Chandra 02 Some Issues of Cohesive zone

download Chandra 02 Some Issues of Cohesive zone

of 13

Transcript of Chandra 02 Some Issues of Cohesive zone

  • 8/21/2019 Chandra 02 Some Issues of Cohesive zone

    1/29

    Some issues in the application of cohesive zone models

    for metal–ceramic interfaces

    N. Chandra   a,*, H. Li   a, C. Shet   a, H. Ghonem   b

    a Department of Mechanical Engineering, College of Engineering, Florida State University, Tallahassee, FL 32310, USAb Department of Mechanical Engineering, University of Rhode Island, Kingston, Rhode Island, USA

    Received 29 November 2000

    Abstract

    Cohesive zone models (CZMs) are being increasingly used to simulate discrete fracture processes in a number of 

    homogeneous and inhomogeneous material systems. The models are typically expressed as a function of normal and

    tangential tractions in terms of separation distances. The forms of the functions and parameters vary from model to

    model. In this work, two different forms of CZMs (exponential and bilinear) are used to evaluate the response of in-

    terfaces in titanium matrix composites reinforced by silicon carbide (SCS-6) fibers. The computational results are then

    compared to thin slice push-out experimental data. It is observed that the bilinear CZM reproduces the macroscopic

    mechanical response and the failure process while the exponential form fails to do so. From the numerical simulations,

    the parameters that describe the bilinear CZM are determined. The sensitivity of the various cohesive zone parameters

    in predicting the overall interfacial mechanical response (as observed in the thin-slice push out test) is carefully ex-amined. Many researchers have suggested that two independent parameters (the cohesive energy, and either of the

    cohesive strength or the separation displacement) are sufficient to model cohesive zones implying that the  form (shape)

    of the traction–separation equations is unimportant. However, it is shown in this work that in addition to the two

    independent parameters, the form of the traction–separation equations for CZMs plays a very critical role in deter-

    mining the macroscopic mechanical response of the composite system.    2002 Elsevier Science Ltd. All rights reserved.

    Keywords: Cohesive zone models; Metal–ceramic interface; Push-out test; Thermal residual stress; Interface failure

    1. Introduction

    Metallic and intermetallic matrix composites (MMCs and IMCs) are being considered as potential

    material systems of choice for advanced propulsion systems of future air and space craft because of their

    high specific stiffness and strength. A critical issue in the successful application of these composites is the

    mechanical characteristics of fiber–matrix interfaces. Interfaces are narrow regions separating well-defined

    domains and are primarily responsible for a range of key properties including stiffness, strength, and

    fracture behavior (Jayaraman et al., 1993). The role of interfaces is very vital to the stress transfer between

    International Journal of Solids and Structures 39 (2002) 2827–2855

    www.elsevier.com/locate/ijsolstr

    * Corresponding author. Fax: +1-850-410-6420.

    E-mail address:  [email protected] (N. Chandra).

    0020-7683/02/$ - see front matter 

     2002 Elsevier Science Ltd. All rights reserved.PII: S0 0 2 0 -7 6 8 3 (0 2 )0 0 1 4 9 -X

  • 8/21/2019 Chandra 02 Some Issues of Cohesive zone

    2/29

    the fiber and matrix and interfaces influence the mechanical performance and fracture behavior of com-

    posites under various loading conditions (Kim and Mai, 1991a,b).

    Interfaces have been modeled in a number of ways, e.g., a narrow region of continuum with graded

    properties, an infinitely thin surface separated by springs, and cohesive zones with specific traction–sepa-

    ration relations. One of the key research issues is to determine the best way to characterize interfaces within

    the framework of continuum mechanics rather than using ad hoc methods just to facilitate numerical

    implementations, e.g. springs in finite element methods. In the past few years, Chandra and co-workers

    (Ananth and Chandra, 1995a,b; Chandra and Ananth, 1995; Mukherjee et al., 1997, 1998; Chandra and

    Ghonem, 2001) simulated the interfacial mechanical behavior of composites in thin-slice push-out tests

    incorporating the spring layer model and fracture mechanics approach. They analyzed both the initiation

    and the propagation of interfacial failure to explain many of the experimental observations of fiber push-

    out behavior at room and elevated temperatures. In the spring layer model (Ananth and Chandra, 1995b), a

    stress based criterion for debonding and a frictional resistance based criterion for interfacial sliding have

    been used to capture debonding and sliding. Debonding is postulated to occur under the combined action

    of normal tensile stress (mode I) and shear stress (mode II) at the interface. In the fracture mechanics

    approach strain energy release rate at the tip of the interface cracks is computed using the equivalentdomain integral method (Mukherjee et al., 1997). In this approach a critical strain energy release rate is

    used as the criterion to determine the propagation of the crack.

    Recently the cohesive zone approach is being increasingly used in describing fracture and failure be-

    havior in a number of material systems. CZM has been used in the past to study crack tip plasticity and

    creep under static and fatigue loading conditions, crazing in polymers, adhesively bonded joints, interface

    cracks in bimaterials, and crack bridging due to fibers and ductile particles in composites. Interface mod-

    eling using CZM has the distinct advantage compared to other global approaches (e.g. shear lag model), in

    that it is based on a micro-mechanical approach. CZM was originally proposed by Barenblatt (1959, 1962)

    as a possible alternative to the concept of fracture mechanics in perfectly brittle materials. Later, Dugdale

    (1960) extended this concept to perfectly plastic materials by postulating the existence of a process zone at

    the crack tip. Based on the Barenblatt and Dugdale models, Hillerborg and co-authors (van Mier, 1996)proposed the fictitious crack model (FCM) for analyzing crack growth in cementitious composites. FCM

    has made rapid strides in the analysis of concrete structures in the field of civil engineering.

    CZM has spawned a plethora of models in fracture of metals, ceramics, and polymers and their com-

    posites (Needleman, 1990a,b; Rice and Wang, 1989; Wappling et al., 1998). It is not the purpose of this

    paper to review all of those models and applications, but to outline some of the key works relevant to the

    present study. Needleman was one of the first to use polynomial and exponential types of traction–sepa-

    ration equations to simulate the particle debonding in metal matrices (Needleman, 1987, 1990a, 1990b,

    1997). Xu and Needleman (1993, 1994, 1995) further used the above models to study the void nucleation at

    the interface between particle and matrix, fast crack growth in brittle materials under dynamic loading, and

    dynamic crack growth at the interface of bimaterials. Tvergaard (1990) used a quadratic traction–dis-

    placement jump form to analyze interfaces. Tvergaard and Hutchinson (1992) used a trapezoidal shape of 

    the traction–separation model to calculate the crack growth resistance. Camacho and Ortiz (1996) em-

    ployed a linear traction–separation equation with an additional fracture criterion to propagate multiple

    cracks along arbitrary paths during impact damage in brittle materials. Geubelle and Baylor (1998) have

    utilized a bilinear CZM to simulate the spontaneous initiation and propagation of transverse matrix cracks

    and delamination fronts in thin composite plates subjected to low-velocity impact. In all the CZMs (except

    Dugdale’s model and Camacho et al.’s model; see Table 1), the traction–separation relations for the in-

    terfaces are such that with increasing interfacial separation, the traction across the interface reaches a

    maximum, then decreases and eventually vanishes permitting a complete decohesion. The main difference

    lies in the shape and the constants that describe that shape. In this context, the term ‘shape’ is used in a

    loose sense to describe the normal (or tangential) traction vs. the normal (or tangential) separation response

    2828   N. Chandra et al. / International Journal of Solids and Structures 39 (2002) 2827–2855

  • 8/21/2019 Chandra 02 Some Issues of Cohesive zone

    3/29

    Table 1

    Various cohesive zone models and their parameters

    Author

    (year)

    Proposed model Model parameters Problem

    solved

    Model c

    Barenblatt

    (1959, 1962) K  ¼ R d þd 0

    0

    G 1ðt Þ dt  ffit 

    ¼ ffiffiffiffiffiffiffi

    p ET 1m2

    q   (ductile)

    ¼ ffiffiffiffiffiffiffi

    p ET 01m2q 

      (brittle)

    T  ¼  T 0 þT 1T 0  is work of separation

    for brittle materials

    T 1   is work of plastic deformation

    Perfectly

    brittle

    materials

    Dugdale

    (1960) sa ¼ 2sin2   p

    4T ry

    For small value of  T =ry

     s1 ¼ 1:23   T 

    ry

    2Yielding of 

    thin ideal

    elastic–plastic

    steel sheets

    containing slits

    Plastic z

    ranges f

    0.042 to

    (in.)

    Needleman

    (1987)

    /sep  is work of separation

    d  are normalizing parametersrmax   is cohesive strength

    Particle-matrix

    decohesion

    d ¼ 109

    108

    m cenergy 1

    rmax ¼ 11

    ry ¼ 350

    Rice and

    Wang

    (1989)

     E 0   is initial Young’s modulus

    h  is normalizing parameter

    rmax   is maximum stress

    a   is constant   E 0ha2 ¼ 2c

    Solute

    segregation

    Needleman

    (1990a)

    /sep  is work of separation

    d  are normalizing parameters

    rmax   is cohesive strength

    Particle-matrix

    decohesion

    d ¼ 109108

  • 8/21/2019 Chandra 02 Some Issues of Cohesive zone

    4/29

    Table 1 (continued )

    Author

    (year)

    Proposed model Model parameters Problem

    solved

    Model c

    Needleman

    (1990b)

    /n,  / t  are work of normal

    and shear separation

    dn,  d t   are critical displacements

    rmax   is cohesive strength

    Decohesion of 

    interface under

    hydrostatic

    tension

    dn ¼ dt ¼2 102 10

     J IC=/n ¼0:57 – 2:

    rmax=r0 

    Tvergaard

    (1990)

    dn,  d t   are critical displacements

    rmax   is cohesive strength

    Interfaces of 

    whisker rein-

    forced metal

    matrix com-

    posites

    dn ¼ dt ¼1 10

     E  ¼ 60 ðYoungry= E  ¼ 0rmax=ry 

    Tvergaardand Hutch-

    inson (1992)

    C0   is work of separationdc   is critical displacement

    rmax   is peak normal traction

    =interface strength d1,d2  are factors governing shape

    Crack growthin elasto-plas-

    tic material,

    peeling of ad-

    hesive joints

    Css=C0 ¼ðCss ¼   Pdcn=d

    ct ¼  

    rmax=ry d2 ¼ 0:1ry= E  ¼ 

    Xu and

    Needleman

    (1993)

    /n  is work of normal separation

    /t  is work of shear separation

    dn,  d t   are critical displacements

    rmax   is cohesive strength

    Particle-matrix

    decohesion

    dn ¼ dt ¼2  102  10

  • 8/21/2019 Chandra 02 Some Issues of Cohesive zone

    5/29

    Camacho

    and Ortiz

    (1996)

    r0,  s0   are normal and shear stress

    at fracture initiation

    drcr; dscr   are critical normal and

    shear opening displacement

    G C ¼   fracture energy

    Impact Alumin

    r0 ¼ 40drcr ¼ 1

    107 msteel:

    r0 ¼ 15drcr ¼ 5

    Geubelle

    and Baylor

    (1998)

    /n,  / t  are work of normal and

    tangential separation

    Dn,  D t   are normal and tangential

    displacement jump

    dn,  d t   are normal and tangential

    interface characteristic length

    Delamination

    by low-velocity

    impact

    rmax ¼ E E =10 crnormal

    ment ju

    Dcn ¼ 10

  • 8/21/2019 Chandra 02 Some Issues of Cohesive zone

    6/29

    as typically used in the literature. A more precise concept would be to examine the functional form of the

    relation between the traction vector and displacement vector that exists between the separating surfaces

    during the fracture process. However, for the sake of simplicity the term ‘shape’ will be used in this work.

    The magnitude of the parameters in CZMs vary widely, ranging from MPa to GPa for traction, J to kJ

    for energy, and nanometers to micrometers for separation distance. Table 1 describes the details of some

    popular CZMs mentioned above with specific attention focused on the shape and the values of the model

    parameters.

    The basic question raised in this paper is whether the detailed form, i.e., the shape of traction–separation

    equations, affects the macroscopic mechanical response of a material system. We attempt to answer this

    question by comparing the predictions of two different models to the experimental results of thin slice push

    out tests. In this research work, the metal–ceramic interfaces in Titanium Beta 21S/SCS composites have

    been studied by using exponential and bilinear forms of CZMs. This paper is arranged as follows: in Section

    2, formulation of the interface problem is presented; in Section 3, critical issues of various CZMs are

    discussed; in Section 4, two CZMs used in the present work are outlined; following Section 4, the finite

    element approach is given in Section 5; finally, the detailed results are discussed in Section 6, which includes

    comparison between experimental observations and computational results and comparison between sim-ulation results from two different models. The effects of various parameters that describe CZMs are in-

    vestigated systematically to examine the sensitivity of the interfacial parameters of CZM. Currently, many

    researchers (e.g. Rahulkumar et al. (2000), Mohammed and Liechti (2000), Hutchinson and Evans (2000))

    have assumed that two independent parameters (cohesive energy, and either of the cohesive strength or the

    separation displacement) are sufficient to model interfaces using CZM. Series of parametric studies indicate

    that this assumption leads to meaningless responses (see Figs. 10 and 12). The responses clearly show (by

    the theory of contradiction) that those two parameters are not sufficient to represent the physics of the

    problem of the interface separation process. In order to accurately simulate the interface and reproduce the

    macroscopic mechanical behavior of composites, CZM should include, apart from those two parameters,

    the shape of traction–separation law.

    2. Interface problem

    In general, continuum models of materials are stated as boundary value problems in which we seek

    quantities such as displacements, velocities, stresses or temperatures at each point in a given domain. The

    governing equations of the continuum are identified with the balance laws of classical physics such as

    conservation of energy, mass and momentum. The geometries of deformation are prescribed in term of a

    kinematic relationship between strain and displacements or deformation gradients. The specific choice of 

    the relationship depends on whether the body is undergoing finite deformation or not. To complete the

    description of the boundary value problem a constitutive relationship between the kinematic quantity

    (strain, deformation gradient) and the kinetic quantity (stress, stress rate) is postulated. This constitutive

    equation is specific to the body under consideration within the range of temperature, rate of loading, type of 

    loading and environment and is usually obtained phenomenologically.

    If the boundary value problem consists of domains where one set of equations cannot be applied, then

    different sets are used to represent the individual domains. A case in point is the modeling of heterogeneous

    composites in which different forms of constitutive relations are used. For such bodies, there is a distinct

    separation between the various regions where the usual theories of continuous media are valid. Such a

    separation leads to discontinuity in at least a few of the field quantities, e.g. strains. In the sense of con-

    tinuum mechanics, this separation defines the interface that needs to be independently characterized. In

    some cases, interfaces are not modeled explicitly and the two regions are rigidly bonded leading to con-

    tinuous displacement fields with discontinuous strain and stress fields across the interface.

    2832   N. Chandra et al. / International Journal of Solids and Structures 39 (2002) 2827–2855

  • 8/21/2019 Chandra 02 Some Issues of Cohesive zone

    7/29

    Thus, the thermo-mechanical characterization of interfaces should be explicitly included for determining

    the response of bodies represented by regions of different materials. This prescription is even more critical if 

    the interfaces were to physically separate (open up or slide) during the loading or unloading process.

     2.1. Generic formulation

    Consider the two solid bodies  X1   and  X2   separated by a common boundary  S  as shown in Fig. 1(a),

    where  S  can be considered as the same surface  S 1 2 X1   and  S 2 2 X2, in the initial configuration, i.e.  S 1 ¼S 2 ¼ S . Mathematically, we would like to define  S  as an infinitesimally thin 3-D domain with surfaces  S 1and  S 2   being the part of  X1   and  X2  respectively, before separation occurs. For all practical purposes the

    surface S 1  or  S 2  can be identified as a single surface as a part of either of the domains. A material particle

    initially located (within either of the domains  X1   and  X2) at some position  X , moves to a new location   x x,

    with a one to one correspondence between    x x   and   X   given by the equation of motion    x x ¼ vð X ; t Þ   or xi ¼ vð X  j; t Þ.

    In a generic sense,  S  defines the interface between any two domains. If  X1   is a metal and  X2  a ceramic,

    then  S  represents a metal–ceramic interface; if  X1  and  X2  belong to the same material depicting grains of 

    different orientations then S  is a grain boundary, and if  X1  and  X2  represent the same domain X1 [X2 ¼ X,then  S  is an internal surface not yet separated.

    In any one of those cases, if  S  separates (fractures) to and as shown in Fig. 1(b), then the process

    creates a new internal/external surface violating the fundamental laws of continuity. Obviously the newly

    formed region cannot be uniquely mapped from the undeformed configuration. The equation of motion of 

    Fig. 1. Conceptual frame work of cohesive zone model for interface.

    N. Chandra et al. / International Journal of Solids and Structures 39 (2002) 2827–2855   2833

  • 8/21/2019 Chandra 02 Some Issues of Cohesive zone

    8/29

    the body xi ¼ vð X  j; t Þ cannot identify the new region. This is the fundamental problem in modeling fracture(creation of internal/external surface) in the framework of the mechanics of continuous media.

    The surface S  represented by the unit normal N ð N 1 2 S 1, and N 2 2 S 2 and  N  ¼  N 1 ¼ N 2) acting along theboundary separating the domain prior to deformation is as shown in Fig. 1(a). In the deformed configu-

    ration as shown in Fig. 1(c)   n̂n1  and   n̂n2  represent the unit normal of the surfaces (separated or otherwise).

     2.2. Domains  X1   and  X 2

    Consider two neighboring particles located at  X   and X  þ d X   in the initial configuration. Let d X  deformto d x x   such that

    d x x ¼ e F  F  d X ;   ð1Þwhere e F  F   is the deformation gradient evaluated with respect to the material coordinates. When constitutiveequations of materials are formulated in the rate form (e.g. elastoplasticity, viscoplasticity) then it is

    convenient to define velocity gradient e L L   such thate L L ¼  ovvo x x

    ;   ð2Þ

    where the gradient of velocity vector vv is evaluated with respect to spatial coordinates. e L L is decomposed intoa symmetric part e D D   and antisymmetric part eW  W    such thate L L ¼ e D D þ eW  W   ;   ð3Þwheree D D ¼   1

    2ðe L L þ e L LT Þ   and   eW  W   ¼   1

    2ðe L L  e L LT Þ;   ð4Þ

    where

     e D D  is the rate deformation tensor and

     eW  W    is the spin tensor.

    In order to formulate an appropriate constitutive relation to describe the thermo-mechanical behavior of 

    material, balance laws (conservation of mass, linear and angular momentum) and thermodynamic laws(first and second law) should be satisfied. Additionally, the equations should remain invariant under a

    change of frame of reference. This condition can be satisfied by selecting quantities that are objective (frame

    invariant) in nature.

    If the material is governed by Cauchy elasticity, which states that the current stress depends only on the

    current deformation, then the Cauchy’s stress can be written in terms of right stretch tensor eU U , such that~rr ¼ e R RG ð eU U Þe R RT ;   ð5Þ

    where   G   is the material response function and deformation gradient e F  F  ¼ e R R eU U   and e R R   is the orthogonalrotational tensor. However, if the material is hyperelastic (possesses a strain energy density function   W ,

    that depends on the current configuration), then

    W   ¼ W  ðe E  E Þ   and eS S  ¼  oW  oe E  E  ;   ð6Þ

    where eS S  is second Piola–Kirchoff stress tensor, and e E  E  is the Green–Lagrangian strain tensor. If the strainenergy density is  W   ¼ ð1=2Þ E ijC ijkl E kl, theneS S  ¼ eC C e E  E ;   ð7Þwhere eC C   is the fourth order elasticity tensor. In the finite deformation inelastic models, the constitutiveequations are formulated in the rate form based on hypoelasticity. In this case, it is assumed that the

    material has no memory and that the Cauchy stress rate D~rr=Dt  only depends on the current state   ~rr  andvelocity gradient

     e L L. Thus

    2834   N. Chandra et al. / International Journal of Solids and Structures 39 (2002) 2827–2855

  • 8/21/2019 Chandra 02 Some Issues of Cohesive zone

    9/29

    D~rr

    Dt  ¼  g ð~rr; e L LÞ:   ð8Þ

    In order to fulfill the objectivity principle (principle of the material-frame indifference), the above equation

    can be written as

    r ¼ hð~rr; e D DÞ;   ð9Þ

    where  r

    is the objective Jaumann rate of Cauchy stress tensor, given by

    r ¼ D~rr

    Dt   eW  W   ~rr  ~rr eW  W    T :   ð10Þ

    It is noted that D~rr=Dt  is the material derivative of the Cauchy stress. If  r

    is a linear function of the rate of 

    deformation, then

    r ¼

     eC C ð~rrÞ

    e D D:   ð11Þ

    The above hypo-elastic formulation can be extended to inelastic material by adopting an additive de-

    composition of the rate of deformation tensore D D ¼ e D DEl þ e D DIn;   ð12Þwhere e D DEl and e D DIn are the elastic and inelastic parts of the rate of deformation tensor. Although the validityof the above equation has been questioned, for the present analysis the formulation is acceptable. Addi-

    tionally, the equation has been implemented in most of the commercial codes including the one used in the

    present work. The constitutive equation can then be written as

    r ¼ eC C ðe D D  e D DInÞ:   ð13Þ

    The elasticity tensor

     eC C  is assumed to be isotropic. Since the rate of deformation tensor

     e D D  is objective, the

    above equation satisfies the principle of material frame indifference.

     2.3. Interface S 

    If  S  continues to be a part of  X1  and  X2  (having points/particles common to both), then the motion of  S 

    can be uniquely defined by the motion of either  X2 or  X1. Though the surface normal b N  N  would have rotatedand deformed to  n̂n, in the sense of the kinematics b N  N  ¼  Ke F  F  1n̂n, (K ¼ d s=dS  ¼  stretch ratio,  S  and  s  are thelengths of a small segment in the original and deformed configurations, respectively). Thus, we have a one-

    to-one relationship between the deformed and undeformed configurations.

    However, if  S  is to be separated as shown in Fig. 1(b), then we have created a pair of new surfaces in the

    traditional sense of the term. Consider the region bounded by and belonging to a new domain X.Assume that X  is a 3-D domain made of extremely soft glue, which can be shrunk to an infinitesimally thinsurface but can be expanded to a 3-D domain. The constitutive relation of  X   is expressed quite differentlyfrom that of a typical 3-D solid (e.g.  X1  or  X2). The two surfaces that are initially part of  X1  and  X2  (S 1  and

    S 2) have normal b N  N 1   and b N  N 2  in the undeformed configuration; b N  N 1   and b N  N 2  are equal and opposite. Duringdeformation the surfaces rotate to new normal  n̂n1  and  n̂n2. As the surfaces separate we have two surfaces

    (   2 X1 \X) and (   2  X2 \ X). The constitutive equation is written in terms of the normal andshear tractions and their corresponding separation displacements. For a narrow region (crack tip region)

    the directions of  n̂n1 and  n̂n2 are approximated to be same. A typical constitutive relation of  X is given by T  – dd

    relations (see Fig. 1(d)), such that

    if  j ddj 

  • 8/21/2019 Chandra 02 Some Issues of Cohesive zone

    10/29

    Beyond a separation distance of  jddj > jddsepj, the traction is identically zero within  X,jddjP jddsepj;   ~rrn̂n ¼ T  ¼  0:   ð15Þ

    It can also be construed that whenjddj >

     jdd

    sepjin the domain X, the stiffness C 

    ijkl  0. In order to implement

    the vectorial inequalities given in Eqs. (14) and (15) typically two separate identities are postulated for the

    normal and tangential components with limits set for each of them. Other methods of implementation are

    equally possible.

    In the numerical schemes, such as in finite element method, interfaces have been modeled as contact

    surfaces or surfaces separated by spring elements. Such approaches usually have ad hoc criteria for sep-

    aration and sliding and usually do not conserve energy. For example, when a nodal spring constant is

    reduced to zero upon satisfying a failure criterion, stored energy is not appropriately dissipated.

    In the cohesive zone approach, an independent specification of a local constitutive relation is established.

    This relationship reflects the coupling of stresses (or tractions) at the interfaces to the displacements they

    suffer. One class of interfacial constitutive model is based on the existence of a potential  /ðDÞ that measuresthe energy cost to displace the adjacent planes across an interface by a relative displacement D  (Needleman,

    1992). The resulting tractions are given by

    T  ¼  o/ðDÞoðDÞ :   ð16Þ

    3. Critical issues in the application of cohesive zone models

    Since CZMs provide a convenient way to model fracture (or for that matter any separation) process, it is

    instructive to critically review the available models. It is our opinion that at this stage of their development,

    CZMs should be truly called models rather than sanctifying them as  constitutive equations  or as   cohesive

    laws. We examine the most popular CZMs in terms of their forms, physical significance and applications.We examine not only the functional form of traction–separation distance relation but also the various

    parameters that describe such relations, e.g. the rising part of the curve, the peak values (both the maximum

    traction, rmax, and corresponding dmax), and the value of  dsep when complete separation occurs. Usually the

    T  – dd relation is expressed in the form of  T n  vs.  Dn=dc. The normalizing process ðDn=dcÞ is necessitated by therequirement that when traction is obtained as a function of cohesive strength, the multiplier has to be a

    non-dimensional quantity. An important issue is whether ‘dc’ is just a scaling parameter or if it has any

    other physical significance. It should be noted that  dc varies anywhere from 1010 to 105 m and rmax ranges

    from MPa to GPa. A final issue to be addressed is the physical significance of the area under the curve.

    Does it represent just the surface energy or all the dissipative work associated with the separation process?

    Barenblatt (1959, 1962) was the first to propose the concept of CZM while analyzing brittle fracture. He

    postulated that the molecular force of cohesion exists near the edge of a crack. The intensity of the mo-

    lecular force of cohesion ‘T n’ is found to vary as shown in row 1 of Table 1. Very interestingly, in this work

    Barenblatt did not suggest any specific form of  T  – dd relationship. He postulated the existence of the cohesive

    stress (thus reducing the singular stress). Crack propagation occurs when the integral effect of the cohesive

    force acting ahead of the crack tip (along the crack plane) for the length equal to the cohesive zone size ‘ d ’

    reaches a set value for the modulus of the cohesion, ‘ K ’. While investigating the yielding of an elastic–plastic

    steel sheet, Dugdale (1960) identified the existence of cohesive force (termed as internal stresses) acting at

    the tip of the crack for a length equal to plastic zone size.

    Needleman (1987) was one of the earliest to apply the traction–separation cohesive relationship to model

    void nucleation and debonding of inclusions in metallic materials. A number of potentials have been

    suggested in order to obtain different forms of response between   T   and   uu. For example, a polynomial

    2836   N. Chandra et al. / International Journal of Solids and Structures 39 (2002) 2827–2855

  • 8/21/2019 Chandra 02 Some Issues of Cohesive zone

    11/29

    function in 1987’s model (row 3, Table 1), an exponential type in 1990a’s model (row 5, Table 1), an ex-

    ponential function for the normal traction and periodic function for the tangential traction in 1990b’s

    model (row 6, Table 1) have all been proposed by Needleman and co-workers. In 1993, Needleman again

    used exponential forms for both normal and shear tractions (row 9, Table 1). While the polynomial form

    was chosen for analytical convenience, the exponential form was motivated by atomistic considerations

    following the work of Rose et al. (1981). The periodic response for shear was postulated to model the

    Pieriels shear stress associated with slips.

    The area under the normal traction vs. separation distance is assumed to be the work of separation  /sepand is in the range of 1 to 10 J/m2. In most of their (Needleman and co-investigators) work, the maximum

    stress   rmax   is typically assumed to be about three times that of yield stress (ry) of the matrix material.

    Typical values of  rmax ¼ 1000–1400 MPa, and dmax ¼ 109 to 108 m have been used in their analysis. In thecase of the exponential model, separation dsep theoretically occurs as d ! 1 . However, for practical reasonit can be assumed that complete separation occurs when  dsep   is about 6–7 times  dmax. For the polynomial

    model  dmax ¼ dsep=3.Rice and Wang (1989) analyzed embrittlement (or ductilization) effects of atomic scale separation of 

    solutes to grain boundaries. The normal traction vs. separation distance proposed in their work is shown inrow 4 of Table 1. The area under the curve was assumed to be

    G  ¼Z   1

    0

    rðdÞdd ¼ 2cint ¼ A;   ð17Þ

    where G  is the strain energy release rate in an elastic-brittle material. Rice raises a very important question

    regarding the area (A) under the curve. It can be interpreted as the adhesion energy or work of fracture, i.e.,

    the energy consumed in opening up two new surfaces from a single one. As a first approximation, for

    perfectly brittle materials the area under the curve can be assumed to be  G . However, for ductile materials

    we should recognize that if  J -integral represents the strain energy release rate  J  ¼ oU =o A, then J IC

     ¼ W  p

     þ2cint

     ¼ W  p

     þ A;

      ð18

    Þwhere W  p  is the plastic (or viscoplastic) work ahead of the crack tip. While in brittle inclusions (carbides in

    steel), W  p is about 1.3–2.5 times that of  A, in the failure of polycrystalline ductile materials  W  p is about 500– 

    1000 times that of  A.

    Tvergaard (1990) proposed a simplified interface model, where the traction shows a quadratic variation

    with respect to the traction separation, having a steep ascending curve and gradual softening curve as

    shown in row 7 of Table 1. Tvergaard and Hutchinson (1992) proposed a CZM as shown in row 8 of Table

    1, for simulating crack growth processes in elastic–plastic solids. The  T  – d   response is made up of linear

    segments with separate slopes for rising and falling parts with a horizontal segment at the traction value of 

    rmax. For mode I cracks, in elastic–plastic solids, Tvergaard and Hutchinson (1992) have noted that the

    shape parameters ðd1=dc; d2=dcÞ  are relatively unimportant and only  rmax  and the area under the curve  C0play a key role.

    Camacho and Ortiz (1996) used CZM concepts in their finite element implementation of ballistic impact

    problems. They departed from the earlier works and defined a separate multi-axial fracture criterion based

    on local normal and tangential stresses. Once this criterion is satisfied a cohesive relationship, as shown in

    row 10 of Table 1, is specified. The area under the curve is assumed to be the fracture energy, discounting

    the energy consumed prior to the stress at the tip of the crack reaching  rmax. The parameters used are

    dsep ¼ 5:43  lm,  rmax ¼ 1500 MPa,  G  ¼ 4074 J/m2 for steel (impactor), and  dsep ¼ 17  lm,  rmax ¼ 400 MPa,G  ¼ 34 J/m2 for alumina. This work deviates once again from the others discussed in that it attempts tosolve a macroscopic boundary value problem using CZM only to model the separation and not the initi-

    ation of cracks. Here G  inherently includes the surface energy and all other inelastic energies associated with

    the elastic–plastic processes.

    N. Chandra et al. / International Journal of Solids and Structures 39 (2002) 2827–2855   2837

  • 8/21/2019 Chandra 02 Some Issues of Cohesive zone

    12/29

    Geubelle and Baylor (1998) have simulated the delamination process in a composite plate subjected to

    low velocity impact using CZMs. They proposed bilinear models both for normal and tangential separation

    processes and we have followed this model in the present work. The area under the normal traction– 

    separation curve is considered as  G IC

     and that due to shear traction–separation is  G IIC

    . In applying this

    model to graphite/epoxy delamination the following values have been adapted: for PMMA,  rmax ¼ 324MPa,  G IC ¼ 352:3 J/m2,   dsep ¼ 2:17  lm; and for epoxy,   rmax ¼ 91 MPa,  G IC ¼ 88 J/m2,  G IIC ¼ 315 J/m2(interface),  G 

    plyIC ¼ 147 J/m2.

    A number of other investigators have followed the above models in solving a variety of boundary value

    problems. Rahulkumar et al. (2000) simulated large deformation peel tests of a fully relaxed polymer using

    the model shown in row 8 of Table 1. They use  /sep ¼ 5 – 10 J/m2 and  dsep ¼ 25–50  lm and have a corre-sponding rmax ¼ 0:147 MPa. Espinosa et al. (2000) simulated dynamic impact of woven fiber compositesusing models shown in row 7, 8 and 11 of Table 1. They use a value of  rmax ¼ 50 MPa,  /sep ¼ 200–2000 J/m2,  dsep ¼ 4–8  lm for the PMMA/epoxy interface. Foulk et al. (2000) modeled SCS-6/Timetal 21s metalmatrix composite interface using the model 8 in Table 1 with values of   rmax ¼  513–1500 MPa anddsep

     ¼ 0:1–0.2 lm. From the above discussion it is abundantly clear that there are many forms of cohesive

    zone models and the parameters in each of them vary by orders of magnitude.

    4. The cohesive zone models

    In this work, two CZMs, exponential (Xu and Needleman, 1993) and bilinear (Geubelle and Baylor,

    1998) forms are used to simulate the interfacial mechanical response by considering push-out test process as

    the example.

    For the exponential CZM, the interfacial potential is given by

    /ðDn;DtÞ ¼ /n þ /n expðDn=dnÞf½1 r þ Dn=dn½ð1 qÞ=ðr  1Þ

    ½q

    þ ½ðr qÞ=ð

    1ÞD

    n=d

    nexp

    ðD2

    n=d

    2

    t Þg;   ð19

    Þwith q ¼ /t=/n,  r  ¼ Dn=dn, where  /n  and /t  are the work of normal and shear separation, respectively;  Dnand   Dt   are the normal and tangential displacement jumps, respectively;   dn   and   dt   are the normal and

    tangential interface characteristic-lengths; and  Dn  is the value of  Dn  after complete shear separation takesplace under the condition of normal tension being zero,  T n ¼ 0.

    Using Eq. (19) in Eq. (16) the interfacial tractions are obtained as

    T n ¼ ð/n=dnÞ expðDn=dnÞfðDn=dnÞ expðD2t =d2t Þþ ½ ð1 qÞ=ðr  1Þ½1 expðD2t =d2t Þ½r  Dn=dng;ð20Þ

    T t ¼ ð/n=dnÞð2dn=dtÞDt=dtfq þ ½ðr  qÞ=ðr  1ÞDn=dng expðDn=dnÞ expðD2

    t =d

    2

    t Þ:   ð21ÞThe works of normal and shear separations are related to  rmax   and  smax, respectively,

    /t ¼ ffiffiffiffiffiffiffi

    e=2p 

      smaxdt;   /n ¼ rmaxedn;   ð22Þwhere e ¼ expð1Þ.

    With Dt ¼ 0, the normal traction–separation relation obtained from Eq. (20) is shown in Fig. 2(a), andthe variation of  T t  with  Dt  given by Eq. (21) for  Dn ¼ 0 is shown in Fig. 2(b).

    The interfacial constitutive relations for the bilinear CZM are given below:

    For  dn  >  0,

    2838   N. Chandra et al. / International Journal of Solids and Structures 39 (2002) 2827–2855

  • 8/21/2019 Chandra 02 Some Issues of Cohesive zone

    13/29

    T n ¼rmaxdmax

    dn;   ðd6 dmaxÞrmaxd

    1d1dmax

    dn;

      ðd >  dmax

    Þ;

      ð23Þ

    T t ¼rmaxdmax

    DcnDctdt;   ðd6 dmaxÞ

    rmaxd

    1d1dmax

    DcnDctdt;   ðd >  dmaxÞ:

    (  ð24Þ

    For  dn ¼ 0,

    T t ¼rmaxdmax

    DcnDctdt;   ðd6 dmaxÞ

    rmaxd

    1d1dmax

    DcnDctdt;   ðd >  dmaxÞ;

    (  ð25Þ

    where  rmax   and   smax   are interface normal and tangential strength, respectively;   dmax   is interface charac-

    teristic length parameter;  Dcn   and  Dct  are the critical normal and tangential separations at which complete

    separation is assumed; and   dn,   dt   and   d   denote the non-dimensional normal, tangential and total dis-placement jumps respectively, defined by

    dt ¼ DtDct

    ;   dn ¼ DnDcn

    ;   d ¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffid

    2t þ d2n

    q   :   ð26Þ

    For pure opening ðDt ¼ 0Þ  and pure shear separation ðDn ¼ 0Þ, the variation of normal and tangentialtractions with respect to  Dn   and  Dt   are shown in Fig. 3(a) and (b), respectively. The normal ðCnÞ   andtangential ðCtÞ  works of separation per unit area of interface are given by

    Fig. 2. (a) Variation of normal traction,  T n, with  Dn   for D t ¼ 0. (b) Variation of shear traction,  T t, with  Dt   for  Dn ¼ 0.

    Fig. 3. (a) Normal traction  T n  as a function of the normal separation  dn   for dt  ¼ 0, (b) shear traction  T t   vs. shear separation  d t   fordn ¼ 0.

    N. Chandra et al. / International Journal of Solids and Structures 39 (2002) 2827–2855   2839

  • 8/21/2019 Chandra 02 Some Issues of Cohesive zone

    14/29

    Cn ¼ rmaxDcn=2;   Ct ¼ smaxDct=2:   ð27Þ

    5. Finite element approach

    To evaluate the interfacial properties of composites, the single fiber push-out test has emerged as the de-

    facto experimental tool. The objective of any test for characterizing the interface is to extract qualitative as

    well as quantitative information about the properties. But many factors make it clear that extracting

    quantitative data from the single push-out test is not trivial.

    The thin-slice push-out test is modeled as an axisymmetric problem (Kallas et al., 1992), since the

    loading (both thermal and mechanical) produces axially symmetric displacement fields. The discretized

    mesh has 4800 axisymmetric four-node elements to model the fiber and matrix. Duplicate nodes are created

    at the interface on the fiber and matrix sides. 240 axisymmetric cohesive elements with each having four

    nodes and zero thickness in the direction normal to the interface are used to model the interface behavior.

    The general-purpose commercial code ABAQUS (1998) is employed to carry out the analysis due to itscapability of handling non-linear problems and also, because of its flexibility in allowing user-defined

    subroutines to be linked to the main program. The cohesive element model is input as a user-defined el-

    ement subroutine UEL into ABAQUS. The mesh used in the simulation is shown in Fig. 4.

    To simulate the single fiber push-out test, the elastic constitutive behavior is assumed for the fiber, and

    the matrix is assumed to be a rate independent elastic–plastic material. The temperature dependency of the

    elastic and inelastic properties (Nimmer et al., 1991; Kroupa and Neu, 1994) of the constituent phases is

    included in a piece-wise linear manner. The material properties used in the analysis are given in Fig. 5. The

    analysis is done in the following three steps:

    1. In step 1, the cooling process after the composite consolidation at high temperature is modeled. A ref-

    erence temperature  T ref   is assumed above which the composite is stress free, which is taken as 815   C.

    Thermal residual stress (TRS) is induced due to the coefficient of thermal expansion mismatch whenthe sample is cooled down from   T ref   to ambient temperature. The boundary conditions for this step

    are shown in Fig. 6(a). The nodes on the top surface are tied to have equal displacement to create a gen-

    eralized plane strain condition.

    2. When a thin slice is cut out of the bulk composite plate, the residual stresses tend to relax. This is sim-

    ulated by removing the existing tying constraints and boundary conditions from the top and bottom, and

    Fig. 4. The model used for push-out test simulation and details of interface element.

    2840   N. Chandra et al. / International Journal of Solids and Structures 39 (2002) 2827–2855

  • 8/21/2019 Chandra 02 Some Issues of Cohesive zone

    15/29

    allowing the stresses in the specimen to relax and reach equilibrium. The symmetry boundary conditions

    used for this step are shown in Fig. 6(b). This process results in shear stresses at the interface due to the

    differential axial residual strains between the fiber and matrix, as shown in the Fig. 7.

    3. Axial displacement is applied to the punch shown in Fig. 6(c) until the fiber slides out of the matrix after

    the entire length of the fiber debonds.

    6. Results and discussion

    6.1. Distribution of thermal residual stresses

    The role of TRS is often ignored in analytical and experimental considerations of interfacial effects in

    metallic based composite materials. This oversight is unfortunate because the resultant interpretation of 

    properties and behaviors is usually misleading. TRS are inherent characteristics of composite materials

    arising from the differing coefficients of thermal expansion of the component materials. Since composites

    Fig. 5. Variation of material properties with temperature.

    N. Chandra et al. / International Journal of Solids and Structures 39 (2002) 2827–2855   2841

  • 8/21/2019 Chandra 02 Some Issues of Cohesive zone

    16/29

    are invariably used at a different temperature than that at which they are fabricated, the differing thermal

    expansions or contractions of the fiber and matrix set up the thermally induced residual stresses around the

    interface on cool down from consolidation temperature. Typical MMCs and IMCs are fabricated at

    temperatures which are quite high relative to ambient (above 815  C for Ti-MMCs (Yang et al., 1991; Sohi

    et al., 1991)), and thus hold the possibility of producing very high stress levels that influence the fracture

    and failure process of the material system.

    The distribution of TRS induced during the cooling process (step 1) is shown in Fig. 7. It can be seen that

    the fiber and matrix are under uniform axial compression and tension respectively to satisfy the force

    equilibrium. Also, the fiber has similar distribution for radial and tangential stresses and the matrix is

    subjected to compressive radial and tensile tangential stresses.

    Fig. 7. Distribution of TRS along cross-section after cooling at ambient for sample thickness of 0.63 mm.

    Fig. 6. Boundary conditions used in the finite element analysis.

    2842   N. Chandra et al. / International Journal of Solids and Structures 39 (2002) 2827–2855

  • 8/21/2019 Chandra 02 Some Issues of Cohesive zone

    17/29

    The evolution of shear stress distribution along the interface from step 2 to 3 (see Fig. 6) of the push-out

    test is shown in Fig. 8.

    6.2. Comparison between experimental results and numerical predictions

    The non-aged Timetal-21S/SCS-6 composite system was fabricated and push-out tested as a part of this

    work. The reinforcing fibers are made of SCS-6 materials with 142  lm diameter, with the volume fraction of 

    fiber being 35%. The following three cases selected from a set of experimental data of Osborne et al. (2000),

    are used in the current study:

    test temperature: 25   C; sample thickness: 0.63 mm,

    test temperature: 500   C; sample thickness: 1.07 mm,

    test temperature: 650   C; sample thickness: 1.10 mm.

    Force–displacement data obtained from the experiments are shown in Fig. 9, and are the average of at

    least two data sets in each case.

    Finite element simulation was carried out on the model shown in Fig. 4 as per the three steps discussed

    earlier. The bilinear CZM (see Eqs. (23)–(27)) for the interface is characterized by four parameters:  dmax,

    smax, Dct   and D

    cn. Since shear failure is by far the dominant failure mechanism in current study, it is assumed

    that  Dcn  is equal to  Dct . Therefore, only three parameters need to be determined in the simulation. To de-

    termine smax and Dct , the work of shear separation ð/tÞ is first estimated based on the area under the force– 

    displacement curves obtained from the experiment (Fig. 9). The values of  smax  for various temperatures are

    chosen so that the experimental response is closely matched. And then  Dct  is determined by using Eq. (27).

    The values used in the simulation are listed in Table 2.

    Fig. 8. Shear stress distribution along the interface. (a) After slicing, (b) beginning of loading, (c) debonding occur and (d) complete

    debonding. Note: the displacement is applied at the left side of fiber (see arrow).

    N. Chandra et al. / International Journal of Solids and Structures 39 (2002) 2827–2855   2843

  • 8/21/2019 Chandra 02 Some Issues of Cohesive zone

    18/29

    The experimental and computational results for various temperatures (i.e. 25, 500 and 650  C) are shown

    in Fig. 9. As noted above, the samples for 500 and 650  C are approximately of the same thickness and can

    be compared directly. However, due to maximum force limitations, the sample for room temperature is

    thinner. First, it can be seen from both experiments and simulations that the peak load decreases with the

    increase of temperature due to the relaxation of compressive residual stresses which act on the fiber as a

    clamping force at elevated temperatures (Davidson, 1992). The simulation response for various tempera-

    tures matches the experimental results very well. Also, it can be seen that it is linear for room temperature

    until the maximum load is reached. This is followed by catastrophic and complete debonding. For elevated

    temperatures, each curve has an initial linear portion, and as the load continues to increase, a reduction in

    slope can be observed (especially for the plot at 650   C). It was reported that the change in slope from

    linearity for elevated temperatures corresponds to the initiation of the interfacial debond crack (Kerans and

    Parthasarathy, 1991). However, from the current simulation results, it is found that the debonding does not

    take place until the peak load is reached. It was reported (Davidson, 1992) that the matrix material behaves

    elastically at lower temperature. At higher temperature (>400   C), the ability of the matrix phase to flowunder the applied stress is enhanced. It is shown by simulation that the plastic deformation in the matrix

    indeed takes place along the entire interface during the loading process at elevated temperatures. Even at

    ambient temperature, the plastic deformation occurs near the bottom along the interface. However, it is just

    not enough to cause the slope change in macroscopic force–displacement curve. Hence, it should be rea-

    sonable to conclude that the change of slope in the curves during loading at elevated temperatures is due

    not to the initiation of interfacial debonding, but to plastic deformation of the matrix material at the in-

    Fig. 9. Experimental measurement and numerical predictions, at different temperatures. Note: displacement for each temperature is

    relative to the starting position.

    Table 2

    The values of CZM parameters used for simulation

    Temperature   dmax   smax  (MPa)   Dct   (m)   D

    cn   (m)   /t   (J m

    2)

    25   C 1.0 200 3.5E5 3.5E5 3000500   C 1.0 100 1.3E5 1.3E5 650650   C 1.0 90 1.42E5 1.42E5 640

    2844   N. Chandra et al. / International Journal of Solids and Structures 39 (2002) 2827–2855

  • 8/21/2019 Chandra 02 Some Issues of Cohesive zone

    19/29

    terface. It is clear that the choice of  smax   is very critical to reproduce the macroscopic mechanical response

    of the material system.

    6.3. Comparison between two cohesive zone models

    At the outset it appears that the CZM for the interface and conventional field equations with appro-

    priate constitutive equations for the bulk (matrix and fiber) will reproduce the test results. Consequently,

    the agreement leads to a set of cohesive zone parameters (see Table 2) that can then be used for other

    geometric and loading configurations. Before such a positive assertion can be made, it is worthwhile to

    analyze the uniqueness of the shape of the CZM and the sensitivity of the chosen parameters on matching

    the experimental data.

    For the purpose of examining whether the form of the CZM is important if the energy (area under force– 

    displacement curve) is maintained at the same level, two forms (exponential and bilinear) are chosen. An

    energy value of  /n ¼ /t ¼ 3000 J/m2 was used for the two forms. The simulation results for bilinear andexponential forms are shown in Fig. 10(a) and (b), respectively. The value of  smax  was selected as 200 MPa(as in Table 2 for ambient temperature case), and then altered to 300 and 400 MPa, which corresponds to

    curves 1–3 (Fig. 10(a) and (b)), respectively. It should be noted that a bilinear CZM with  smax ¼ 200 MPamatches the experiment very well.

    Fig. 10(a) and (b) bring home some very important conclusions regarding a number of issues. Fig. 10(a)

    illustrates that even though the cohesive energy is the same, increasing the interfacial shear strength changes

    the F  – DU   response of the push-out test. This result is in contrast to the common belief that the separation

    process is governed mainly by the energy of separation. Again, focusing on  smax ¼ 300 and 400 MPa in Fig.10(a), there is a non-linear response when the force is above 50 N. On closer examination, it is observed that

    matrix yields under those conditions, leading to the slope change in the macroscopic response. Arrows in

    the Fig. 10(a) indicate the loading at which the initial plastic deformation occurs. The initial debonding for

    all three cases, however, takes place at the peak load. Hence, the average shear stress required for failure

    savg ¼ P peak=2pd L is different in the three cases. The present discussions again clearly reveal that selection of smax  is critical in determining the macroscopic mechanical response.

    Fig. 10. Load–displacement plots simulated using exponential and bilinear CZMs. (a) Bilinear model (Geubelle et al., 1998), (b) ex-

    ponential model (Xu and Needleman, 1993). Note: Pd –– plastic deformation; Db –– debonding.

    N. Chandra et al. / International Journal of Solids and Structures 39 (2002) 2827–2855   2845

  • 8/21/2019 Chandra 02 Some Issues of Cohesive zone

    20/29

    Fig. 10(b) shows the overall response of the push-out test for the exponential model keeping the cohesive

    energy and interfacial strength identical to those of Fig. 10(a). It is immediately clear that the predicted

    response is in no way similar to the experimental observations. Arrows in the Fig. 14(b) indicate the loading

    at which the initial debonding occurs. The important question is why there is such a drastically different

    response. In fact, no amount of parametric variations on the exponential CZM can make it match with

    experimental results making it unsuitable for modeling the interface in this problem. Focusing on Fig.

    10(b), for smax ¼ 200 MPa, the peak load (about 53 N) does not correspond to either the initial debondingor the matrix yield stress. As the push-out load is increased (truly the push-out displacement, see Fig. 8(a)),

    the interface responds to the external force according to its constitutive equation. When the peak stress in

    each element is reached (Fig. 10(b)), the slope is reversed and the interface becomes more and more

    compliant. Thus the overall stiffness of all the interfaces in tandem is continually altered with different

    cohesive elements changing their stiffness as the push out displacement increases. Since the separation in

    this model takes place with a significantly large displacement, the first debonding does not occur until very

    late in the process.

    It is interesting to note that in both Fig. 10(a) and (b), the shape of the overall force–displacement re-

    sponse ( F  – DU , along the longitudinal fiber axis) is similar to that of input traction–separation curve of 

    CZM (T  – d, normal to the interface in the radial direction, see Figs. 2 and 3). This enigmatic coincidence can

    be explained on the basis of the geometry of the push-out test and the loading process. A simplified model is

    presented in the next section to explain why the two responses show a close resemblance.

    6.4. A simplified model for push-out test based on cohesive zone model 

    In this section, we develop a simplified model of the push-out test problem. The circular ceramic fiber is

    connected to the metallic matrix by means of interfaces. While the ceramic domain  X1  is modeled as purely

    elastic, the metallic matrix  X2  is modeled as temperature-dependent elastic-linearly strain hardening plastic

    material governed by von Mises yield criterion.  X  is modeled by cohesive zone elements discretized as  N 

    four-node axisymmetric elements governed by traction–separation laws, through  T  – d relations in the finite

    element analysis. When the push-out load F  is applied, the end of the fiber moves by DU . We would like to

    establish a relationship between the global  F  – DU  and the cohesive zone  T  – d  relations.

    As shown in Fig. 11, we have simplified cohesive zone finite elements as nonlinear springs. We have pairs

    of nodes connected by springs at f1; 10g, f2; 20g; . . . ; fn; n0g co-located in the fiber and the matrix before theapplication of load. Let T i  be the traction at a spring  i  after some global load  F  is applied. Without loss of 

    generality, let f i  be the shear force given by f i ¼ pdT ili, where d  is the fiber diameter, T i  is the average sheartraction in the region of spring ði ði 1ÞÞ=2 to ði þ ði þ 1ÞÞ=2 and  li  the axial length of the region. Theshear displacement of spring   i  is governed by  T  – d  relation.

    From the equilibrium of force, we have

     F  ¼ X N 

    i¼1  f i:   ð28Þ

    6.4.1. Case (1): bilinear model  d <  dmaxAssume that the matrix has not yielded; for   ith spring,  T i ¼ k idi   and  f i ¼ ðpdiliÞk id , so that

     F  ¼  pdk iX N i¼1

    dili:   ð29Þ

    Since the effective length of the spring remains constant, the total length can be expressed as a linear

    combination of each of them. Also

    P N 

    1  di ¼ DU , leading to

    2846   N. Chandra et al. / International Journal of Solids and Structures 39 (2002) 2827–2855

  • 8/21/2019 Chandra 02 Some Issues of Cohesive zone

    21/29

     F  ¼  K 0 DU :   ð30ÞThus the global  F  – DU  curve will be linear with a different slope,  K 0.

    6.4.2. Case (2): exponential model  d <  dmax

    As shown in Fig. 11, since  k  ¼ k ðdÞ, we have the following equation, F  ¼  pd 

    X N i¼1

    k ðdiÞdili;

     F  ¼  K ðDÞDU :ð31Þ

    It is thus seen that we have a monotonically increasing curve, with the slope changing with  DU , i.e.  K ðDÞ.

    6.4.3. Case (3):  dP dmax  (for some springs)

    Since only a few springs will have dP dmax, the traction of others will still be below the peak value. Thus

    in both cases (bilinear and exponential) the slope of   F  – DU  will change when any of the spring exceeds

    d >  dmax.

    6.4.4. Case (4): matrix yielding 

    Whenever the matrix yields, the stress–strain response of the matrix becomes non-linear. Consequently

    the slope of the  F  – DU  curve also changes.

    6.4.5. Case (5): spring failure,  dP dmaxSince the number of loaded springs will be less than  n, the slope of  F  – DU  will change.

    Thus, the simplified analysis shows that a change in the global response of  F  – DU  may be caused by (1)

    change in slope (change in sign) of traction–separation (T  – d) law of CZM, (2) matrix yielding, or (3) co-

    hesive zone failure.

    Fig. 11. (a) The schematic free body diagram of all cohesive zone elements. (b) The bilinear and exponential traction–separation

    relations of CZM.

    N. Chandra et al. / International Journal of Solids and Structures 39 (2002) 2827–2855   2847

  • 8/21/2019 Chandra 02 Some Issues of Cohesive zone

    22/29

    6.5. Parametric study

    In Section 6.2, we evaluated a set of parameters (Table 2) within the framework of the bilinear CZM that

    matched the experimental results. However, numerical simulation clearly shows that there are many other

    parameters that significantly alter the progression of debonding and hence the load–displacement behavior.

    In this section, the effects of various parameters, i.e., work of shear separation ð/tÞ, interfacial shearstrength ðsmaxÞ, test temperature and interface characteristic length ðdmaxÞ are investigated systematically byusing bilinear CZM in terms of load–displacement response of the push-out test. The main purpose is not

    only to simulate a specific material system, but also to provide some insights on the effects of the chosen

    parameters on the deformation and failure behavior of the materials in general.

    6.5.1. Effect of interface characteristic length,  dmaxIn this section, the effect of interface characteristic length,   dmax, on the force–displacement curve is

    studied and the results are shown in Fig. 12. The values of  dmax  are selected to be 0.1, 0.4, 0.7 and 1.0; the

    corresponding F  – 

    DU  curves are I, II, III and IV, respectively. It is interesting to find that the shape of theforce–displacement curve is dependent on the value of interface characteristic length, which determines the

    shape of shear traction vs. shear separation. The arrows in the figure indicate the loading at which deb-

    onding occurs at the interface. Comparing the numerical results with the experimental data, it can be

    concluded that for the current material system, the suitable value for  dmax  is 1.0. Therefore, it is reasonable

    to conclude that the  shape  of the traction–separation curve for the CZM is an important factor in deter-

    mining the macroscopic mechanical behavior of the interface in addition to the other interfacial parameters

    such as the cohesive energy. Thus, it is obvious that setting dmax ¼ 1:0 fixes the shape of the CZM; in otherwords, any other values of   dmax   corresponding to different shapes will yield unacceptable results for the

    present problem of metal–ceramic interfaces. This is an important point to be considered while selecting

    specific shapes of CZM for a given problem.

    Fig. 12. Effect of interface characteristic length   dmax  on load–displacement plots (I, II, III and IV represent corresponding force– 

    displacement curves to different values of characteristic lengths 0.1, 0.4, 0.7 and 1.0, respectively).

    2848   N. Chandra et al. / International Journal of Solids and Structures 39 (2002) 2827–2855

  • 8/21/2019 Chandra 02 Some Issues of Cohesive zone

    23/29

    Having established that the shape is critical, we proceed to conduct parametric study of other parameters

    (/t,  smax   or  dsep) that describe the bilinear model.

    6.5.2. Effect of work of shear separation,  /t The effect of the work of shear separation,  /t, on the force–displacement curve is discussed in the fol-

    lowing section and the results are shown in Fig. 13. The work of shear separation is varied from 100 to 3000

    J/m2 and the curves 1–4 correspond to the work of shear separation of 100, 500, 1000 and 3000 J/m2,

    respectively. The interfacial shear strength smax (200 MPa) is kept constant. The total area under the force– 

    displacement curve represents the overall work required to separate the fiber from the matrix (complete

    debonding). It is found from the simulation that the debonding begins to occur at the right end (Fig. 8) just

    when the peak load is reached. And the subsequent debonding process along the interface from right to the

    left takes place at a very fast pace (corresponding to the unloading portion of curves in Fig. 13). The slope

    of the  F  – DU  curve is steeper for the case with lower  /t  and this slope reduces with an increase in  /t. As

    happens in the push-out test the entire fiber is debonded soon after the first cohesive element (at the left

    end) fails. The actual critical separation distances  D

    c

    t  at which the cohesive element debonds completely areDct ¼  1 106, 5 106, 1 105, 3 105 m, corresponding to  /t ¼ 100, 500, 1000, 3000 J/m2 respec-tively. Thus, sufficiently large tangential displacement occurs at the left element before the entire fiber is

    being pushed out.

    6.5.3. Effect of shear strength,  smaxIn this section, simulations were carried out by keeping   /t   constant and varying   smax   (and corre-

    spondingly critical tangential displacement Dct  through the Eq. (27) for the bilinear CZM). The value of  dmaxis set equal to 1 as explained earlier.

    Fig. 14 shows the macroscopic response of force–displacement curves with different interfacial shear

    strengths,  smax. It is seen that when  smax6 200 MPa, there is a linear relationship between force and dis-

    placement. For cases when  smax ¼ 300 or 400 MPa, the slope of the ascending part of the curve is found to

    Fig. 13. Effect of work of shear separation on force–displacement curves.

    N. Chandra et al. / International Journal of Solids and Structures 39 (2002) 2827–2855   2849

  • 8/21/2019 Chandra 02 Some Issues of Cohesive zone

    24/29

    change when the load reaches about 55 N. It is observed from the results that yielding in the matrix occurs

    at this load and is responsible for the change in slope. It is worthwhile to note that at critical points where

    debonding initiates, the matrix yields when the stress due to applied loading combined with the TRS

    reaches the yield stress of 1043 MPa for a given external load (55 N in this case). It can also be noted that

    the force–displacement curve response for   smax ¼ 200 MPa is one closer to the experimental result. It isgenerally assumed that given the cohesive energy  /t   and the cohesive strength  smax, the behavior of the

    CZM can be uniquely described. But it is clear from the present work, apart from  /t  and  smax, the shapeðdmax ¼ 1Þ is required to uniquely and completely describe the CZM. On the other hand when  dmax 6¼ 1, orother CZMs are used, the response will be quite different from that of experimental results (see Figs. 10(b)

    and 12).

    6.5.4. Effect of temperature

    For sample with thickness of 0.63 mm, the simulation is performed to study the effect of temperature on

    the force–displacement curve and the results are shown in Fig. 15(a) and (b) for cohesive energies of 3000

    and 1000 J/m2, respectively. As expected, the peak value of load decreases as the temperature increases. It

    can also be seen that magnitude of peak force reduces with the decrease in the cohesive energy (Fig. 15(b)

    compared to Fig. 15(a)). For the simulation at room temperature, the load keeps increasing linearly with

    displacement until the load reaches the peak value, and then debonding initiates from the left end. On theother hand, at higher temperatures the load increases linearly initially; then the slope changes with the onset

    of plastic deformation in the matrix. Also, it should be noted that at higher temperatures, the temperature-

    dependent yield stress value is considerably lower than that at room temperature which results in plastic

    deformation in the matrix at a much lower external load (

  • 8/21/2019 Chandra 02 Some Issues of Cohesive zone

    25/29

    that there is a significant amount of residual shear stress even before the application of any load. So it is

    interesting to examine the absorption of cohesive energy by various elements along the length, especially at

    both ends. The growth of interfacial shear traction and work of shear separation in the cohesive elements

    during mechanical loading (step 3), are shown in Figs. 16 and 17 for two cases (case 1:  dmax ¼ 0:1; case 2:dmax ¼ 1:0). The evolution of shear traction   T t   and interfacial cohesive energy  /t   in Figs. 16 and 17 isplotted along the interface at various stages of loading as indicated by 1–5 in Fig. 12.

    Figs. 16(a) and 17(a) show that the distribution of shear traction at early stages of mechanical loading is

    anti-symmetric with respect to the sample center; the right side (top) is in positive shear and the left side(bottom) is in negative shear, indicated by the curve 1. With slight increase in loading, the shear traction of 

    Fig. 15. The effect of temperature on force–displacement curve (a)  / t ¼ 3:0 103 and (b)  / t ¼ 1:0 103.

    Fig. 16. (a) The evolution of shear traction and (b) work of shear separation for case 1 ðdmax ¼ 0:1, /t ¼ 3000, smax ¼ 200Þ (1–5 indicatevarious loading positions shown on curve I in Fig. 12).

    N. Chandra et al. / International Journal of Solids and Structures 39 (2002) 2827–2855   2851

  • 8/21/2019 Chandra 02 Some Issues of Cohesive zone

    26/29

    all elements along the interface turns to negative shear traction and also reaches the maximum value  smax as

    indicated by 2 and 3 in Figs. 16(a) and 17(a). These stages indicated by 1–3 lie along the ascending part of 

    the T  – d curve (Fig. 3). After reaching the peak value  smax the shear traction begins to reduce to zero and the

    interface debonds as, indicated by curves 4 and 5 in Fig. 16(a) and curve 4 in Fig. 17(a). The stages in-

    dicated by 4 and 5 lie along the descending part of  T  – d   curve. It is interesting to note that mechanical

    loading is applied at the right end (right side in the plot) and the shear traction first reaches zero in a few

    elements on the bottom (left side on the plot), indicating that the fracture has initiated from the left end,which is consistent with the experimental results (Ghosn et al., 1994).

    Figs. 16(b) and 17(b) show the evolution of the work of shear separation (cohesive energy) in the co-

    hesive elements along the interface. In both figures the cohesive energy along the interface is zero at the

    beginning of loading and gradually increases with the increase in the mechanical loading, as indicated by

    curves 1–5 in Fig. 16(b) and curves 1–4 in Fig. 17(b). As observed earlier, with the increase of mechanical

    loading, the work of shear separation reaches its critical value  /t  first for a few elements near the bottom

    end (left side in the plot) rather than that at right side (top) indicating that the debonding always begins to

    occur at the bottom for the current material system. The cause of bottom fracture initiation during thin-

    slice push-out tests of MMCs has been found to be predominantly due to TRSs (Ananth et al., 1995;

    Chandra et al., 1995).

    7. Summary and conclusions

    This work can be summarized as follows:

    •   The concept of continuum mechanics is extended to include a zone of discontinuity modeled by cohesive

    zones. A comprehensive analysis of some of the popular CZMs has been presented with respect to the

    form of the equations and the magnitude of the parameters. It is clearly demonstrated that energy values

    ranging from J to kJ, traction values ranging from MPa to GPa and separation distances ranging from

    nm to lm are employed to describe interfaces. We illustrate some of the outstanding issues, e.g., physical

    Fig. 17. (a) The evolution of shear traction and (b) work of shear separation for case 1ðdmax ¼ 1:0, /t ¼ 3000, smax ¼ 200). (1–4 indicatevarious loading positions shown on curve IV in Fig. 12.)

    2852   N. Chandra et al. / International Journal of Solids and Structures 39 (2002) 2827–2855

  • 8/21/2019 Chandra 02 Some Issues of Cohesive zone

    27/29

    interpretation of the area under the curve, form of the equations (shape of the traction curves), cohesive

    strength and scales in CZMs.

    •   Two CZMs (exponential and bilinear) have been employed to model the interface failure in metal matrix

    composites. The bilinear CZM is suitable to simulate the push-out test process and the calculated results

    match the experiment data very well. The macroscopic mechanical behavior ( F  – DU   curve shape) of 

    interfaces in the material system has a striking resemblance to that of CZM (shape of interfacial trac-

    tion–separation law) provided no plastic deformation in the matrix takes place. This implies that the

    macroscopic mechanical response of the interface of the material system is very sensitive to the shape

    of the interfacial constitutive relation of the CZM in addition to the other interfacial parameters.

    •   Debonding of cohesive elements takes place when the accumulated cohesive energy reaches the critical

    value, as prescribed in the model. It has been observed during the push-out tests, when the load/displace-

    ment is applied at the top end, a few cohesive elements near the bottom end first attain the critical en-

    ergy, leading to initiation of fracture at the bottom end.

    •   CZMs can be used to describe a broad range of fracture processes in a wide variety of material systems.

    However, in order to have truly predictive capabilities, this phenomenological failure model must rely on

    carefully designed and conducted experiments to select a proper cohesive zone constitutive relation. Sucha relation should include the shape and appropriate material parameters such as interfacial strength,

    work of separation and critical displacement discontinuities, and response during loading and unloading.

    •   It is our view that the CZM represents the physics of the interface separation process and hence the

    shape of the CZM should in some sense depend on the inelastic processes occurring at the micromechan-

    ical level. Hence when using CZMs to model separation in a given material system, an appropriate shape

    (form), depending on the type of material system and the inelastic micromechanical processes, should be

    used. Otherwise, the CZM based modeling and simulation will not yield meaningful results.

    Acknowledgements

    The authors wish to acknowledge the Air Force Office of Scientific Research (F49620-99-10275) for

    providing partial financial assistance in support of this research.

    References

    ABAQUS, 1998. Manual, version 5.8, Habbit, Karlsson & Sorensen, Inc., USA.

    Ananth, C.R., Chandra, N., 1995a. Evaluation of interfacial shear properties of metal matrix composites from fiber push-out tests.

    Mechanics of Composite Materials and Structures 2, 309–328.

    Ananth, C.R., Chandra, N., 1995b. Numerical modeling of fiber push-out test in metallic and intermetallic matrix composites –– 

    Mechanics of the failure processes. Journal of Composite Materials 29 (11), 1488–1514.

    Barenblatt, G.I., 1959. The formation of equilibrium cracks during brittle fracture. General ideas and hypothesis. Axially-symmetric

    cracks. Prikl. Matem. I mekham 23, 434–444.

    Barenblatt, G.I., 1962. Mathematical theory of equilibrium cracks in brittle fracture. In: Dryden, H.L., von Karman, T. (Eds.),

    Advances in Applied Mechanics, vol. VII. Academic Press, New York, pp. 55–125.

    Camacho, G.T., Ortiz, M., 1996. Computational modeling of impact damage in brittle materials. International Journal of Solids and

    Structures 33, 2899–2938.

    Chandra, N., Ananth, C.R., 1995. Analysis of interfacial behavior in MMCS and IMCS using thin-slice push-out tests. Composite

    Science and Technology 54 (1), 87–100.

    Chandra, N., Ghonem, H., 2001. Interfacial mechanics of push-out tests: theory and experiments. Composites Part A –– Applied

    Science and Manufacturing 32 (3–4), 575–584.

    Davidson, D.L., 1992. The micromechanics of fatigue crack growth at 25  C in Ti-6Al-4V reinforced with SCS-6 fibers. Metallurgical

    Transactions 23A, 865–879.

    N. Chandra et al. / International Journal of Solids and Structures 39 (2002) 2827–2855   2853

  • 8/21/2019 Chandra 02 Some Issues of Cohesive zone

    28/29

    Dugdale, D.S., 1960. Yielding of steel sheets containing slits. Journal of the Mechanics and Physics of Solids 8, 100–104.

    Espinosa, H.D., Dwivedi, S., Lu, H.C., 2000. Modeling impact induced delamination of woven fiber reinforced composites with

    contact/cohesive laws. Computational Methods in Applied Mechanics and Engineering 183 (3–4), 259–290.

    Foulk, J.W., Allen, D.H., Helms, K.L.E., 2000. Formulation of a three-dimensional cohesive zone model for application to a finite

    element algorithm. Computational Methods in Applied Mechanics and Engineering 183, 51–66.Geubelle, P.H., Baylor, J., 1998. Impact-induced delamination of laminated composites: a 2D simulation. Composites Part B

    Engineering 29 (5), 589–602.

    Ghosn, L.J., Eldridge, J.I., Kantzos, P., 1994. Analytical modeling of the interfacial stress state during push-out testing of SCS-6/

    Ti-based composites. Acta Metallurgica et Materialia 42, 3895–3908.

    Hutchinson, J.W., Evans, A.G., 2000. Mechanics of materials: top-down approaches to fracture. Acta Materialia 48 (1), 125–135.

    Jan, G.M. van Mier, 1996. Fracture Processes of Concrete, Assessment of Material Parameters for Fracture Models. CRC press, Boca

    Raton, Florida, p. 291.

    Jayaraman, K., Reifsnider, K.L., Swain, R.E., 1993. Elastic and thermal effects in the interphase: part II. Comments on modeling

    studies. Journal of Composites Technology & Research 15 (1), 14–22.

    Kallas, M.N., Koss, D.A., Hahn, H.T., Hellman, J.R., 1992. Interfacial stress state present in a thin-slice fibre push-out test. Journal of 

    Material Science 27 (14), 3821–3826.

    Kerans, R.J., Parthasarathy, T.A., 1991. Theoretical analysis of the fiber pullout and push out tests. Journal of the American Ceramic

    Society 74 (7), 1585–1596.

    Kim, J.K., Mai, Y.W., 1991a. High strength high fracture toughness fiber composites with interface control –– 

    A review. CompositesScience Technology 41 (4), 333–378.

    Kim, J.K., Mai, Y.W., 1991b. The effect of interfacial coating and temperature on the fracture behaviors of unidirectional KFRP and

    CFRP. Journal of Materials Science 26 (17), 4702–4720.

    Kroupa, J.L., Neu, R.W., 1994. The non-isothermal viscoplastic behavior of a titanium-matrix composite. Composites Engineering 4

    (9), 965–977.

    Mohammed, I., Liechti, K.M., 2000. Cohesive zone modeling of crack nucleation at bimaterial corners. Journal of the Mechanics and

    Physics of Solids 48, 735–764.

    Mukherjee, S., Ananth, C.R., Chandra, N., 1997. Effect of residual stresses on the interfacial fracture behavior of metal-matrix

    composites. Composite Science and Technology 57 (11), 1501–1512.

    Mukherjee, S., Ananth, C.R., Chandra, N., 1998. Effect of interface chemistry on the fracture properties of titanium matrix

    composites. Composites Part A –– Applied Science and Manufacturing 29 (9–10), 1213–1219.

    Needleman, A., 1987. A continuum model for void nucleation by inclusion debonding. Journal of Applied Mechanics 54,

    525–531.

    Needleman, A., 1990a. An analysis of tensile decohesion along an interface. Journal of the Mechanics and Physics of Solids 38, 289– 

    324.

    Needleman, A., 1990b. An analysis of decohesion along an imperfect interface. International Journal of Fracture 42, 21–40.

    Needleman, A., 1992. Micromechanical modeling of interfacial decohesion. Ultramicroscopy 40 (3), 203–214.

    Needleman, A., 1997. Numerical modeling of crack growth under dynamic loading conditions. Computational Mechanics 19 (6), 463– 

    469.

    Nimmer, R.P., Bankert, R.J., Russel, E.S., Smith, G.A., Wright, P.K., 1991. Micromechanical modeling of fiber-matrix

    interface effects in transversely loaded SiC/Ti-6-4 metal matrix composites. Journal of Composites Technology & Research 13

    (1), 3–13.

    Osborne, D., Ghonem, H., Chandra, N., 2000. Interface behavior of Ti matrix composites at elevated temperature. Composites Part

    A –– Applied Science and Manufacturing 32 (3–4), 545–553.

    Rahulkumar, P., Jagota, A., Bennison, S.J., Saigal, S., 2000. Cohesive element modeling of viscoelastic fracture: application to peel

    testing of polymers. International Journal of Solids and Structures 37, 1873–1897.

    Rice, J.R., Wang, J.-S., 1989. Embrittlement of interfaces by solute segregation. Material Science and Engineering A 107, 23–40.Rose, J.H., Ferrante, J., Smith, J.R., 1981. Universal binding energy curves for metals and bimetallic interfaces. Physical Review

    Letters 47 (9), 675–678.

    Sohi, M., Adams, J., Maahapatra, R., 1991. Transverse constitutive response of titanium aluminum metal matrix composites. In:

    Desai, C.D. (Ed.), Constitutive Laws for Engineering Materials. ASME Press, New York, pp. 617–626.

    Tvergaard, V., 1990. Effect of fibre debonding in a whisker-reinforced metal. Material Science and Engineering A 125 (2), 203–213.

    Tvergaard, V., Hutchinson, J.W., 1992. The relation between crack growth resistance and fracture process parameters in elastic–plastic

    solids. Journal of the Mechanics and Physics of Solids 40 (6), 1377–1397.

    Wappling, D., Gunnars, J., Stahle, P., 1998. Crack growth across a strength mismatched bimaterial interface. International Journal of 

    Fracture 89, 223–243.

    Xu, X.P., Needleman, A., 1993. Void nucleation by inclusion debonding in a crystal matrix. Modelling and Simulation in Materials

    Science and Engineering 1 (2), 111–132.

    2854   N. Chandra et al. / International Journal of Solids and Structures 39 (2002) 2827–2855

  • 8/21/2019 Chandra 02 Some Issues of Cohesive zone

    29/29

    Xu, X.P., Needleman, A., 1994. Numerical simulation of fast crack growth in brittle solids. Journal of the Mechanics and Physics of 

    Solids 42, 1397–1434.

    Xu, X.P., Needleman, A., 1995. Analysis of ductile crack growth by means of a cohesive damage model. International Journal of 

    Fracture 81, 99–112.

    Yang, J.M., Jeng, S.M., Yang, C.J., 1991. Failure mechanisms of fiber reinforced titanium alloy matrix composites, part I: interfacebehavior. Material Science and Engineering A 138, 155–167.

    N. Chandra et al. / International Journal of Solids and Structures 39 (2002) 2827–2855   2855