Computational Aspects of Cohesive Zone Models

download Computational Aspects of Cohesive Zone Models

of 8

Transcript of Computational Aspects of Cohesive Zone Models

  • 8/20/2019 Computational Aspects of Cohesive Zone Models

    1/18

    COMPUTATIONAL ASPECTS OF

    COHESIVE–ZONE MODELS

    René de Borst, Joris J.C. Remmers

    Faculty of Aerospace Engineering, Delft University of Technology

    P.O. Box 5058, NL-2600 GB Delft, The Netherlands

    e-mail: [email protected] / [email protected]

    Alan Needleman

    Division of Engineering, Brown University

    Providence, Rhode Island 02912, U.S.A.

    e-mail: [email protected]

    Abstract

    The importance of the cohesive–zone approach to analyse localisation and fracture in engi-

    neering materials is emphasised and ways to incorporate the cohesive–zone methodology in

    computational methods are discussed. Numerical implementations of cohesive–zone mod-

    els suffer from a certain mesh bias. For discrete representations this is caused by the initial

    mesh design, while for smeared representations it is rooted in the ill–posedness of the rate

    boundary value problem that arises upon the introduction of decohesion. A proper represen-

    tation of the discrete character of cohesive–zone formulations which avoids any mesh biascan be obtained elegantly when exploiting the partition-of-unity property of finite element

    shape functions. A recent development, so-called cohesive crack segments, which are par-

    ticularly suited to simulate the entire process of crack nucleation, growth and coalescence is

    presented. The effectiveness of this approach is demonstrated by some examples for crack 

    nucleation and growth in heterogeneous materials and for fast crack growth.

    Introduction

    Linear elastic fracture mechanics applies when there is a crack–like flaw in an otherwise

    linear elastic solid and the singularity associated with that flaw is characterised by a non-

    vanishing energy release rate. For example, linear elastic fracture mechanics concepts do

    not apply for dynamic crack growth in a brittle solid in the intersonic regime. The fracture

    and any dissipative processes must also remain confined to a small region in the vicinity of 

    the crack tip. If these conditions are not met, linear elastic fracture mechanics concepts do

    not apply and another fracture framework is needed.

    When the region in which the separation and dissipative process take place is not small

    compared to a structural dimension, but any nonlinearity is confined to a surface emanating

    from a classical crack tip (i.e. one with a non-vanishing energy release rate), cohesive zone

    models which were introduced by Barenblatt [1] and Dugdale [2] can be applied. Subse-

  • 8/20/2019 Computational Aspects of Cohesive Zone Models

    2/18

    σ  

    u

     f t

    σ  

     f t

    u

    Figure 1: Stress–displacement curves for a ductile solid (left) and a quasi–brittle solid (right).

    quently, the cohesive zone approach was extended by Hillerborg et al.  [3] and Needleman [4]

    to circumstances where: (i) an initial crack–like flaw need not be present or, if one is present,

    it need not be associated with a non-vanishing energy release rate; and (ii) non-linear de-

    formation behaviour (but not separation) may occur over an extended volume. In this viewof cohesive modelling, the continuum is characterised by two constitutive relations: a vol-

    umetric relation that relates stress and strain and a cohesive relation that relates traction to

    separation across a set of cohesive surfaces. Thus, for a cohesive formulation to apply, what-

    ever fracture mechanism is occurring, it must be appropriate to idealise it as taking place

    over a surface of negligible thickness.

    Figure 1 shows some commonly used decohesion relations, one for ductile fracture (left)

    [5] and one for quasi-brittle fracture (right) [6]. For ductile fracture, the most important

    parameters of the cohesive–zone model appear to be the tensile strength  f t  and the work of separation or fracture energy G c  [7], which is the work needed to create a unit area of fully

    developed crack. It has the dimensions J/m2 and is formally defined as:

    G c =

       ∞

    u=0

    σdu   (1)

    with σ  and u   the stress and the relative displacement across the fracture process zone. Formore brittle decohesion relations, as shown for instance in the right part of Figure 1, i.e.,

    when the decohesion law stems from micro-cracking as in concrete or ceramics, the  shape

    of the stress–separation relation plays a larger role and is sometimes more important than the

    value of the tensile strength f t [8, 9].

    Numerical representations of cohesive–zone models

    When the crack path is known in advance, either from experimental evidence, or because of 

    the structure of the material (such as in laminated composites), cohesive–zone models have

    been used with considerable success. In those cases, the mesh can be constructed such that

    the crack path  a priori  coincides with the element boundaries. Such a cohesive crack can

    be modelled by inserting interface elements between continuum elements along the potential

    crack path.

    In such analyses, interface elements equipped with a cohesive constitutive relation are

    inserted a priori in the finite element mesh. Introducing interface elements from the out-

    set of an analysis requires a finite stiffness prior to the onset of cracking. This changes the

  • 8/20/2019 Computational Aspects of Cohesive Zone Models

    3/18

    (b)(c)

    (a)

    Figure 2: Experimentally observed ‘diffuse’ crack pattern [17] and possible numerical rep-

    resentation.

    compliance of the structure and gives rise to deformations in the interface before crack ini-

    tiation. Furthermore, in dynamic analyses, wave speeds change. Although these effects canbe limited by choosing a sufficiently high initial stiffness, they still remain. For example, de-

    pending on the chosen spatial integration scheme, a high initial stiffness can lead to spurious

    traction oscillations, which may cause erroneous crack patterns, e.g. [10].

    An example where the potential of cohesive–zone models can be exploited fully using

    traditional discrete interface elements, is the analysis of delamination in layered composite

    materials [11, 12]. Since the propagation of delaminations is then restricted to the interfaces

    between the plies, inserting interface elements at these locations permits an exact simulation

    of the failure mode.

    To allow for a more arbitrary direction of crack propagation, Xu and Needleman [13]

    have inserted interface elements equipped with a cohesive–zone model between  all  contin-uum elements. A related method, using remeshing, was proposed by Camacho and Or-

    tiz [14]. Although such analyses provide much insight, they suffer from a certain mesh bias,

    since the direction of crack propagation is not entirely free, but is restricted to interelement

    boundaries [15]. Another drawback is that the method is not suitable for large-scale anal-

    yses. For these reasons smeared numerical representations of cohesive–zone models have

    appeared, including the emergence of some, initially not foreseen mathematical difficulties,

    which can only be overcome in a rigorous fashion by resorting to higher-order continuum

    models, see [16] for a more in-depth analysis and an overview of possible remedies.

    Crack propagation in heterogeneous media

    The physics of crack initiation and crack growth in a heterogeneous quasi–brittle material are

    illustrated in Figure 2 [17]. The heterogeneity of the material, i.e., the presence of particles of 

    different sizes and stiffnesses leads to a complex stress field where new cracks nucleate (‘a’

    in Figure 2) and existing cracks branch (‘b’). Moreover, cracks may propagate across phase

    boundaries (‘c’), or along interfaces (‘d’). Smeared crack models do not properly capture

    these processes of crack initiation, growth, coalescence and branching, because essential

    features can be lost in the smoothing process.

    More detail is preserved if we model the initiation, growth and eventual coalescence

  • 8/20/2019 Computational Aspects of Cohesive Zone Models

    4/18

    of the cracks at the mesoscopic level of observation in Figure 2   separately. Hitherto this

    could not be carried out, not only because of the high computational effort that this would

    require, but also because a suitable numerical framework was lacking. The exploitation of 

    the partition-of-unity property of finite element shape functions [18, 19, 20] can make such

    calculations feasible. As a further extension to a continuous cohesive crack that runs through

    an existing finite element mesh without bias [21, 22, 23], one can define  cohesive segments

    that can arise at arbitrary locations and in arbitrary directions and allow for the resolution of 

    complex crack patterns including crack nucleation at multiple locations, followed by growth

    and coalescence [24].

    The partition-of-unity approach

    Governing equations

    A finite element method accommodating the propagation of discrete cracks through elements

    in an arbitrary manner was proposed by Belytschko and his co-workers [19, 20], exploiting

    the partition-of-unity property of finite element shape functions [18]. Since finite element

    shape functions ϕi form partitions of unity,n

    i=1 ϕi = 1 with n the number of nodal points,a field u  can be interpolated as

    u =n

    i=1

    ϕi

    āi +

    m j=1

    ψ jãij

      (2)

    with  āi   the ‘regular’ nodal degrees-of-freedom,  ψ j   the enhanced basis terms, and  ãij   the

    additional degrees-of-freedom at node i  which represent the amplitude of the j th enhancedbasis term ψ j .A basic assumption of the method is that a crack can be regarded as a discontinuity in the

    displacement field. Consider a domain Ω that is crossed by a single discontinuity at Γd. Thedisplacement field  u  can be written as the sum of two continuous displacement fields  ū andũ:

    u =  ū + HΓdũ   (3)

    where  HΓd   is the Heaviside step function centred at the discontinuity. The displacement

    decomposition in eq. (3) has a structure similar to the interpolation in eq. (2). This can be

    seen directly by rewriting and specialising eq. (2) as:

    u =  N(ā + HΓdã) =  Nā + HΓdNã =  ū + HΓdũ   (4)

    where N  contains the standard shape functions, and ā and ã collect the conventional and the

    additional nodal degrees-of-freedom, respectively. Accordingly, the partition-of-unity prop-

    erty of finite element shape functions can be used in a straightforward fashion to incorporate

    discontinuities, and thus discrete crack models, in a manner that preserves the discontinuous

    character of cracks.

     Example: delamination–buckling

    To exemplify the possibilities of this approach to model the combined failure mode of delam-

    ination growth and local buckling we consider the double cantilever beam of Figure 3 with

  • 8/20/2019 Computational Aspects of Cohesive Zone Models

    5/18

     

     

     

     

     

     

     

     

     

     

     

    P 0

    P 0, u

    a0   = 10mm

    b  = 1mm

    h  = 0.2mm

    l   = 20mm

    Figure 3: Double cantilever beam with initial delamination under compression.

    Debonding (coarse mesh)

    Debonding (dense mesh)

    Perfect bond 

    1 2 3 4 5 6 7 8u (mm)

    0

    1

    2

    3

    4

    Figure 4: Load–displacement curves for delamination–buckling test.

    an initial delamination length a = 10 mm. Both layers are made of the same isotropic linearelastic material with Young’s modulus E   = 135000 N/mm2 and Poisson’s ratio ν   = 0.18.Due to symmetry in the geometry of the model and the applied loading, delamination prop-

    agation can be modelled with an exponential mode–I decohesion law:

    tndis = tult exp

    −tultG c

    vndis

    ,   (5)

    where tndis and vndis are the normal traction and displacement jump, respectively. The ultimate

    traction tult is equal to 50 N/mm2

    , the work of separation is G c = 0.8 N/mm.This case, in which failure is a consequence of a combination of delamination growth and

    structural instability, has been analysed using conventional interface elements in [25]. The

    beam is subjected to an axial compressive force 2P , while two small perturbing forces P 0 areapplied to trigger the buckling mode. Two finite element discretisations have been employed,

    a fine mesh with three elements over the thickness and 250 elements along the length of the

    beam, and a coarse mesh with only one (!) element over the thickness and 100 elements

    along the length. Figure 4 shows that the calculation with the coarse mesh approaches the

    results for the fine mesh closely. For instance, the numerically calculated buckling load is in

    good agreement with the analytical solution. Steady–state delamination growth starts around

    a lateral displacement u  = 4   mm. From this point onwards, delamination growth interacts

  • 8/20/2019 Computational Aspects of Cohesive Zone Models

    6/18

    Figure 5: Deformation of coarse mesh after buckling and delamination growth (true

    scale) [26].

     

    0

    0.5

    1

    1.5

    2

    2.5

    3

    0 1 2 3 4 5 6

    Perfect bond

    Debonding (Gc=0.8)

    u [mm]

       P   [   N   ]

    Figure 6: Load–displacement curve and deformations of shell model after buckling and de-

    lamination growth (true scale) [27].

    with geometrical instability. Figure 5 presents the deformed beam for the coarse mesh at a

    tip displacement u  = 6  mm. Note that the displacements are plotted at true scale, but that

    the difference in displacement between the upper and lower parts of the beam is mainly dueto delamination and the strains remain small.

    The good results obtained in this example for the coarse discretisation have motivated the

    development of a layered plate/shell element in which delaminations can occur inside the el-

    ement between each of the layers [27]. The example of Figure 3 has been reanalysed with

    a mesh composed of eight node enhanced solid–like shell elements [27]. Again, only one

    element in the thickness direction has been used. In order to capture delamination growth

    correctly, the mesh has been refined locally. Figure 6 shows the lateral displacement u  of the beam as a function of the external force  P . The load–displacement response for a speci-men with a perfect bond (no delamination growth) is given as a reference. The numerically

    calculated buckling load is in agreement with the analytical solution. Steady delamination

  • 8/20/2019 Computational Aspects of Cohesive Zone Models

    7/18

    t p

    Ω+2

    Γ

    nΓd,1

    Γd,2

    +

    1

    nΓd,2

    Γd,1

    Γt

    u p

    Γu

    Figure 7: Domain crossed by two discontinuities Γd,1 and Γd,2.

    growth starts around a lateral displacement u ≈ 4 mm, which is in agreement with previoussimulations [25].

    Cohesive crack segments

     Formulation

    A key feature of the cohesive segments approach is the possible emergence of multiple

    cohesive segments in a domain. Consider a domain Ω   which contains  m  discontinuitiesΓd,j, j  = 1,....,m, see Figure 7. Each discontinuity splits the domain in two parts, denotedas Ω−

     j  and Ω+

     j , such that Ω−

     j  ∪ Ω+

     j

      = Ω. Generalising eq. (3), the displacement field can bewritten as the sum of  m + 1  continuous displacement fields  ū and ũ j  [28]:

    u =  ū +m

     j=1

    HΓd,jũ j   (6)

    with HΓd,j  separating the continuous displacement fields  ū  and  ũ j . Restricting attention tosmall displacement gradients, the strain field follows by differentiation of (6):

     = ∇sū +

    m

     j=1HΓd,j∇

    sũ j +  δ Γd,j(ũ j ⊗ nΓd,j)

    s

      (7)where δ Γd,j  denotes the Dirac delta function placed at the  j

    th discontinuity Γd,j  and the su-perscript s  denotes the symmetric part of the tensor.

    One proceeds by defining test functions for the displacements in a Bubnov–Galerkin

    sense:

    ηηη = η̄ηη +

    m j=1

    HΓd,jη̃ηη j   (8)

    substitutes them into the momentum equation

    ∇ ·σσσ =

     0  (9)

  • 8/20/2019 Computational Aspects of Cohesive Zone Models

    8/18

    with σσσ the Cauchy stress tensor, and integrates over the domain  Ω:

     Ωη̄ηη +m

     j=1

    HΓd,jη̃ηη j · ∇ · σσσdΩ = 0   (10)Following a standard procedure, one applies the divergence theorem, uses the external bound-

    ary conditions, eliminates the Heaviside functions by changing the integration domain from

    Ω to Ω+ j   and eliminates the Dirac functions by transforming the volume integral into a surfaceintegral:

     Ω

    ∇η̄ηη :  σσσdΩ +m

     j=1

     Ω+j

    ∇η̃ηη j   : σσσdΩ +

     Γd,j

    η̃ηη j · td,jdΓ

    = Γη̄ηη +

    m

     j=1

    HΓd,j

    η̃ηη j · t p dΩ

    (11)

    with   td,j  the interface traction at  Γ j . The trial functions u  and the test functions ηηη  can bediscretised in a fashion similar to the left part of eq. (4):

    u   =   N

    ā +

    m j=1

    HΓd,j ã j

      (12)

    ηηη   =   N

    w̄ +

    m

     j=1HΓd,j w̃ j

      (13)

    Requiring that the result holds for all admissible  w̄   and  w̃ j, eq. (11) can be separated inm + 1  sets of equations:

    f intā

    f intã1

    ...

    f intãm

    =

    f extā

    f extã1

    ...

    f extãm

    ,   (14)

    where the internal forces are defined as:

    f intā   =

     Ω

    BTσσσdΩ ;   f intãj   =

     Ω+j

    BTσσσdΩ +

     Γd,j

    NTtd,jdΩ   j  = 1...m,   (15)

    and the external forces are equal to

    f extā   =

     Γ

    NTt pdΓ ;   f 

    extãj

      =

     Γ

    HΓd,jNTt pdΓ   j  = 1...m.   (16)

    The linearisation needed for the incremental-iterative solution procedure is described in

    Remmers et al. [24].

  • 8/20/2019 Computational Aspects of Cohesive Zone Models

    9/18

    The cohesive segments approach shares the advantages of the method that exploits the

    partition-of-unity property of finite element shape functions to describe continuous cohesive

    crack growth. Examples are the insensitivity of the direction of crack propagation to the

    structure of the underlying discretisation and bypassing the need to define a high initial

    stiffness at the interface, which can result in numerical artefacts such as interface traction

    oscillations and spurious wave reflections.

     Implementation aspects

    When the criterion for the initiation of decohesion is met — currently, a principal stress crite-

    rion is used, a cohesive segment is inserted through the integration point. In the applications

    so far, its direction has been taken to be orthogonal to the direction of the major principal

    stress. The segment is taken to extend throughout the element to which the integration point

    belongs and into the neighbouring elements, see Figure 8(a). The magnitude of the displace-

    ment jump is determined by a set of additional degrees of freedom which are added to allnodes whose support is crossed by the cohesive segment. The nodes of the element boundary

    that is touched by one of the two tips of the cohesive segment are not enhanced in order to

    ensure a zero opening at these tips [21]. Subsequently, the evolution of the separation of 

    the cohesive segment is governed by a decohesion constitutive relation in the discontinuity.

    When the criterion for the initiation is met at one of the two tips, the cohesive segment is ex-

    tended into a new element, as demonstrated in Figure 8(b). The extension is straight within

    the element, but does not necessarily have to be aligned with previous parts of the segment,

    so that curved crack paths can be simulated.

    Each cohesive segment is supported by its own set of additional degrees of freedom.

    When two segments meet within a single element, the nodes that support the element are

    enhanced twice, once for each cohesive segment. In the situation depicted in Figure 9(a),

    segment A is only extended until it touches segment B, which can be regarded as a free edge.

    This implies that there is no crack tip, so that all four nodes of the element are enhanced. A

    special case is shown in Figure 9(b). When two segments approach as shown in this figure,

    they are simply joined.

    Because the crack is not taken as a single entity  a priori   in the cohesive segments ap-

    proach, the method can equally naturally simulate distributed cracking which frequently

    occurs in a heterogeneous solid. Thus, the cohesive segments approach embraces both ex-

    tremes, distributed cracking with crack nucleation, growth and eventual coalescence at mul-

    tiple locations as well as the initiation and propagation of a single, dominant crack without

    requiring special assumptions. It is only needed to specify the conditions for crack nucleationand for the crack propagation direction, and a decohesion relation at the crack.

     Example: double-cantilever beam

    Some features of the cohesive segments method are illustrated in the following example.

    Consider the double cantilever beam with a small notch as shown in Figure 10. The beam

    is subjected to bending. The two layers of the beam are assumed to be isotropic linear

    elastic materials and have identical elastic properties: Young’s modulus  E  = 20.0 GPa andPoisson’s ratio ν   = 0.2. The cohesive tensile stress of the material is  f layt   = 2.5  MPa, thework of separation is G lay

    c

      = 40.0 N/m. It is assumed that the fracture mode is purely mode-I

  • 8/20/2019 Computational Aspects of Cohesive Zone Models

    10/18

    (b)(a)

    Ω+j

    Ω−j

    nd,j

    Figure 8: (a) A single cohesive segment in a quadrilateral mesh. The segment passes through

    an integration point (⊗) where the fracture criterion is violated. The solid nodes contain

    additional degrees of freedom that determine the magnitude of the displacement jump. The

    gray shade denotes the elements that are influenced by the cohesive segment. (b) A cohesive

    segment is extended into a new element (dashed line). The gray nodes contain degrees of freedom that have just been added to support the extension of the cohesive segment.

    (a) (b)

    B

    A

    Figure 9: (a) Interaction of two cohesive segments. Segment A is extended (dashed line)

    until it touches segment B. Since this can be regarded as a free edge, there will be no crack 

    tip for segment A. Consequently, all four nodes of the element will be enhanced in order to

    support the displacement jump of segment A (denoted by the gray nodes). (b) Two segments

    are connected (dashed line).

    and is modelled by the decohesion relation:

    tn  =  f layt   exp

    −f 

    layt

    G layc

    vn  ,   (17)

    where tn  and vn  are the normal traction and displacement jump respectively. The adhesivethat bonds the two layers is modelled with a mixed mode delamination model with a non-zero

    compliance prior to cracking. The traction across a cohesive segment follows from

    t = ∂ G 

    ∂ v .   (18)

    Here, assuming the normal and the shear work of separation are identical, the potential G   is

  • 8/20/2019 Computational Aspects of Cohesive Zone Models

    11/18

    P, u

    50 mm

    10 mm

    50 mm

    10 mm

    Figure 10: Geometry and loading conditions of a double-cantilever beam with an initial

    notch.

    tip displacement   u [mm]

        a    p    p      l     i    e      d      l    o    a      d     P

          [     N      ]

    0

    1

    2

    3

    4

    5

    0.1 0.3 0.4 0.50.2

    Figure 11: Load–displacement curve of the double-cantilever beam with an initial notch.

    given by [29]:

    G  = G adhc

    1 − (1 +

     vnδ n

    )exp

    −vnδ n

    exp

    −v2sδ 2s

      (19)

    where G adhc   is the work of separation, vn and  vs  are the normal and shear components of thedisplacement jump  v  and δ n  and δ s   the corresponding characteristic lengths, which followfrom:

    δ n  =  G adhc

    e tn,max;   δ s =

      G adhc 12e ts,max

    (20)

    with e  = exp(1) and tn,max and ts,max the normal and shear cohesive strengths, respectively.In this example, we assume that G adh

    c

      = 10.0 N/m, tn,max = ts,max = 1.0 MPa.The specimen is analysed with a mesh having 99×21 elements. The notch is simulated by

    removing a single element from the mesh. The interface between the two layers is modelled

    by a cohesive segment, which is added to the mesh beforehand. This implies that one part of 

    the element that is crossed by this segment belongs to the top layer of the double-cantilever

    beam, the other part belongs to the bottom layer.

    The tip displacement u is plotted against the applied load in Figure 11. When the appliedload is equal to F   ≈  2.4 N, a crack nucleates at the notch in the top layer. A new cohesivesegment is added to the finite element model as shown in Figure 12 (a). Upon further loading,

    the crack propagates towards the interface, Figure 12(b). At this point the top layer has

    completely debonded and the interface is now loaded in nearly pure mode-II. When the

  • 8/20/2019 Computational Aspects of Cohesive Zone Models

    12/18

    (a) (b)

    Figure 12: Position of cohesive segments during the simulation at crack initiation in the top

    layer (a) and after total failure of the top layer (b).

    Figure 13: Final deformation of the specimen (amplification factor 100.0).

    shear tractions in the interface exceed the decohesion strength, the two layers start to debond.

    Figure 13 shows the deformation of the specimen at the final stage of loading. The highlydistorted elements in Figure 13 contain the displacement jumps   v j   that govern the open

    cracks. The actual deformation of the material modelled by these elements is of the same

    order of magnitude as the deformation in the surrounding elements.

    Fast crack propagation

    Another application of the cohesive segment method is the simulation of fast crack growth.

    When the crack propagation speed approaches the Rayleigh wave speed of he material, the

    fracture process is characterised by intermittent crack propagation and micro-crack nucle-

    ation or branching in the vicinity of the main crack tip.

     Formulation

    In order to take into account the inertia effects in these analyses, the equilibrium equation in

    (9) needs to be extended as follows:

    ρü + ∇ · σσσ =  0 ,   (21)

    where ρ  is the density of the material and  ü denotes the acceleration of a material which isthe second derivative of the displacement field in eq. (6) with respect to time:

    ü = ∂ 2u

    ∂t2  = ¨̄u +

    m

     j=1

    HΓd,j¨̃u j .   (22)

  • 8/20/2019 Computational Aspects of Cohesive Zone Models

    13/18

    Following the same procedure as before, we arrive at the following system of discrete equi-

    librium equations:

    Māā   Māã1   . . .   Māãm

    Mã1ā   Mã1ã1   . . .   Mã1ãm

    ...  ...

      . . .  ...

    Mãmā   Mã1ãm   . . .   Mãmãm

    ¨̄a

    ¨̃a1

    ...

    ¨̃am

    =

    f extā

    f extã1

    ...

    f extãm

    f intā

    f intã1

    ...

    f intãm

    .

    (23)

    The internal and external force vectors are defined in eqs. (15) and (16). The terms in the

    mass matrix M  are:

    Māā = Ω

    ρNTNdΩ ;   Māãj  = Ω+j

    ρNTNdΩ ;   Mãj ãk= 

    Ω+j ∩Ω+

    k

    ρNTNdΩ .   (24)

    Note that these matrices are symmetric so thatMāãj  =  Mãj ā. The equilibrium equations canbe discretised in the time domain with a variant of the Newmark-β  explicit time integrationscheme, as for example in [30]:

    ḋt+1

    2∆t =   ḋt +

     1

    2∆t d̈t ,

    dt+∆t =   dt + ∆t  ḋt+

    1

    2∆t ,

    d̈t+∆t =  M−

    1(f ext,t+∆t − f int,t+∆t) ,   (25)

    ḋt+∆t =   ḋt+

    1

    2∆t +

     1

    2∆t d̈t+∆t ,

    where ∆t is the discrete time step and  d represents the total array of displacement compo-nents:  d = [ā, ã1, . . . , ãm]

    T.

     Implementation aspects

    Traditionally, the calculation time of an explicit time integration procedure is reduced by

    using a diagonal (lumped) mass matrix in eq. (25). Numerical experiments show that in

    the case of the cohesive segments method, such an approach leads to inaccurate results. Bylumping the mass matrix, essential information on the coupling of the regular and additional

    degrees of freedom is lost, which gives rise to the spurious transmission of stress waves

    through cohesive surfaces. In the example presented here, the consistent mass matrix is

    used and the system of equations is solved using a direct solver. Although this strategy is

    less efficient, the additional computational effort is relatively small. The most expensive

    operation in this approach, the factorisation of the mass matrix, is only performed when this

    matrix is modified, e.g. when an existing segment is extended or a new segment is created.

    Considering the use of small time increments, this will only occur occasionally.

    The critical time increment for explicit iteration procedures is related to the length of the

    smallest element in the domain and the elastic properties of the material. As a rule of thumb,

  • 8/20/2019 Computational Aspects of Cohesive Zone Models

    14/18

    hle

    le

    Figure 14: Overview of the algorithm that prevents a discontinuity to cross an element

    boundary in the vicinity of a supporting node. The gray dashed line denotes the originalposition of the discontinuity. Since it crosses the node within a distance  hle (denoted by thedashed circle), it is moved away from the node.

    the following relation is used for a stable time increment:

    ∆t =  α∆tcrit = αlecd

    ,   (26)

    where le is the dimension of the smallest element in the domain,  cd is the dilational speed of the material and α is a reduction factor that accounts for destabilising effects due to nonlinearmaterial behaviour. In high-speed fracture mechanics, this factor is normally taken to be 0.1.

    In the case of the cohesive segments method however, it appears that when an element is

    crossed by a discontinuity, the two separated parts can be considered as individual elements,

    each with a smaller effective length than the original element. In principle, an element can be

    crossed by a discontinuity in such a way that, according to relation (26), the time increment

    for a stable solution procedure will become infinitesimal. In order to avoid such a situation,

    a discontinuity is not allowed to cross an element boundary within a certain distance hle   toa node, where h  is a factor between 0  and 0.5, see Figure 14. A side effect of this additionalcriterion is that the exact orientation of a new segment is slightly modified. However, it can

    be demonstrated that for reasonably small values of  h  (in the example, we assume h  = 0.1)

    the deflection is in the order of a few degrees. In the current model, we use the followingrelation to obtain the stable time increment:

    ∆t =  αhlecd

    .   (27)

     Example: dynamic shear failure

    An experimental technique for subjecting edge cracks in plate specimens to high rate shear

    loading was proposed by Kalthoff and Winkler [31] and related configurations have been

    used subsequently in a variety of investigations, e.g. Mason et al. [32] and Ravichandar [33].

  • 8/20/2019 Computational Aspects of Cohesive Zone Models

    15/18

    2L

    v0

    a

    2W 

    Figure 15: Geometry and loading conditions of the specimen for the dynamic shear failure

    test.

    At sufficiently high rates of loading, fracture in this configuration takes place by cleavage

    cracking at an angle of approximately 60◦ to 70◦ with respect to the initial crack. This typeof failure has been modelled by Needleman and Tvergaard [34] and Li et al. [35]. Here,

    we explore the use of the cohesive segments method to analyse the configuration shown in

    Figure 15.

    Various calculations were carried out. In some calculations numerical instabilities were

    encountered, while in other calculations there were difficulties associated with resolving near

    crack tip fields with the mesh used. These issues are being addressed as the development of 

    the cohesive segments method for dynamics problems proceeds. Nevertheless, at this early

    stage, good results have been obtained for a specimen that has dimensions  L   = 0.003 mand W   = 0.0015 m and has an initial crack with length  a  = 0.0015 mm. It is made of anisotropic linear elastic material with Young’s modulus 3.24 · 109 N/m2, Poisson’s ratio 0.35and density  ρ   = 1190.0 kg/m3. The corresponding dilatational, shear and Rayleigh wavespeeds are cd = 2090 m/s, cs = 1004 m/s and cR = 938 m/s. The ultimate normal traction of the material is equal to 100.0 · 106 N/m2 and the fracture toughness is  700 N/m. The lowerpart of the specimen is subjected to an impulse load in positivex direction which is modelledas a prescribed velocity with magnitude v0 = 10 m/s and a rise time tr  = 0.1 µs.

    The specimen is modelled with linear quadrilateral elements. In the region around the

    crack tip, the mesh is locally refined and the length of the elements is  le   = 15.0 µm. Ac-cording to equation (27), the time increment is set to  ∆t  = 1.0 · 10−10 s. At the start of thesimulation, the model has 11587 degrees of freedom. The initial crack is modelled as a trac-

    tion free discontinuity and therefore, in contrast with the simulations in [34], the crack tip is

    sharp. Possible crack nucleation away from the main crack tip is not taken into account.

    As can be seen in Figure 16, the crack propagates roughly at an angle around  70◦, whichis in agreement with experimental observations [31]. The small fluctuations in the path are

    most likely caused by stress waves reflections.

    Concluding remarks

    Cohesive–zone models constitute a powerful approach to analyse fracture, in particular for

  • 8/20/2019 Computational Aspects of Cohesive Zone Models

    16/18

    70◦

    60◦

    Figure 16: Position of crack (bold line) at  t  = 5.0 ms. The dashed lines are the projectionsof straight cracks at angles of  60◦ and 70◦.

    heterogeneous materials and for dynamic fracture. The bulk and cohesive constitutive re-

    lations together with appropriate balance laws and boundary (and initial) conditions com-

    pletely specify the problem. Fracture, if it takes place, emerges as a natural outcome of the

    deformation process.The partition-of-unity property of finite element shape functions enables a natural imple-

    mentation of cohesive–zone models, unbiased by the initial mesh design. The formulation

    can be used in a wide variety of numerical applications. In particular, cohesive surfaces of 

    a finite size can be defined, which can be placed at arbitrary locations and in arbitrary di-

    rections. This enables diffuse crack patterns to be modelled. On the other hand, a single,

    continuous crack can also be represented, thus enabling the simulation of the initiation and

    growth of a dominant crack without ad hoc assumptions. Nevertheless, in dynamic simula-

    tions, some problems related to the numerical stability of the time-integration technique and

    the resolution of the near tip fields were encountered. These issues will be addressed as the

    development of the method proceeds.

    References

    [1] Barenblatt, G.I., The mathematical theory of equilibrium cracks in brittle fracture, Ad-

    vances in Applied Mechanics,  7, 55–129, 1962.

    [2] Dugdale, D.S., Yielding of steel sheets containing slits, Journal of the Mechanics and 

    Physics of Solids,  8, 100–108, 1960.

  • 8/20/2019 Computational Aspects of Cohesive Zone Models

    17/18

    [3] Hillerborg, A., Modeer, M. and Petersson, P.E., Analysis of crack formation and crack 

    growth in concrete by means of fracture mechanics and finite elements,  Cement and 

    Concrete Research,  6, 773–782, 1976.

    [4] Needleman, A., A continuum model for void nucleation by inclusion debonding, Jour-

    nal of Applied Mechanics,  54, 525–531, 1987.

    [5] Tvergaard, V. and Hutchinson, J.W., The relation between crack growth resistance

    and fracture process parameters in elastic-plastic solids, Journal of the Mechanics and 

    Physics of Solids,  40, 1377–1397, 1992.

    [6] Reinhardt, H.W. and Cornelissen, H.A.W., Post-peak cyclic behaviour of concrete in

    uniaxial and alternating tensile and compressive loading,   Cement and Concrete Re-

    search,  14, 263–270, 1984.

    [7] Hutchinson, J.W. and Evans, A.G., Mechanics of materials: top-down approaches to

    fracture, Acta Materialia,  48, 125–135, 2000.

    [8] Rots, J.G., Strain-softening analysis of concrete fracture specimens. In Fracture Tough-

    ness and Fracture Energy of Concrete, edited by F.H. Wittmann, Elsevier Science Pub-

    lishers, Amsterdam, 1986, 137–148.

    [9] Chandra, N., Li, H., Shet, C. and Ghonem, H., Some issues in the application of co-

    hesive zone models for metal-ceramic interfaces,  International Journal of Solids and 

    Structures, 39, 2827–2855, 2002.

    [10] Schellekens, J.C.J., de Borst, R., On the numerical integration of interface elements,

     International Journal for Numerical Methods in Engineering,  36, 43–66, 1992.

    [11] Schellekens, J.C.J. and de Borst, R., A non-linear finite element approach for the anal-

    ysis of mode-I free edge delamination in composites,  International Journal of Solids

    and Structures,  30, 1239–1253, 1993.

    [12] Allix, O. and Ladevèze, P., Interlaminar interface modelling for the prediction of de-

    lamination. Composite Structures,  22, 235–242, 1992.

    [13] Xu, X.P. and Needleman, A., Numerical simulations of fast crack growth in brittle

    solids, Journal of the Mechanics and Physics of Solids,  42, 1397–1434, 1994.

    [14] Camacho, G.T. and Ortiz, M., Computational modeling of impact damage in brittle

    materials, International Journal of Solids and Structures,  33, 2899–2938, 1996.

    [15] Tijssens, M.G.A., Sluys, L.J. and van der Giessen, E., Numerical simulation of quasi-

    brittle fracture using damaging cohesive surfaces,   European Journal of Mechanics:

     A/Solids,  19, 761–779, 2000.

    [16] de Borst, R., Sluys, L.J., Mühlhaus, H.-B., Pamin, J., Fundamental issues in finite

    element analysis of localisation of deformation,  Engineering Computations,   10, 99–122, 1993.

    [17] van Mier, J.G.M., Fracture Processes of Concrete, CRC Press, Boca Raton, 1997.

    [18] Babuska, I. and Melenk, J.M., The partition of unity method, International Journal for 

     Numerical Methods in Engineering,  40, 727–758, 1997.

    [19] Belytschko, T. and Black, T., Elastic crack growth in finite elements with minimal

    remeshing, International Journal for Numerical Methods in Engineering, 45, 601–620,

    1999.

    [20] Moës, N., Dolbow, J. and Belytschko, T., A finite element method for crack growth

    without remeshing,  International Journal for Numerical Methods in Engineering,  46,

    131–150, 1999.

  • 8/20/2019 Computational Aspects of Cohesive Zone Models

    18/18

    [21] Wells, G.N. and Sluys, L.J., A new method for modeling cohesive cracks using finite el-

    ements, International Journal for Numerical Methods in Engineering,  50, 2667–2682,

    2001.

    [22] Wells, G.N., de Borst, R. and Sluys, L.J., A consistent geometrically non–linear ap-

    proach for delamination, International Journal for Numerical Methods in Engineering,

    54, 1333–1355, 2002.

    [23] Moës, N. and Belytschko, T., Extended finite element method for cohesive crack 

    growth, Engineering Fracture Mechanics,  69, 813–833, 2002.

    [24] Remmers, J.J.C., de Borst, R., Needleman, A., A cohesive segments method for the

    simulation of crack growth,  Computational Mechanics,  31, 69–77, 2003.

    [25] Allix, O. and Corigliano, A., Geometrical and interfacial non-linearities in the analy-

    sis of delamination in composites,  International Journal of Solids and Structures,  36,

    2189–2216, 1999.

    [26] Wells G.N., Remmers J.J.C., Sluys L.J. and de Borst R., A large strain discontinuous

    finite element approach to laminated composites, IUTAM symposium on computational

    mechanics of solid materials at large strains, 355–364, 2003.

    [27] Remmers, J.J.C., Wells, G.N. and de Borst, R., A solid–like shell element allowing for

    arbitrary delaminations, International Journal for Numerical Methods in Engineering,

    58, 2013–2040, 2003.

    [28] Daux, C., Moës, N., Dolbow, J., Sukumar, N. and Belytschko, T., Arbitrary branched

    and intersecting cracks with the extended finite element method, International Journal

     for Numerical Methods in Engineering,  48, 1741–1760, 2000.

    [29] Xu, X.P. and Needleman, A., Void nucleation by inclusion debonding in a crystal ma-

    trix,  Modelling and Simulation in Materials Science and Engineering ,   1, 111–132,

    1993.[30] Belytschko T., Chiapetta R.L. and Bartel H.D., Efficient large scale non-linear transient

    analysis by finite elements, International Journal for Numerical Methods in Engineer-

    ing,  10, 579–596, 1976.

    [31] Kalthoff J.K. and Winkler S., Failure mode transition at high rates of loading. Proceed-

    ings of the International Conference on Impact Loading and Dynamic Behaviour of 

     Materials, Chiem C.Y., Kunze H.D. and Meyer L.W. eds., 43-56, Deutsche Gesellshaft

    für Metallkunde, 1988.

    [32] Mason J.J., Rosakis A.J. and Ravichandran G., Full field measurements of the dynamic

    deformation field around a growing adiabatic shear band at the tip of a dynamically

    loaded crack or notch.  Journal of the Mechanics and Physics of Solids,  42, 1679-1697,1994.

    [33] Ravichandar K., On the failure mode transitions in polycarbonate under dynamic

    mixed-mode loading.  International Journal of Solids and Structures,   32, 925–938,

    1995

    [34] Needleman A. and Tvergaard V., Analysis of a brittle-ductile transition under dynamic

    shear loading International Journal of Solids and Structures,  32, 2571–2590, 1995.

    [35] Li S.F., Liu W.K., Rosakis A.J., Belytschko T. and Hao W., Mesh-free Galerkin sim-

    ulations of dynamic shear band propagation and failure mode transition.   International

     Journal of Solids and Structures,  39, 1213–1240, 2002.