the rotations in the or-
2 thogonal group O(3), containing rotations and reflections.
The anisotropic response of a unidirectional FRP compos- The length–scale parameter l controls the breadth of the
ite requires the description of an additional invariant. To this crack. This approximation can be extended to a class of
end, we introduce a reference unit vector f 0 for the fiber ori- anisotropic materials as
entation in the reference configuration and its spatial coun- Z
terpart Γl (d) = γ(d, ∇d; L)dV, where
B
f = F f 0, (6) (10)
1
γ(d, ∇d; L) = (d2 + ∇d · L∇d) ,
which idealizes the micro-structure of the unidirectional 2l
FRP composite. We can express the related Lagrangian and is the anisotropic crack surface density function with the
Eulerian forms of the structure tensors A and Af as follows condition γ(d, Q∇d) = γ(d, ∇d), ∀Q ∈ G ⊂ O(3), where
G designates a symmetry group as a subset of O(3). The
A = f0 ⊗ f0 and Af = f ⊗ f . (7)
second-order anisotropic structure tensor L is given as
The physically meaningful additional fourth–invariant
L = l2 (I + ωf0 f 0 ⊗ f 0 ), (11)
I4 = f · gf , (8)
which aligns the crack with the orientation of fibers in the
measures the square of the stretch along the mean fiber di- continuum, see Fig. 3. Therein, the anisotropy parameter ωf0
rection in the unidirectional FRP composite. regulates the transition from weak to strong anisotropy. For
A phase–field model for fracture of unidirectional fiber–reinforced polymer matrix composites 5

0
e1

f
(a) L = l2 I (b) L = l2 (I + e1 ⊗ e1 ) (c) L = l2 (I + f 0 ⊗ f 0 )

0 d 1

Fig. 3: Damage field on a square block: (a) isotropic damage field, (b) anisotropic damage field with fiber orientation θ = 0◦ stated by the unit
base vector e1 , (c) anisotropic damage field with fiber orientation θ = 45◦ given by the unit vector f 0 .

isotropic solids ωf0 = 0, whereas for a general anisotropic of the block leads to damage fields depicted in Fig. 3 for
continuum, it must lie in an open range, i.e. −1 < ωf0 < various structure tensors L associated with different fiber
∞ in order to satisfy the ellipticity condition for Γl (d), see orientations. In order to demonstrate the influence of wf0 ,
Gültekin [19] for an elaborate discussion. four representative cracks for two different anisotropy fac-
tors, namely wf0 = 0 in Fig. 4(a),(c) and wf0 = 10 in
Fig. 4(b),(d) are depicted. The introduced centroidal cracks
2.4 Euler–Lagrange equations of the phase–field problem (dˆ = 1) are vertical in the first and horizontal in the sec-
ond row. When the anisotropy parameter equals to zero, i.e
From a purely geometrical perspective, the boundary of the wf0 = 0, the crack smears isotropically so that the geomet-
domain under interest can be decomposed into Dirichlet and ric resistance to crack propagation is identical in all direc-
Neumann-type boundaries such that ∂B = ∂B d ∪ ∂B q and tions. However, wf0 = 10 smears the crack considerably
∂B d ∩ ∂B q = ∅. By considering (10), we can state the min- more in the transverse plane towards the fiber direction e1 ,
imization principle as see Fig. 4(b), whereas the minimum smearing occurs around
n o the cracks parallel to the fiber direction, see Fig. 4(d), as ob-
d(X) = Arg d∈W inf Γl (d) , (12) served in the second column. This means that cracks prop-
agating across the fibers are penalized and the crack prop-
along with the Dirichlet–type boundary constraint
agating along the fibers are favored. This is due to the fact
W = {d | d(X) ∈ B ∧ d = dˆ on ∂B d} . (13) that the energy threshold for the cracks propagating across
the fibers are higher for wf0 > 0. The converse applies for
Whilst an already existing crack is given by dˆ = 1, the in- −1 < wf0 < 0.
tact state is described by dˆ = 0. Although the boundary
value problem admits any meta-states dˆ ∈ [0, 1] on ∂B d,
we confine ourselves for the two ideal states. The Euler-
Lagrange equations are obtained after employing the min-
imization principle. Thus,
1
[d − Div(L∇d)] = 0 in B,
l (14)
L∇d · N = 0 on ∂B q , 3 Governing Equations of the Anisotropic Fracture
where the divergence term interpolates d between the intact
and the ruptured state of the material. In (14)2, N denotes This section deals with the coupled equations of the elastic–
the unit surface normal oriented outward in the reference fracture problem for finite and small strains, where the clas-
configuration. Solving only the phase–field problem in the sical balance of linear momentum is accompanied by the
purely geometrical context, a 2–D square block with dˆ = 1 evolution equation of the crack phase–field; the strong forms
at the single centroidal point and L∇d = 0 on the sides of the boundary–value problem are presented.
6 Funda Aksu Denli et al.

e1 e1

(a) wf 0 = 0 (b) wf0 = 10

e1 e1

(c) wf 0 = 0 (d) wf0 = 10

0 d 1

Fig. 4: Damage field on a square block with a single fiber family whose orientation is characterized by the unit base vector e1 : (a),(b) for a vertical
centroidal crack; (c),(d) for a centroidal horizontal crack where wf0 = 0 in the first column and wf0 = 10 in the second column.

3.1 Rate–dependent variational formulation based on (17)


power balance
The free–energy function Ψ defined in (17) characterizes a
3.1.1 Finite-strain setting degrading continuum with

As a point of departure, we introduce the viscous rate–type Ψ (g, F , Af ; d) := g(d)Ψ0 (g, F , Af ), (18)
potential Πη as
Πη = E + Dη − P. (15) where Ψ0 is the effective free–energy function of the hypo-
thetically intact solid. In (18), a monotonically decreasing
The first term E on the right–hand side of (15) represents the quadratic degradation function, i.e.
rate of energy storage functional, i.e.
Z
˙ = (P : Ḟ − f ḋ)dV, g(d) := (1 − d)2 , (19)
E(ϕ̇; d) (16)
B
where the work conjugate variables to ϕ and d are the first describes the degradation of the solid with the evolving
Piola–Kirchhoff stress tensor P and the scalar energetic crack phase–field parameter d together with the following
force f , respectively, i.e. growth conditions:

P = ∂F Ψ (g, F , Af ; d), f = −∂d Ψ (g, F , Af ; d). g ′ (d) ≤ 0 with g(0) = 1, g(1) = 0, g ′ (1) = 0. (20)
A phase–field model for fracture of unidirectional fiber–reinforced polymer matrix composites 7

The first condition ensures degradation, while the second 3.1.2 Small-strain setting
and third condition set the limits for the intact and the rup-
tured state, and the final condition ensures the saturation at The displacement field u = x−X is described at a material
d → 1. The second term Dη on the right–hand side of (15) is point X ∈ B ⊂ R3 and at time t ∈ T , i.e.
a viscous regularized dissipation functional due to fracture, (
B × T → R3 ,
i.e. u(X, t) : (29)
Z (X, t) 7→ u(X, t).
˙ β; d) = [β d˙ − 1 hχ(β; d, ∇d)i2 ]dV,
Dη (d, (21)
B 2η The rate of energy storage functional E on the right–hand
where the artificial viscosity η ≥ 0 regulates the scalar vis- side of (15) in the small-strain setting reads
cous over–stress χ, which reads Z
E(u̇; ḋ) = ( σ : ε̇ − f ḋ )dV, (30)
χ(β; d, ∇d) = β − gc [δd γ(d, ∇d; L)]. (22) B

where the stress tensor σ is the work conjugate variable of


The Macaulay brackets in (21) filter out the positive values,
the small-strain measure ε = sym ∇u. In (30), the energetic
χ > 0, while gc in (22) stands for the critical fracture en-
force f appears as the work conjugate of the damage vari-
ergy. Finally, the last term P on the right–hand side of (15)
able d. The stress tensor σ and the scalar energetic force f
denotes the (classical) external power functional acting on
are then expressed as
the body according to
Z Z σ = ∂ε Ψ (ε, A; d), f = −∂d Ψ (ε, A; d). (31)
P(ϕ̇) = ρ0 γ · ϕ̇dV + T · ϕ̇dA, (23)
B ∂Bt The term f can also be interpreted as the local crack driving
where ρ0 , γ and T represent the material density, the pre- force. The free–energy function Ψ in (31) characterizes a
scribed body force and the surface traction, respectively. degrading continuum with
Now, with the rate–type potential Πη at hand, we propose a Ψ (ε, A; d) := g(d)Ψ0 (ε, A), (32)
mixed variational principle of the evolution problem as fol-
lows where Ψ0 is the effective free–energy function of the hypo-
  thetically intact solid. The regularized dissipation functional
˙ β} = Arg
{ϕ̇, d, inf inf sup Πη , (24) (21) and the scalar viscous over–stress function (22) remain
˙
ϕ̇∈Wϕ̇ d∈W β≥0

unchanged. Finally, the (classical) external power functional
with the admissible domains for the primary variables P stated on the right–hand side of (15) can now be written
in the form
Wϕ̇ = {ϕ̇ | ϕ̇ = 0 on ∂Bϕ }, Z Z
(25) P(u̇) = ρ0 γ · u̇dV + T · u̇dA, (33)
Wd˙ = {d˙ | d˙ = 0 on ∂Bd }. B ∂Bt

Afterwards, the variation of the potential Πη with respect to where ρ0 , γ and T represent the material density, the pre-
˙ β} along with simple algebraic manipula-
the fields {ϕ̇, d, scribed body force and the surface traction, respectively.
tions via elimination and substitution of the respective terms With the expressions for the rate–type potential Πη at hand,
(see Gültekin et al. [19] for more details) results in the strong we propose a mixed variational principle of the evolution
form of the field equations, i.e. problem as follows
 
1: Div P + ρ0 γ = 0 , inf inf sup Πη ,
{u̇, ḋ, β} = Arg u̇∈W (34)
˙
u̇ d∈W β≥0
(26) ḋ

2: η d˙ = 2(1 − d)H − d + Div(L∇d).


with the admissible domains for the primary variables
The first equation in (26) simply describes the balance of lin- Wu̇ = {u̇ | u̇ = 0 on ∂Bu },
ear momentum, whereas the latter states the evolution equa- (35)
tion for the crack phase–field in which H indicates the crack Wḋ = {ḋ | d˙ = 0 on ∂Bd }.
driving source term such that
Afterwards, the variation of the potential Πη with respect to
Ψ0 ˙ β} and substitution of the respective terms
H= . (27) the fields {u̇, d,
gc /l (see Miehe et al. [27] for more details) we obtain the strong
The evolution of the phase–field parameter can be recast into form of the field equations, i.e.
the form
1: Div σ + ρ0 γ = 0 ,
1 (36)
d˙ = [2(1 − d)H̄ − d + ∇d · L∇d], (28)
η 2: η d˙ = 2(1 − d)H − d + Div(L∇d).
8 Funda Aksu Denli et al.

3.2 A note on the weak formulation and numerical To characterize the local anisotropic mechanical re-
implementation sponse of a unidirectional FRP composite, the free–energy
function can be stated as
On the numerical side, a canonical Galerkin-type finite el-
Ψ0 (g, F , Af ) := Ψ0iso (J, I1 ) + Ψ0ani (I4 ). (37)
ement procedure renders the weak forms of the coupled
balance equations given in (26). The nonlinearities due to For the isotropic part of the mechanical response of the com-
the geometry and the constitutive law, as subsequently de- pressible polymer matrix we adopt the generic compressible
scribed, necessitates a linearization process employed on the neo-Hookean free–energy function. Thus,
weak forms. Afterwards, an identical temporal and spatial λ µ
discretization scheme is employed for the deformation map Ψ0iso (J, I1 ) := (lnJ)2 + (I1 − 2lnJ − 3). (38)
2 2
and the crack phase–field. The field variables are appropri-
For the anisotropic part we use the standard reinforcing
ately discretized with isoparametric shape functions so as
model in the sense of Qiu & Pence [39], i.e.
to transform the continuous integral equations of the non-
µf
linear weighted-residuals and their linearizations to a set of Ψ0ani (I4 ) := (I4 − 1)2 . (39)
coupled, discrete algebraic equations. Finally, this set of al- 4
gebraic equations is solved by a one–pass operator–splitting In (38), λ denotes the Lamé’s first constant, whereas µ de-
algorithm in a Newton-type iterative solver that successively notes the Lamé’s second constant or the shear modulus. In
updates the history field described by the failure criterion, the anisotropic term (39), µf stands for a stress-like material
the crack phase–field and the deformation field. For a more parameter associated solely with the fibrous content.
elaborate numerical treatment of the respective problem, the Let us now exploit the Coleman–Noll procedure on the
readers are referred to, e.g., Gültekin et al. [18,19]. Clausius–Planck inequality, and use the form of the free–
energy function Ψ , as introduced in (18), so that we can re-
trieve the Kirchhoff stress tensor τ as
4 Constitutive Equations of the Multi–Field Problem τ := P F T = 2∂g Ψ = g(d)τ 0 , τ 0 = 2∂g Ψ0 . (40)
In this section, we elucidate: (i) the constitutive equations Therein, g(d) is a monotonically decreasing quadratic
that capture the nonlinear anisotropic response of a unidi- degradation function as provided in (19). For the relevant
rectional FRP composite and (ii) the related energy–based nonlinear continuum mechanics used see, e.g., Holzapfel
anisotropic failure criterion capturing the state of the mate- [23]. By substituting (37) along with (38) and (39) into the
rial at which the cracking starts/propagates. definition (40)2 we get the stress expression for the intact
material, i.e.
τ 0 = λlnJg −1 + µ(b − g −1 ) + 2ψ4 f ⊗ f , (41)
4.1 The constitutive model for the unidirectional FRP
composites where we have introduced the (deformation-dependent) con-
stitutive function ψ4 by
The free–energy function of isotropic solids can be modeled µf
through the three invariants I1 , I2 , I3 , which constitute the ψ4 := ∂I4 Ψ0 = (I4 − 1). (42)
2
integrity basis of the deformation tensors C or b, see e.g., The change in the Kirchhoff stress tensor is provided by the
Spencer [44]. For incompressible materials the two invari- elasticity tensor here given in the spatial form as
ants I1 and I2 are enough to describe the isotropic deforma-
2 2
tion. C := 4∂gg Ψ = g(d)C0 , C0 = 4∂gg Ψ0 , (43)
For transversely anisotropic solids, one can introduce for which the effective elasticity tensor C0 has the explicit
the additional set of invariants I4 and I5 with the help of expression
the structural tensors that satisfy the objectivity requirement
under superimposed rigid body rotations, see e.g., Betten C0 = λg −1 ⊗ g −1 + 2(µ − λlnJ)Ig−1 + 4ψ44 M, (44)
[7], Boehler [9] and Schröder & Neff [42]. In unidirectional where the symmetric fourth-order identity tensor Ig−1 has
fiber–reinforced materials, the stored energy can be obtained the index representation (Ig−1 )ijkl = (δ ik δ jl + δ il δ jk )/2.
in terms of a free energy of the unreinforced base matrix In addition, the constitutive function ψ44 reads
with the arguments I1 and I3 augmented by a storage func- µf
tion that involves the fourth invariant I4 related to the fiber ψ44 := ∂I4 ψ4 = , (45)
2
stretch. The latter function is also known as the standard re-
and the fourth-order structure tensor takes on the following
inforcing model, see Qiu & Pence [39]. A similar approach
form
can be adopted for the modeling of soft biological tissues,
see, e.g., Holzapfel et al. [23]. M := f ⊗ f ⊗ f ⊗ f . (46)
A phase–field model for fracture of unidirectional fiber–reinforced polymer matrix composites 9

4.2 Linearization of the constitutive model: small-strain For the rate–independent case for which η → 0, the expres-
setting sions (51) and (52) engender distinct evolution equations of
the crack phase–field in relation to the ground–matrix and
The linearized form of the free–energy function (37) can be the fibrous content, i.e.
represented as iso
2(1 − d)H = d − Div(Liso ∇d),
Ψ0 (ε, A) := Ψ0iso (ε) + Ψ0ani (ε, A). (47) (53)
ani ani
2(1 − d)H = d − Div(L ∇d).
The isotropic and the anisotropic parts take on the simple
quadratic forms What remains is to superpose the two distinct failure pro-
cesses (53), which leads to the rate–independent evolution
λ equation of the phase–field, i.e.
Ψ0iso (ε) := (tr ε)2 + µ(ε : ε),
2 (48)
1
Ψ0ani (I4 ) := µf (ε : A)2 . (1 − d)H = d − Div(L∇d), (54)
2
The linear stress tensor σ o := ∂ε Ψ0 of the intact solid can along with the specific form of the dimensionless crack driv-
then be derived from (48), i.e. ing source term
σ 0 = λ(tr ε)1 + 2µε + 2µf (ε : A)A. (49)  
H(t) = maxs∈[0,t] hH(s) − 1i ,
The related elasticity moduli C0 = ∂ε σ 0 of the intact solid (55)
iso ani
can be derived as H=H +H .

C0 = λ1 ⊗ 1 + 2µI + 2µf A ⊗ A. (50) Relation (55) indicates an irreversible and positive crack
driving source term such that the maximum positive value
Therein, the fourth order symmetric identity tensor I has the of H(s) − 1 is tracked down for the entire deformation his-
following index representation (Iijkl = (δ ik δ jl + δ il δ jk )/2. tory s ∈ [0, t]. The Macaulay brackets filter out the positive
The proposed Ansatz is the simplest form of transverse values for H(s) − 1 and keeps the solid intact until the fail-
isotropy with only one additional material parameter µf to ure surface is reached, which denotes the energetic criterion
describe the axial reinforcement due to unidirectional fibers. proposed by Gültekin et al. [18,19]. Finally, in view of (54),
This particular choice assumes identical shear response in we specify the rate–dependent case, i.e.
the planes including the fibers and the transverse plane. It
1
also excludes the coupling effect between the bulk response η d˙ = (1 − d)H − [d − Div(L∇d)], (56)
and the fiber reinforcement. |{z} | {z } | 2 {z }
Crack evolution Driving force
Geometric resistance

where the evolution of the crack is characterized by the bal-


4.3 Energy–based anisotropic failure criterion
ance between the crack driving force and the geometric re-
sistance to the crack, see Miehe et al. [29]. A closer exam-
Following Gültekin et al. [18,19], we start with the assump-
ination of (56) shows that the geometric resistance is direc-
tion that two distinct failure processes govern the cracking
tional dependent. In Fig. 4, for instance, the energy threshold
of the ground matrix and the fibers, whereby the anisotropic
ratio of the crack in the direction f 0 to that in the transverse
structure tensor L in (11) is additively decomposed as
direction to f 0 is Gk /G⊥ = L11 /L22 = (wf0 + 1)/1 lead-
L = Liso + Lani with ing to an isotropic crack resistance for wf0 = 0.
(51)
Liso = l2 I, and Lani = l2 ωf0 f 0 ⊗ f 0 .
5 Representative Numerical Examples
Next we introduce gciso
and gcani
corresponding to the critical
fracture energies attributed to the ground–matrix (isotropic) In this section we demonstrate the utility of the proposed dif-
fusive fracture model for FRP composites. The model is ca-
and the fibrous content (anisotropic) of FRP, respectively,
which homogenize the distinct mechanical resistance of pable of capturing anisotropic fracture, which is illustrated
the respective interactions against rupture. Hence, the crack for a spectrum of benchmark problems such as single edge–
driving source term (27) can be decomposed as notched specimens with various fiber orientations subjected
to Mode–I and Mode–II loadings (Section 5.1). Although
iso Ψ0iso ani Ψ0ani the strain levels remain small until the onset of cracking,
H = , H = . (52) the finite strain version of the theory is adopted in order to
gciso /l gcani /l
10 Funda Aksu Denli et al.

20 40 response at a single Gauss point from which the model pa-


rameters are obtained. For the related values and units see
20 Table 1. Figure 6 compares the model results with the ex-
perimental data of the AS4/3501-6 epoxy lamina. The nu-
20 merical results agree favorably with the experimental data
(a) both under tensile and compressive loads, a gradually dimin-
F ishing mechanical response under compression is observed
upon reaching the ultimate stress value.
The analyses for the Mode–I and Mode–II tests are now
F conducted for six different fiber angles namely 0◦ , 15◦ , 30◦ ,
45◦ , 60◦ and 90◦ (measured from the horizontal direction),
F which are shown in Fig. 7. In the Mode–I test, an incre-
mental load is applied at the beginning for every specimen
(b) (c)
to find out the node–specific displacements, which are con-
F sidered as the node–specific displacement increments dur-
ing the rest of the analysis, thereby retaining the smoothness
Fig. 5: Single edge–notched specimen with Mode–I and Mode–II load-
ing: (a) dimensions of the specimen with a notch; (b) Mode–I load case;
of the surfaces on which the displacements are exerted. As
(c) Mode–II load case. All dimensions are in millimeter. for Mode–II the computations are performed with constant
displacement increments. The crack patterns pertaining to
Mode–I and Mode–II are illustrated in Fig. 7, in which it
Table 1: Model parameters λ, µ, µf , gciso and gcani with related values can be seen that the propagation of the crack mostly follows
and units.
the orientation of the fibers for all cases simulated.
Parameter Value Unit The respective load–displacement curves are depicted in
λ 5.2 × 103 [MPa] Fig. 8. The curves attributed to Mode–I (M1) suggest that a
µ 4.04 × 103 [MPa] 0◦ fiber angle exhibits the highest stiffness response, while
µf 64.6 × 103 [MPa]
a 90◦ fiber angle experiences the highest strength. Part of
gciso 5.5 [MPa mm]
the reason for such a distinct behavior may be that the flanks
gcani 80.0 [MPa mm]
of the domain act, in a sense, as a cantilever beam under
bending due to Mode–I, thereby leading to a higher stiffness
consider (i) the geometrical nonlinearities during the crack
value for the fiber orientation characterized by 0◦ degree.
propagation phase, and (ii) the large strains observed around
As for a 90◦ fiber angle, the branching of the crack upon the
the crack tip. Finally a (more) realistic test case of a unidi-
onset probably causes the highest strength among the cases
rectional laminate with an initial notch undergoing mixed–
tested. However, both the stiffness and strength values are
mode bending (MMB) is examined (Section 5.2).
in favor of the 0◦ fiber angle when it comes to the Mode–II
(M2) test. The crack needs to rupture more fibers on its way
for a unit vertical distance.
5.1 Mode–I and Mode–II tests for single edge–notched
specimens with various fiber orientations

Consider a rectangular plate with a horizontal notch placed 5.2 Mixed–mode bending of a unidirectional CFRP
in the middle of its height starting from the left edge. The
dimensions of the specimen with the notch together with the In this example the emphasis is on a thin rectangular carbon
Mode–I and Mode–II load cases are depicted in Fig. 5. For a fiber–reinforced polymer composite (CFRP) with a notch
discretization with 15 000 standard displacement finite ele- in the middle subject to MMB. The information regard-
ments, an element size of h = 0.4 mm is used over the whole ing the test apparatus, test procedure and the results are
domain and the length–scale parameter l is chosen to be 2.5 obtained from Crews & Reeder [12] and Naghipour et al.
times the element size. For the analysis, the plane strain as- [36]. In MMB, the delamination of the CFRP occurs under
sumption is used, and only one element spans the thickness combined influence of normal (Mode–I) and shear/sliding
of the plate. The anisotropy parameter ωf0 is set to unity. The (Mode–II) stresses. MMB tests make it possible to describe
material is chosen to be a unidirectional AS4/3501-6 epoxy the delamination resistance of a CFRP specimen and to ac-
lamina with the (reference) fiber direction [f 0 ] = [1, 0, 0] count for the effects of combined stresses by using a single
(horizontal fibers). In order to test how the proposed frame- test apparatus. A geometrical sketch of the testing device
work captures experimental data, we make use of a set of consisting of a load F and a loading lever with length c is
data provided by Soden et al. [43], and simulate the model shown in Fig. 9(a) in the undeformed configuration.
A phase–field model for fracture of unidirectional fiber–reinforced polymer matrix composites 11

2000 −250

Num. Num.
1600 −200
Exp. Exp.

1200 −150
σ1 [MPa]

σ2 [MPa]
800 −100

400 −50

0 0
−1 0 1 2 3 4 −4 −3 −2 −1 0
(a) ε2 [%] ε1 [%] (b) ε2 [%] ε1 [%]

Fig. 6: Numerical prediction versus experimental data for an AS4/3501-6 epoxy lamina – stress σ versus strain ε in the 1 (horizontal) and 2
(vertical) direction: (a) tension in the 1 direction; (b) compression in the 2 direction.

A more detailed illustration of the loading acting on the


Table 2: Model parameters λ, µ, µf , gciso and gcani of an APC2 lamina
hinge supports along with the superposition of the loads de- with related values and units.
lineating Mode–I and Mode–II is shown in Fig. 10. Therein,
Parameter Value Unit
a stands for the initial crack length (distance between the
λ 5.84 × 103 [MPa]
loading direction and the crack tip) acting as a delamination
µ 3.61 × 103 [MPa]
initiator, whereas L characterizes the specimen half-span.
µf 62.0 × 103 [MPa]
In particular, the loading position c can be manipulated to
gciso 1.82 [MPa·mm]
generate a (pure) Mode–II loading case where F is directly
gcani 33.0 [MPa·mm]
above the beam mid-span (c = 0). By removing the load-
ing lever and pulling up the hinge, one can achieve a (pure)
Mode–I loading scenario. It is worth to note that the relation- cross-head displacement rate of 0.5 mm/min. Therein, dif-
ship between the deflection δc at the specimen half-span, at ferent mode mixtures were considered. The related lengths
the hinge δhinge and the total displacement δMMB that occurs of the loading lever c for each mode mixture, namely 30%
at the loading point is and 50%, are 98.5 mm and 65.0 mm, respectively, and the
c resulting load–displacement (F -δMMB ) behavior of the uni-
δMMB = δc + δhinge with δhinge = δc +δMode−I , (57) directional layup is shown as dashed curves in Fig. 12.
L
where δMode−I is the displacement at the hinge associated
5.2.2 Numerical simulation
with Mode–I loading. For an illustration of the deformed
state see Fig. 9(b).
The material properties of the APC2 lamina are obtained
at a Gauss point, which are summarized in Table 2, while
5.2.1 Properties of the test specimen, experiments the (reference) fiber direction is [f 0 ] = [1, 0, 0] (horizon-
tal fibers). Figure 13 provides the fitting to the experimen-
The used material for the test specimen is APC2-prepreg, tal data, i.e. the in–plane stress values pertaining to the
which is composed of AS4-fibers (60% of the total vol- fiber (horizontal) and transverse (vertical) directions, see
ume) embedded in a polyether ether ketone (PEEK) matrix. Naghipour et al. [35]. In accordance with the loading de-
Each prepreg layer possesses a thickness of 140 µm. In total, scriptions characterized in Fig. 9, an in silico replica of the
24 carbon/PEEK unidirectional laminae ([0]24 ) are consid- specimen is made and then discretized. In order to better re-
ered in the layup yielding a final specimen size of 25 mm solve the crack pattern the mesh is refined around the crack
width, 150 mm length and 3.12 mm thickness. Furthermore, tip yielding a discretization of 6 000 brick elements with an
a 50 mm film is placed as a delamination initiator in the mid– effective element size of h = 0.065 mm in the refined zone.
plane, as indicated in Fig. 11. In accordance with the stan- Appropriate boundary conditions are applied to avoid rigid
dards established by ASTM [1], Naghipour et al. [35] con- body motions, see Fig. 11. In the analyses the plane strain
ducted mixed–mode experiments by using the MMB test- assumption is applied with a single element used in the di-
ing device set up according to Fig. 9, and by applying a rection of the width. Two sets of analyses are carried out: (i)
12 Funda Aksu Denli et al.

θ
θ
Mode–I
Mode–II

(a) θ = 0◦ (b) θ = 15◦ (c) θ = 30◦

θ θ θ
Mode–I
Mode–II

(d) θ = 45◦ (e) θ = 60◦ (f) θ = 90◦


0 1

Fig. 7: Evolution of the crack phase–field d for Mode–I and Mode–II tests in single edge–notched specimens with different fiber angles θ: (a)
θ = 0◦ ; (b) θ = 15◦ ; (c) θ = 30◦ ; (d) θ = 45◦ ; (e) θ = 60◦ ; (f) θ = 90◦ .

with the anisotropy parameter ωf0 = 1 and (ii) with ωf0 = fying the empirical requirement that l ≥ 2h. In other words,
30. Two different mode mixtures are taken into account in the length–scale parameter considered is greater than two
which the length–scale l is considered to be 0.15 mm satis- times the minimum element size. The applied load F and
A phase–field model for fracture of unidirectional fiber–reinforced polymer matrix composites 13

2.1 2.1
M1: θ = 45◦
1.8 1.8 M1: θ = 60◦
M1: θ = 90◦
1.5 1.5

Load F [kN]
Load F [kN]

1.2 1.2

0.9 0.9

0.6 0.6
M1: θ = 0◦
0.3 M1: θ = 15◦ 0.3
M1: θ = 30◦
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Displacement u [mm] Displacement u [mm]
(a) (b)

9.0 9.0
M2: θ = 45◦
7.5 7.5 M2: θ = 60◦
M2: θ = 90◦
6.0 6.0
Load F [kN]

Load F [kN]

4.5 4.5

3.0 3.0
M2: θ = 0 ◦

1.5 M2: θ = 15◦ 1.5


M2: θ = 30◦
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4

Displacement u [mm] Displacement u [mm]

(c) (d)

Fig. 8: Relationships between load F and displacement u for Mode–I (M1) and Mode–II (M2) tests on single edge–notched specimens with
various fiber angles, namely θ = 0◦ , 15◦ , 30◦ , 45◦ , 60◦ and 90◦ : (a),(b) Mode–I; (c),(d) Mode–II.

the chosen specimen half-span L are 1 N and 62.5 mm, re-


Table 3: Lengths c, proportional loads and related displacements for
spectively, from which the proportional loads exerted on the each mode mixture.
specimen are calculated. Afterwards, the displacement ra-
Mode mixture 30% 50%
tios corresponding to the calculated proportional loads are
c [mm] 98.5 65.0
ascertained, and applied at the loading points, as depicted
Hinge load [N]/displacement [mm] 1.58/2.14 1.04/1.34
in Fig. 10. The lengths c, the proportional (hinge/middle)
Middle load [N]/displacement [mm] 2.58/1.0 2.04/1.0
loads and the respective displacement ratios for each mode
mixture are listed in Table 3. As a matter of fact, the analy-
ses are carried out displacement–driven. The loading speed the displacements at the hinge δhinge and the middle δc of the
is 5 mm/min with the time increment ∆t = 0.5 that runs specimen together with the related reaction forces are stored
until the onset of the macro–crack. From this point onward, during each analysis. The applied load F is then calculated
the time increment is reduced to ∆t = 0.05, which is fol- via the balance of moment with respect to the mid–point
lowed by ∆t = 0.005 during the crack propagation. To ob- of the lever, i.e. F = Fhinge L/c, for every time increment,
tain the corresponding load–displacement (F –δMMB ) curve where Fhinge is the reaction force at the hinge. In addition,
from the points on which the loads are applied, the values of (57) is exploited so as to compute the corresponding values
of δMMB .
14 Funda Aksu Denli et al.

c
F
Loading lever
a

Hinge

Specimen

Apparatus base

(a) L L

δMode−I

δc

δMMB

(b)
c L

Fig. 9: (a) Geometrical sketch of the MMB testing device with the specimen in the undeformed configuration; (b) in the deformed configuration
with the corresponding displacements (reconstructed from Crews & Reeder [12].

5.2.3 Results and discussion marily in line with those conducted for ωf0 = 1 during the
initial phase of macro–cracking, giving rise to a sharp reduc-
tion in the load bearing capacity, as indicated in Fig. 12(b).
The anisotropic phase–field approach is examined for uni-
Nevertheless, the material tends to sustain a constant amount
directional composite laminates by comparing the experi-
of load as the crack propagates further, while damaging the
mental data with the finite element results for the two dif-
interlaminar medium between the laminae. The crack path
ferent mode mixtures namely 30% and 50%, see Fig. 12(a),
starts to better trail the direction along which the reinforcing
where a slight anisotropy is incorporated into the phase–field
fibers are embedded, see Fig. 14(b), which stands in sharp
model (ωf0 = 1). A close examination indicates an agree- contrast to the former, i.e. ωf0 = 1.
ment between the experimental and numerical results for the
two mode mixtures. The largely linear initial response of If we are to assess the discrepancy created when a rela-
the unidirectional laminate precedes with a rather abrupt de- tively strong anisotropic geometric resistance to cracking is
cline, leading the fracture process zone to grow. The growth assumed (ωf0 = 30), it is evident that the crack no longer
shows an inclined pattern towards the top of the specimen cut across the laminae located on the top of each other, but
while rupturing several laminae on its way, as presented in starts to evolve at the interface between the laminae towards
Fig. 14(a). The simulations performed for ωf0 = 30 are pri- the right end of the specimen, thereby resembling peel tests,
A phase–field model for fracture of unidirectional fiber–reinforced polymer matrix composites 15

F (L+c)
Fc
L L

L L F (c−L)
F (L+c)
2L
2L

Mode–I loading
= F (3c−L)
4L

F (3c−L)

+ 4L

F (L+c) F (L+c)
L 4L
Mode–II loading

F (L+c) F (L+c)
2L 4L

Fig. 10: Illustration of loading of the MMB test specimen decomposed according to Mode–I and Mode–II cases (reconstructed from Crews &
Reeder [12].

δhinge
δc

25

3.12

25 50

150

Fig. 11: Dimensions of the specimen with a 50 mm film placed in the mid–plane serving as a delamination initiator. Also shown are the boundary
conditions and displacement at the hinge and the middle of the specimen, i.e. δhinge and δc . All dimensions are in millimeters.

see, e.g., Gültekin et al. [19], where a relatively constant top of the specimen, which eventually results in the failure
load drives the peeling of the specimen parallel to the fibers of one of the arms. As a consequence, a more abrupt failure
in the post–cracking phase. However, in the case of ωf0 = 1 of the entire system occurs, which is seen in Fig. 12(a).
the crack evolution is driven across the laminate until the
16 Funda Aksu Denli et al.

300 300
Exp. 30% Exp. 30%
Num. 30% Num. 30%
250 Exp. 50% 250 Exp. 50%
Num. 50% Num. 50%
200 200
Load F [N]

Load F [N]
150 150

100 100

50 50

0 0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5
(a) δMMB [mm] (b) δMMB [mm]

Fig. 12: Experimental (dashed curves) and numerical (solid curves) results for the relationship between the load F and the loading point displace-
ment δMMB for unidirectional laminates with 30% and 50% mode mixtures: (a) ωf0 = 1; (b) ωf0 = 30.

2.5 −200
Num.
Exp.
2 −160
σ2 [MPa]

1.5 −120
σ1 [GPa]

1 −80

Num.
0.5 −40 Exp.

0 0
−1 −0.5 0 0.5 1 1.5 2 2.5 −2.5 −2 −1.5 −1 −0.5 0 0.5
(a) ε2 [%] ε1 [%] (b) ε2 [%] ε1 [%]
Fig. 13: Numerical prediction versus experimental data for an APC2 lamina – stress σ versus strain ε in the 1 (horizontal) and 2 (vertical) direction:
(a) tension in the 1 direction; (b) compression in the 2 direction.

6 Conclusion anisotropic fracture of a single–edged notch plate in re-


sponse to Mode–I and Mode–II loadings by altering the ori-
This study presents one of the few attempts to model fail- entation of fibers, while the second example offers an analy-
ure of engineered composite materials by making use of an sis of a unidirectional CFRP laminate under MMB, touching
anisotropic crack phase–field approach. A concise yet criti- upon different fracture zones in relation to the anisotropy pa-
cal outline of previous contributions pertaining to strength– rameter. It is noteworthy that the fit of the anisotropic model
based and energy–based criteria for failure of composites shows a satisfactory coherence with experimental data. The
is given. Subsequently, the theoretical backbones of the proposed model approach serves a sound basis for more ad-
anisotropic phase–field approach is reviewed which relies vanced analyses of crack initiation and propagation in uni-
on the superposition of the distinct fracture processes asso- directional fiber–reinforced polymer matrix composites.
ciated with the matrix and fibrous content. The phase–field Acknowledgment H.D. gratefully acknowledges the finan-
model of fracture is essentially modular consisting of two cial support from Tübitak grant scheme Bideb 2232 under
sub–problems emanating from the deformation field ϕ and grant number 114C073.
the crack phase–field d. The numerical examples and the
discussions focus on the crack initiation via fitting to ex-
References
perimental data and the direction on which the crack finds
its path through relevant tests, such as Mode–I, Mode–II 1. ASTM International. ASTM D6671-01 standards test method for
and MMB. In particular, the first example scrutinizes the mixed mode I–mode II interlaminar fracture toughness of unidi-
A phase–field model for fracture of unidirectional fiber–reinforced polymer matrix composites 17

30% mode mixture 50% mode mixture

(a)

(b)
0 1

Fig. 14: Evolution of the crack phase–field d with respect to the mode mixture of 30% (left panels) and 50% (right panels) for two anisotropy
parameters: (a) ωf0 = 1; (b) ωf0 = 30.

rectional fiber reinforced polymer matrix composites. ASTM In- 10. J. W. Cahn and J. E. Hilliard. Free energy of a nonuniform system.
ternational, West Conshohocken, PA, USA, 2001. I. Interfacial free energy. J. Chem. Phys., 28:258–267, 1958.
2. R. Alessi and F. Freddi. Phase-field modelling of failure in hybrid 11. J. D. Clayton and J. Knap. Phase field modeling of directional
laminates. Composite Structures, 181:9–25, 2017. fracture in anisotropic polycrystals. Comput. Mat. Sci., 98:158–
3. M. Ambati, T. Gerasimov and L. De Lorenzis. Phase–field mod- 169, 2015.
eling of ductile fracture. Comput. Mech., 55:1017–1040, 2015. 12. J. H. Crews and J. R. Reeder. A mixed–mode bending apparatus
4. A. Arteiro, G. Catalanotti, J. Reinoso, P. Linde and P. P. Camanho. for delamination testing. NASA TM 100662, 1988.
Simulation of the mechanical response of thin-ply composites: 13. R.G. Cuntze. Efficient 3D and 2D failure conditions for UD lam-
from computational micro-mechanics to structural analysis. Arch. inae and their application within the verification of the laminate
Comput. Meth. Engrg., 26:1445–1487, 2019. design. Compos. Sci. and Tech., 66:1081–1096, 2006.
14. H. Dal, O. Gültekin, F. A. Denli and G. A. Holzapfel. Phase–field
5. A. S. Argon. Fracture of composites. Treatise Mater. Sci. Technol.,
models for the failure of anisotropic continua. Proc. Appl. Math.
1:79–114, 1972.
Mech., 17: 91–94, 2017.
6. V. D. Azzi and S. W. Tsai. Anisotropic strength of composites. 15. G. A. Francfort and J.-J. Marigo. Revisiting brittle fracture as an
Exp Mech., 5:283–288, 1965. energy minimization problem. J. Mech. Phys. Solids, 46:1319–
7. J. Betten. Formulation of anisotropic constitutive equations. In 1342, 1998.
J. P. Boehler (ed.), Applications of Tensor Functions in Solid 16. A. A. Griffith. The phenomena of rupture and flow in solids. Phi-
Mechanics Springer-Verlag, Wien, 1987, pp. 228–250. CISM los. T. Roy. Soc. A, 221:163–197, 1921.
Courses and Lectures no. 292. 17. D. M. Grogan, S. B. Leen and C. M. Ó Brádaigh. An XFEM-
8. M. J. Borden, T. J. R. Hughes, C. M. Landis and A. Anvari. A based methodology for fatigue delamination and permeability of
phase–field formulation for fracture in ductile materials: Finite de- composites. Composite Structures, 107:205–218, 2014.
formation balance law derivation, plastic degradation, and stress 18. O. Gültekin, H. Dal and G. A. Holzapfel. A phase-field approach
triaxiality effects. Comput. Meth. Appl. Mech. Engrg. 312:130– to model fracture of arterial walls: Theory and finite element anal-
166, 2016. ysis. Comput. Meth. Appl. Mech. Engrg., 312:542–566, 2016.
9. J. P. Boehler. A simple derivation of representations for non- 19. O. Gültekin, H. Dal and G. A. Holzapfel. Numerical aspects of
polynomial constitutive equations in some cases of anisotropy. failure in soft biological tissues favor energy-based criteria: A
ZAMM – Zeitschrift für Angewandte Mathematik und Mechanik, Rate-dependent anisotropic crack phase–field model. Comput.
59:157–167, 1979. Meth. Appl. Mech. Engrg., 331:23–52, 2018.
18 Funda Aksu Denli et al.

20. Z. Hashin. Failure criteria for unidirectional fiber composites. J. 40. J. Reinoso, A. Arteiro, M. Paggi and P. P. Camanho. Strength
Appl. Mech., 47:329–334, 1980. prediction of notched thin ply laminates using finite fracture me-
21. Z. Hashin. Finite thermoelastic fracture criterion with application chanics and the phase field approach. Compos. Sci. and Tech.,
to laminate cracking analysis. J. Mech. Phys. Solids, 44:1129– 58:205–216, 2017.
1145, 1996. 41. C. Schreiber, C. Kuhn, and R. Müller. Phase field modeling of
22. R. Hill. The mathematical theory of Plasticity. New York: Oxford brittle fracture in materials with anisotropic fracture resistance.
University Press, 1998. PAMM, 18:e201800113, 2018.
23. G. A. Holzapfel. Nonlinear Solid Mechanics. A Continuum Ap- 42. J. Schröder and P. Neff. Invariant formulation of hyperelastic
proach for Engineering. John Wiley & Sons, Chichester, 2000 transverse isotropy based on polyconvex free energy functions.
24. G. A. Holzapfel, T. C. Gasser and R. W. Ogden. A new constitutive Int. J. Solids Structures, 40:401–445, 2003.
framework for arterial wall mechanics and a comparative study of 43. P. D. Soden, M. J. Hinton and A. S. Kaddour. Lamina properties,
material models. J. Elasticity, 61:1–48, 2000. lay-up configurations and loading conditions for a range of fibre
25. B. Li, C. Peco and D. Millán, I. Arias and M. Arroyo. Phase- reinforced composite laminates. In: M. Hinton, P. D. Soden and
field modeling and simulation of fracture in brittle materials with A.-S. Kaddour (eds.), Failure Criteria in Fibre–Reinforced Poly-
strongly anisotropic surface energy. Int. J. Numer. Meth. Engng., mer Composites, Elsevier Ltd, 2004, pp. 30–51.
102:711–727, 2015. 44. A. Spencer. Deformations of Fibre–Reinforced Materials. Oxford
University Press, 1972.
26. J. E. Marsden and T. J. R. Hughes. Mathematical Foundations of
45. S. Teichtmeister, D. Kienle, F. Aldakheel, and M.-A. Keip. Phase-
Elasticity. Dover, New York, 1994.
field modeling of fracture in anisotropic brittle solids. Int. J. Non-
27. C. Miehe, F. Welschinger and M. Hofacker. Thermodynamically Linear Mech., 97:1–21, 2017.
consistent phase-field models of fracture: Variational principles 46. R. Talreja and C. V. Singh. Damage and Failure of Composite
and multi-field FE implementations. Int. J. Numer. Meth. Engng., Materials. Cambridge University Press, New York, 2012
83:1273–1311, 2010. 47. R. Talreja. Assessment of the fundamentals of failure theories for
28. C. Miehe, M. Hofacker and F. Welschinger. A phase field model composite materials. Compos. Sci. and Tech., 105:190–201, 2014.
for rate-independent crack propagation: Robust algorithmic im- 48. S. W. Tsai and E. M. Wu. General theory of strength for
plementation based on operator splits. Comput. Meth. Appl. Mech. anisotropic materials. J. Compos. Mater., 5:58–80, 1971.
Engrg., 199:2765–2778, 2010. 49. A. Turon, C. G. Dávila, P. P. Camanho and J. Costa. An engineer-
29. C. Miehe, L. M. Schänzel and H. Ulmer. Phase field modeling of ing solution for mesh size effects in the simulation of delamination
fracture in multi-physics problems. Part I. Balance of crack sur- using cohesive zone models. Eng. Frac. Mech., 74:1665–1682,
face and failure criteria for brittle crack propagation in thermo- 2007.
elastic solid. Comput. Meth. Appl. Mech. Engrg., 294:449–485, 50. Y. Wang and H. Waisman. Progressive delamination analysis of
2015. composite materials using XFEM and a discrete damage zone
30. C. Miehe, M. Hofacker, L. M. Schänzel and F. Aldakheel. Phase mode. Comput. Mech., 55:1–26, 2015.
field modeling of fracture in multi-physics problems. Part II. Cou- 51. W. E. Wolfe and T. S. Butalia. A strain-energy based failure cri-
pled brittle-to-ductile failure criteria and crack propagation in terion for non-linear analysis of composite laminates subjected to
thermo-elastic-plastic solids. Comput. Meth. Appl. Mech. Engrg., biaxial loading. Compos. Sci. and Tech., 58:1107–1124, 1998.
294:486–522, 2015. 52. Q. Yang and B. Cox. Cohesive models for damage evolution in
31. C. Miehe, H. Dal, L. M. Schänzel, and A. Raina. A phase-field laminated composites. Int. J. Fracture, 133:107–137, 2005.
model for chemo-mechanical induced fracture in lithium-ion bat- 53. S. Yazdani, W. J. H. Rust and P. Wriggers. An XFEM approach
tery electrode particles. Int. J. Numer. Meth. Engng., 106:683–711, for modelling delamination in composite laminates. Composite
2016. Structures, 135:353–364, 2016.
32. R. von Mises. Mechanik der festen Körper im plastisch de- 54. L. Zhao, Y. Gong, J. Zhang, Y. Chen and B. Fei. Simulation of
formablen Zustand. Göttin. Nachr. Math. Phys., 1:582–592, 1913. delamination growth in multidirectional laminates under mode I
33. J. A. Nairn. The strain-energy release rate of composite and mixed mode I/II loadings using cohesive elements. Composite
microcracking–a variational approach. J. Compos. Mater., Structures, 116:509–522, 2016.
23:1106–1129, 1989.
34. P. Naghipour, M. Bartsh and H. Voggenreiter. Simulation and ex-
perimental validation of mixed mode delamination in multidirec-
tional CF/PEEK laminates under fatigue loading. Int. J. Solids and
Structures, 48:1070–1081, 2011.
35. P. Naghipour. Numerical simulations and experimental investiga-
tions on quasi-static and cyclic mixed mode delamination of multi-
directional CFRP laminates. Ph.D. Thesis, University of Stuttgart,
2011.
36. P. Naghipour, J. Schneider, M. Bartsch, J. Hausmann and
H. Voggenreiter. Fracture simulation of CFRP laminates in mixed
mode bending. Eng. Frac. Mech., 76:2821–2833, 2009.
37. T.-T. Nguyen, J. Réthéro, J. Yvonnet, and M.-C. Baietto. Multi-
phase-field modeling of anisotropic crack propagation for poly-
crystalline materials. Comput. Mech., 60:289–314, 2017.
38. A. Puck and H. Schürmann. Failure analysis of FRP laminates by
means of physically based phenomenological models. Compos.
Sci. and Tech., 58:1045–1067, 1998.
39. G. Qiu and T. Pence. Remarks on the behavior of simple direc-
tionally reinforced incompressible nonlinearly elastic solids. J.
Elasticity, 49:1–30, 1997.

View publication stats

Vous aimerez peut-être aussi