Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

18164 NI

INSTITUTE OF PHYSICS PUBLISHING MODELLING AND SIMULATION IN MATERIALS SCIENCE AND ENGINEERING
Modelling Simul. Mater. Sci. Eng. 9 (2001) 143–155 www.iop.org/Journals/ms PII: S0965-0393(01)21951-2

Three-dimensional analyses of ductile failure in metal


reinforced by staggered fibres

Viggo Tvergaard
Department of Solid Mechanics, Technical University of Denmark, Nils Koppels Allé,
Building 404, DK-2800 Kongens Lyngby, Denmark

Received 8 November 2000, accepted for publication 5 February 2001

Abstract
Ductile matrix failure in a metal reinforced by short parallel fibres is studied
numerically in terms of full three-dimensional cell model analyses. The model
is extended here, relative to the simplest possible cell, such that transversely
staggered fibres are accounted for and such that effects of different sizes of two
neighbouring fibres can be studied. The matrix material is described in terms of
a porous ductile material model, which accounts for the nucleation and growth
of voids to coalescence. Most predictions are based on an isotropic hardening
model, but results for kinematic hardening are used to study the effect of a metal
that forms a rounded vertex on the yield surface. The damage evolution tends
to first show the formation of an open crack near the ends of the longer fibres,
by void coalescence in the matrix, and subsequently a similar crack evolves
near the ends of the shorter fibres.

1. Introduction

In metal matrix composites the use of short brittle fibres gives good stiffness and tensile strength
properties, but also results in poor ductility and low fracture toughness relative to that of the
unreinforced metal [1, 2]. This reduced ductility occurs due to either ductile matrix failure,
brittle failure of the reinforcements or debonding of the matrix–fibre interface (Needleman
et al [3]). Several experimental observations have demonstrated these failure mechanisms,
including observations of failure by the nucleation and growth of voids in aluminium alloys
reinforced with Al2 O3 or SiC particles [4–7], and failure by debonding as well as brittle fracture
of elongated particles aligned with the tensile direction in Al–SiC composites [8–10].
The focus in the present investigation is on ductile matrix failure. Micromechanical
studies of this type of failure mechanism in short-fibre-reinforced metals have been carried out
by Llorca et al [7, 11], using axisymmetric cell models to study the influence of different fibre
shapes or plane strain cell models to obtain an impression of the effects of fibre clustering.
The nucleation and growth of voids to coalescence in the matrix has been represented in terms
of the Gurson model [12], and it was shown that ductile failure initiation tends to occur near
sharp fibre edges. Subsequently, Tvergaard and Needleman [13] used a non-local porous

0965-0393/01/030143+13$30.00 © 2001 IOP Publishing Ltd Printed in the UK 143


144 V Tvergaard

Figure 1. Periodic array of transversely staggered fibres. Cross sections of the 3D unit cell model
are shown on two coordinate planes.

ductile material model to study the effect of a material length scale for damage in the matrix
on predicted failure.
Full three-dimensional (3D) cell model analyses have recently been used by
Tvergaard [14, 15] to study ductile matrix failure around short fibres in cases where either
the fibre spacings in the transverse directions are non-uniform or the transverse principal
macroscopic stresses differ, so that an axisymmetric model cannot adequately represent
the composite. These were an extension of previous 3D analyses for metal matrix
composites [16–18], where no damage was accounted for. For some cases where axisymmetric
conditions apply, it was found that the predictions in [7] are in good agreement with the initial
parts of the corresponding curves in [14], but the 3D computations were continued much
further, to reach stages where regions of final void coalescence have developed.
In the present paper the study of [14] for transversely aligned fibres is extended to consider
more complex 3D cell models, which account for transversely staggered arrays of fibres,
analogous to the fibre distributions studied in [19, 20] in terms of axisymmetric cell models. In
addition, the 3D cell models considered here account for different lengths of two neighbouring
fibres, as well as differences of fibre spacings or average tractions in two transverse directions.

2. Material model

The metal matrix composite is here modelled in terms of a 3D cell model, which contains parts
of two elastic fibres and for which the elastic–plastic description of the metal matrix accounts
for ductile failure. Periodicities are assumed, but the cell model applied here allows for different
sizes of the two fibres and for non-uniform spacings and stresses in the transverse directions.

2.1. Transversely staggered fibre array


The aligned short fibres are represented as cylinders with hexagonal cross sections, with 1
and r1 denoting the initial half-length and the radius of the inscribed circle for fibre 1, while
2 and r2 denote the corresponding dimensions of fibre 2. This fibre shape was also used in
the previous analyses for transversely aligned fibres [14], where only one fibre was included
in the unit cell analysed. In the present studies, the two fibres are placed in opposite corners
of the unit cell, as illustrated in figure 1, such that a transversely staggered array of fibres is
analysed, analogous to the fibre arrays considered in [19, 20].
3D analyses of ductile failure in metal reinforced by staggered fibres 145

A Cartesian reference coordinate system x i is chosen such that the x 1 -axis is parallel
with the fibres (see figure 1). The initial dimensions of the unit cell, representing spacings
in the periodic array of fibres, are denoted A0 , B0 and C0 in the x 1 -, x 2 - and x 3 -directions,
respectively. Thus, the volume fractions for fibre 1 and fibre 2 are
√ √
3 1 r12 3 2 r22
v1 = v2 = (2.1)
2 A0 B0 C0 2 A0 B0 C0
and the total fibre volume fraction is v1 + v2 .
In the reference coordinate system ui are the displacement components on the base vectors
and T i are the nominal traction components. It is assumed that the solution satisfies symmetry
on all six sides of the unit cell, such that the boundary conditions are:
u̇1 = 0 Ṫ 2 = 0 Ṫ 3 = 0 for x1 = 0
u̇1 = U̇I Ṫ 2 = 0 Ṫ 3 = 0 for x 1 = A0
u̇2 = 0 Ṫ 1 = 0 Ṫ 3 = 0 for x2 = 0
u̇2 = U̇II Ṫ 1 = 0 Ṫ 3 = 0 for x 2 = B0
u̇ = 0
3
Ṫ = 0
1
Ṫ = 0
2
for x3 = 0
u̇3 = U̇III Ṫ 1 = 0 Ṫ 2 = 0 for x 3 = C0 (2.2)
Here UI , UII and UIII are the uniform displacements on three sides of the cell, and the average
logarithmic strains in the three coordinate directions are 1 = ln(1+UI /A0 ), 2 = ln(1+UII /B0 )
and 3 = ln(1 + UIII /C0 ) respectively. The average true stresses in the three coordinate
directions are
 B0  C0
1
σ1 = [T 1 ]x 1 =A0 dx 3 dx 2
BC 0 0
 A0  C0
1
σ2 = [T 2 ]x 2 =B0 dx 3 dx 1 (2.3)
AC 0 0
 A0  B0
1
σ3 = [T 3 ]x 3 =C0 dx 2 dx 1
AB 0 0

where A, B and C are the current dimensions of the unit cell. In each increment U̇I is the
prescribed quantity and the values of U̇II and U̇III are determined such that the ratios of the
average principal stresses remain fixed
σ2 /σ1 = κ σ3 /σ1 = γ (2.4)
where κ and γ are prescribed constants.
In the fibres, elastic deformations are accounted for, following Hooke’s law. The values
of Young’s modulus and Poisson’s ratio for the fibres are denoted by Ef and νf . Furthermore,
no fibre breakage is accounted for here, and perfect bonding is assumed at the fibre–matrix
interface.

2.2. Ductile failure of the matrix material


An isotropic/kinematic hardening plasticity model for a porous ductile material is used to
describe the matrix material. This material model, suggested by Mear and Hutchinson [21]
and extended by Tvergaard [22] to account for void nucleation, makes use of a family of
isotropic/kinematic hardening yield surfaces of the form (σ ij , α ij , σF , f ) = 0, where f is
the current void volume fraction, σ ij is the average macroscopic Cauchy stress tensor and α ij
146 V Tvergaard

denotes the stress components at the centre of the yield surface. The radius σF of the yield
surface for the matrix material is taken to be given by
σF = (1 − b)σy + bσM (2.5)
where σy and σM are the initial yield stress and the matrix flow stress, respectively, and the
parameter b is a constant in the range zero to one. The constitutive relations are formulated
such that for b = 1 they reduce to Gurson’s [12] isotropic hardening model, whereas a pure
kinematic hardening model appears for b = 0.
The approximate yield condition for the porous solid is of the form
 k 
σ̃ 2 σ̃k
 = e2 + 2q1 f ∗ cosh − 1 − (q1 f ∗ )2 = 0 (2.6)
σF 2σF
where σ̃ ij = σ ij − α ij , σ̃e = (3s̃ij s̃ ij /2)1/2 and s̃ ij = σ̃ ij − Gij σ̃kk /3. For q1 = 1,
expression (2.6) is that proposed by Mear and Hutchinson [21], which coincides with that
of Gurson [12] for b = 1. The parameter q1 > 1 has been proposed to improve the agreement
with full numerical analyses for periodic arrays of voids [23, 24], while the function f ∗ (f )
in (2.6), of the form

f for f  FC

f (f ) = ∗
f − fC (2.7)
 fC + U (f − fC ) for f > fC
fF − f C
has been proposed [25] to model the process of void coalescence, where fF is the void volume
fraction at final fracture and fU∗ = f ∗ (fF ) = 1/q1 . The onset of coalescence and final failure
are taken to occur at the void volume fractions fC = 0.15 and fF = 0.25.
The plastic part of the macroscopic strain increment η̇ijP and the effective plastic strain
increment ˙MP
for the matrix material are taken to be related by [22]
σ̃ ij η̇ijP = (1 − f )σF ˙M
P
. (2.8)
Then, using the uniaxial true stress–natural strain curve for the matrix material, ˙M P
=
(1/Et − 1/E)σ̇M , an expression for the matrix flow stress increment σ̇M is obtained from (2.8).
Furthermore, the change of the void volume fraction during an increment of deformation is
taken to be given by
f˙ = (1 − f )Gij η̇ijP + Aσ̇M + B(σkk )• /3 (2.9)
where the first term results from the growth of existing voids and the last two terms model the
increment due to nucleation [26]. In particular, nucleation controlled by the plastic strain M
P
is
represented by taking A > 0 and B = 0 in (2.9), and is assumed to follow a normal distribution,
with the mean nucleation strain N , the standard deviation s and the volume fraction fN of void
nucleating particles [27]
   P 2

1 1 fN 1 M − N
A= − √ exp − B = 0. (2.10)
Et E s 2π 2 s
Alternative stress rates involving corotation with the crystal substructure spin (the elastic
spin) have been incorporated in a special version of the kinematic hardening constitutive
law [28] for the purpose of avoiding unrealistic oscillatory stress predictions. This alternative
formulation was not found to make any difference in [14], and will not be used here.
The plastic part of the strain rate is taken to be
1 G F ∇ kl
η̇ijP = m m σ (2.11)
H ij kl
3D analyses of ductile failure in metal reinforced by staggered fibres 147


where σ kl denotes the Jaumann rate of the Cauchy stress and the expressions for H and the
tensors mG
ij and mij are given in [22]. It is noted that plastic yielding initiates when  = 0
F

and  > 0 during elastic deformation, while continued plastic loading requires  = 0 and

(1/H )mFkl σ kl  0. The hardening rule, expressing the evolution of the yield surface centre
during a plastic increment, is taken to be

α kl = µ̇σ̃ kl µ̇  0 (2.12)
where the value of the parameter µ̇ is determined by the consistency condition,  ˙ = 0. The
incremental constitutive law of the form τ̇ ij = Lij kl η̇kl , to be used in the principle of virtual
work, is derived from the assumption that the total strain rate is the sum of the elastic and
plastic parts, η̇ij = η̇ijE η̇ijP .
These expressions are based on a Lagrangian formulation of the field equations, with
a material point identified by the coordinates x i in the reference configuration. The metric
tensors in the current configuration and the reference configuration are denoted Gij and gij ,
respectively, with determinants G and g, and ηij denotes the Lagrangian strain tensor. The
contravariant components of the Cauchy √ stress tensor σ ij and the Kirchhoff stress tensor τ ij
are related by the expression τ = G/gσ ij .
ij

The uniaxial true stress logarithmic strain curve for the matrix material, defining the value
of the tangent modulus Et , is taken to be represented by the piecewise power law
σ

 for σ  σy
E
= σ  σ 1/N (2.13)

 y
 for σ > σ y
E σy
where E is Young’s modulus, σy is the initial yield stress and N is the strain hardening exponent.

2.3. 3D numerical analysis


The Lagrangian strain tensor is given by
ηij = 21 (ui,j + uj,i + uk ,i uk,j ) (2.14)
i
in terms of the displacement components u on the reference base vectors, where ( ),j denotes
covariant differentiation in the reference frame. Furthermore, the equilibrium requirement is
expressed by the incremental principle of virtual work
    
{τ̇ ij δηij + τ ij u̇k ,i δuk,j } dV = Ṫ i δui dS − τ ij δηij dV − T i δui dS (2.15)
V S V S
where V and S are the reference volume and surface of the region analysed, T i are the specified
nominal surface tractions and the bracketed terms are the equilibrium corrections.
The cell model analyses carried out here are quite similar to previous 3D analyses [14, 15]
based on the same porous ductile material model. As in [14], the displacement increments u̇i
are approximated in terms of 3D 20-noded isoparametric elements, and the volume integrations
in (2.15) are carried out using a 2 × 2 × 2 point Gaussian quadrature within each element.
Figure 2 shows an example of a 12×9×6 element mesh used for computations with v1 = 0.075
and v2 = 0.055, i.e. a total fibre volume fraction of 0.13. Relatively crude elements are used
inside the elastic fibres, whereas the mesh outside the fibres is strongly graded, so that a fine
mesh appears adjacent to the fibres near the sharp fibre edges, where high stress and strain
peaks tend to develop.
Final ductile failure, when f reaches fF , can be represented numerically by different
procedures, as discussed in [14]. Here it is chosen to follow the constitutive model until
148 V Tvergaard

Figure 2. Mesh used for numerical analysis with v1 = 0.075 and v2 = 0.055. (a) Shows the cell
model surface at x 3 = C0 , while (b) shows the surface at x 3 = 0. The fibre elements are shown
hatched.

f = 0.9fF (at low stress levels near final failure), and subsequently to keep f constant (as
in [14]).
In the present cell model studies, localized damage near the fibre ends develops so abruptly
that there is no stable equilibrium path corresponding to monotonically increasing values of
the elongation UI , as is common in studies of matrix–fibre debonding [19, 20]. A Rayleigh–
Ritz finite element method [29] is used to prescribe an appropriate parameter, such that the
equilibrium solution can also be followed through mechanically unstable equilibria. It is
necessary here to be able to use this special method to prescribe an increasing value of the
difference between displacements on either side of a localized damage region, which occurs
typically at a fibre end, with the critical location changing throughout the failure process.

3. Results

For the model materials to be studied here, with the possibility of two different fibre sizes in
the unit cell, the volume fractions of the two different fibres are specified as v1 and v2 , but the
total fibre volume fraction is taken to be v1 + v2 = 0.13
√ in all cases. Also, the ratio of axial
and transverse spacings is in all cases given by A0 / 1/2B0 C0 = 4, but different values of
the ratio B0 /C0 are considered. The fibre aspect ratio in the case of equal size fibres is taken
to be 1 /r1 = 2 /r2 = 5.10, with 1 = 2 = 0.625A0 . When the values of v1 and v2 are
varied, the transverse fibre dimensions, specified by r1 and r2 , are kept fixed, while the values
of 1 and 2 are varied. For the matrix material the uniaxial stress–strain curve is specified
by taking σy /E = 0.0033 and N = 0.1. Poisson’s ratio is ν = 0.3, and q1 = 1.5 is used in
equation (2.6). The initial void volume fraction is taken to be fI = 0.0, and strain controlled
3D analyses of ductile failure in metal reinforced by staggered fibres 149

Figure 3. Stress–strain curves predicted for γ = κ = 0.4, i.e. for σm /σe = 1, with√different pairs of
fibres corresponding to the total fibre volume fraction v1 +v2 = 0.13, and with A0 / 1/2B0 C0 = 4,
B0 = C0 . Either isotropic hardening or kinematic hardening material models are used.

nucleation is assumed with the mean strain for nucleation, the standard deviation and the
volume fraction of void nucleating particles specified by N = 0.1, s = 0.05 and fN = 0.04,
respectively. For the fibres, the elastic material parameters are specified by Ef /E = 5.71 and
νf = 0.21, which is representative of SiC fibres in an aluminium matrix.
In figure 3 computed stress–strain curves are compared for a number √ of different materials
with the total fibre volume fraction v1 + v2 = 0.13, and with A0 / 1/2B0 C0 = 4, B0 = C0 .
In all cases the ratios of the transverse stresses to the major tensile stress are specified by
γ = κ = 0.4, according to (2.4), so that the overall stress triaxiality is σm /σe = 1. Three
of the stress–strain curves are computed for an isotropic hardening matrix material (b = 1),
while the last curve is obtained for kinematic hardening (b = 0). For the isotropic hardening
cases the difference is that one case has two identical fibres, while the other two cases have
different length and volume fraction of the two types of fibres. It is seen that the three curves
for an isotropic hardening matrix show very similar overall behaviour, even though details of
the failure evolution differ. When the two fibres are identical, v1 = v2 = 0.065, the onset
and evolution of failure at the ends of the two fibres are identical. However, when one fibre in
the unit cell is longer than the other, failure tends to initiate at the end of the longer fibre, and
ductile failure at this location may have resulted in complete separation of the matrix from the
fibre end before the first onset of failure at the end of the shorter fibre. This sequence of events
is clearly seen on the relevant curves in figure 3, where the characteristic sharp stress reduction
associated with the separation at a fibre end occurs twice on each curve. This behaviour is also
seen on the curve for a kinematic hardening matrix, where the two fibres differ significantly
in size, v1 = 0.075 and v2 = 0.055.
Since the total fibre volume fraction, the material parameters used for the matrix and fibres
and the overall stress triaxiality are identical to values used for some cases studied in [14],
the results are comparable. The main difference results from the fact that the simpler unit
cell in [14], containing one fibre, is representative of transversely aligned fibres, whereas the
present analyses consider transversely staggered fibres. As was shown in [30], the transversely
150 V Tvergaard

Figure 4. Contours of constant void volume


√ fraction for γ = κ = 0.4, i.e. for σm /σe = 1, with
b = 1, v1 = 0.075, v2 = 0.055 and A0 / 1/2B0 C0 = 4 with B0 = C0 . State of damage illustrated
for 1 = 0.0139. (a) x 3 /C0 0.03, (b) x 3 /C0 0.24, (c) x 3 /C0 0.76 and (d) x 3 /C0 0.97.

staggered array results in the prediction of lower stress levels, which appear to be in better
agreement with experiments. Also the curve for equal size transversely staggered fibres in
figure 3 shows first failure at a lower stress level and a larger strain level than found for the
corresponding case in [14]. The significantly earlier occurrence of ductile failure in the case of
a kinematic hardening material (b = 0), as seen in figure 3, is in good agreement with the trend
found in [14]. The kinematic hardening material model may be considered representative of a
solid that forms a rounded vertex on the yield surface, which makes the material significantly
less resistant to deviations from proportional stressing. The effect is particularly important
here, because the strongly localized plastic flow that develops in the matrix material along the
fibres, as well as the failure process near fibre ends results in non-proportional stressing.
The distribution of ductile matrix failure is illustrated in figure 4 by contours of constant
void volume fraction corresponding to a point on the curve for b = 1, v1 = 0.075 and
v2 = 0.055, i.e. the full curve in figure 3. The stage considered is at the overall strain
1 = 0.0139, which is just after the second sharp stress reduction on the full curve in figure 3,
associated with rapid evolution of ductile failure near the end of the shorter fibre. In each case
the plot is drawn on a surface parallel to element boundaries, which will not follow a constant
value of x 3 , as is seen in figure 2(c). The values of x 3 /C0 specified in the caption of figure 4
are precise for the part of the plot inside a fibre, and are thus also very good approximations
in the part of the matrix near the fibres, where non-zero values of f are observed. The plots in
figures 4(a) and 4(b) illustrate the damage distribution at two cross sections around the longer
fibre (shown hatched in figure 2(b)), while figures 4(c) and 4(d) show damage distributions at
two cross sections around the shorter fibre (hatched in figure 2(a)). Figure 4(c) also includes
a trace of the damage region around the end of the longer fibre. It is seen that the voids have
grown to full coalescence with loss of stress carrying capacity in the matrix material near the
end of the longer fibre, and that the material is also close to final failure near the end of the
shorter fibre. Also, some void nucleation and growth has occurred along the sides of the fibres,
where large plastic shear strains have developed.
3D analyses of ductile failure in metal reinforced by staggered fibres 151

√ 5. Stress–strain curves predicted for σm /σe = 1, with v1 = 0.075 and v2 = 0.055, and
Figure
A0 / 1/2B0 C0 = 4. Isotropic hardening material (b = 1).

Figure 5 shows some effects of non-uniform transverse stresses and of non-uniform


transverse fibre spacings for cases where the fibre shapes and volume fractions are kept fixed,
v1 = 0.075 and v2 = 0.055, and the stress triaxiality is fixed, σe /σm = 1, for an isotropic
hardening matrix material. Thus, the full curve in figure 5 is identical to the full curve in
figure 3. When the transverse stresses differ, as specified by γ = 0.320 and κ = 0.506, the
peak stress reached is lower and the first separation by ductile fracture near the end of the
longer fibre is somewhat delayed. When the transverse stresses differ even more, γ = 0.242
and κ = 0.805, the peak stress is much reduced and here none of the characteristic sharp stress
reductions are found, even though the ductile damage situation in the matrix material near the
ends of the two fibres reaches a state rather similar to that illustrated in figure 4. Thus, in this
particular stress state the void growth to coalescence is able to develop without the occurrence
of a mechanical instability, as observed in other cases. The curve for B0 = 2C0 represents
non-uniform fibre spacings in the two transverse directions, while the transverse stresses are
equal (γ = κ = 0.4), and it is seen that this has an effect rather similar to that of moderately
non-uniform transverse stresses with uniform transverse spacings.
The effect of the mesh size is studied in figure 6 by comparing predictions for the 12×9×6
mesh used in the previous figures (see also figure 2) with predictions obtained by a cruder
12 × 8 × 4 mesh. One of the cases considered is that with two identical fibres from figure 3,
and the other case is that with the largest non-uniformity of the transverse stresses from figure 5.
It is seen in figure 6 that in both cases the agreement is reasonably good. Using only 8 × 4
elements over the cell cross section is a significant reduction relative to the mesh shown in
figure 2(c), but it is noted that the cruder mesh has been stretched so that the sizes of matrix
elements closest to the fibres are rather similar to those in the finer mesh.
The effect of the stress triaxiality is considered in figure 7 by comparing predictions for
values of σm /σe equal to 1, 2 and 3. The volume fractions of the two fibres are v1 = 0.075
and v2 = 0.055, so that the full curve in figure 7 is identical to the full curves in figures 3
and 5. The effect of increasing stress triaxiality is very similar to that found in [14] for the
transversely aligned array of short fibres. Thus, higher stress triaxiality results in less overall
plastic flow prior to the occurrence of matrix failure at the end of the longer fibre. All three
curves clearly show the two separate sharp stress reductions, the first corresponding to matrix
failure near the end of the longer fibre and the second to the later matrix failure near the end of
152 V Tvergaard

curves predicted for σm /σe = 1, with different pairs of fibres such that
Figure 6. Stress–strain √
v1 + v2 = 0.13, and A0 / 1/2B0 C0 = 4 with B0 = C0 . Isotropic hardening material (b = 1).

Figure 7.√Stress–strain curves predicted for various stress triaxialities, with v1 = 0.075, v2 = 0.055
and A0 / 1/2B0 C0 = 4 with B0 = C0 . Isotropic hardening material (b = 1).

the shorter fibre. After these two sharp stress reductions a more gradual decay of the overall
stress is seen, associated with the development of final void coalescence in the matrix material
at the fibre ends, as f grows from the critical value fC to the value fF at final fracture.
The stress–strain curves in figure 8 compare different cases for γ = κ = 0.625. The full
curve in figure 8 is identical to the curve for σm /σe = 2 in figure 7, and the other two curves
represent a case with fibre spacings that differ in two transverse directions (B0 = 2C0 ) and a
case with equal spacings but with a kinematic hardening matrix material (b = 0). The different
fibre spacings give a trend similar to that found in figure 5 for σm /σe = 1, i.e. a slightly lower
stress level and a small delay in the first onset of a sharp stress reduction. The effect of a
3D analyses of ductile failure in metal reinforced by staggered fibres 153

√ curves predicted for γ = κ = 0.625, i.e. for σm /σe = 2, with v1 = 0.075,


Figure 8. Stress–strain
v2 = 0.055 and A0 / 1/2B0 C0 = 4. Either isotropic hardening or kinematic hardening material
models are used.

rounded vertex on the yield surface, as represented by kinematic hardening, is that the first
onset of failure occurs earlier and that failure develops more rapidly, quite similar to the effect
of kinematic hardening found in figure 3 for σm /σe = 1.

4. Discussion

Three-dimensional cell model studies for ductile matrix failure in a metal reinforced by parallel
short fibres are here extended, relative to the analyses in [14, 15], so that a transversely staggered
array of fibres can be accounted for. In [30] an axisymmetric cell model with special boundary
conditions was developed to represent such transversely staggered fibres, and it was found
that the resulting smaller constraint on plastic flow in the matrix gives a better agreement with
experiments for real fibre distributions. Also the present full 3D cell model results, compared
to those in [14] for transversely aligned fibres, show that the average stress levels are generally
reduced and that the first onset of ductile failure in the matrix near a fibre end occurs at a larger
average strain.
The present cell model also allows for investigating effects of different sizes of two
neighbouring fibres. The tendency found is that ductile failure occurs first near the ends
of the longer fibre, developing into full separation by void coalescence before failure starts
near the ends of the shorter fibre. Corresponding to this sequence of events the predicted stress–
strain curve for the composite shows a characteristic sharp stress reduction when failure occurs
near the ends of the longer fibre and a secondary sharp stress reduction when failure occurs
at the shorter fibre. The average strains, at which the sharp stress reductions occur, change a
little when the fibre spacings or the average principal stresses differ a little in two transverse
directions. However, for significantly different stresses in two transverse directions, a case has
been found where the successive failure evolution near the ends of the different length fibres
occurs without any sharp stress reductions (figure 5).
154 V Tvergaard

The predictions for a kinematic hardening matrix material (b = 0) in figures 3 and 8


are significant because they illustrate the strong sensitivity to details of the elastic–plastic
constitutive law. The kinematic hardening model has been used in a number of previous
studies to represent a material that forms a rounded vertex on the yield surface at the point
of loading, which appears to be quite realistic for real materials. The present comparisons of
predictions for kinematic hardening with those for isotropic hardening clearly show that an
increased local curvature of the yield surface gives a significantly earlier occurrence of ductile
failure. It is emphasized that predictions of the kinematic hardening model coincide with those
of the isotropic hardening model for proportional loading, so the strong difference found here
depends on the deviation from proportional loading in local regions of the matrix material.
As in [14] the present computations have been continued far beyond the point where
dσ1 /d 1 → −∞ occurs first, i.e. where the solution is mechanically unstable even for
prescribed end displacements of the unit cell. Previous axisymmetric computations [7, 11]
for ductile matrix failure focused on the behaviour before this point of instability. The present
3D solutions could only be continued through the unstable regime using a special method to
prescribe an increasing displacement difference over the localized damage region, as described
in section 2.3. It is noted that the failure evolution found in this unstable regime is somewhat
similar to the pattern of failure predicted in studies accounting for fibre–matrix debonding,
see for example [19]. Thus, the present analyses show full separation by ductile failure in
the matrix material near the fibre ends and subsequently ductile failure starts to stretch down
along the sides of the fibres (figure 4), while the analyses of interface failure show the first
debonding at the fibre ends and then a gradual separation by tangential debonding along the
fibre sides as the fibre pull-out process starts.
It should be noted that the porous ductile material models used for the present studies
of ductile matrix failure cannot be expected to give a very accurate description of the failure
process along the fibre sides. The models are based on assuming that voids remain spherical
during growth, which is a good approximation in high triaxiality regions as that near the
fibre end, but in the highly shear dominated regions along fibre sides voids may tend to
elongate in the shear direction during growth, and this would not be very well represented by
the material models. Another implicit assumption in using a porous ductile material model
around discretely represented fibres is that the void size is much smaller than the fibre size,
which is not always true for metal matrix composite materials (e.g. see [3]). Furthermore,
effects of strain gradients are not included in the present analyses, but such effects would tend
to somewhat stiffen the response when voids or fibres are very small (see discussions in [31]).

References

[1] Divecha A P, Fishman S G and Karmarker S D 1981 J. Met. 33 12


[2] McDanels D L 1985 Metall. Trans. A 16 1105
[3] Needleman A, Nutt S R, Suresh S and Tvergaard V 1993 Fundamentals of Metal Matrix Composites ed S Suresh,
A Mortensen and A Needleman (Boston, MA: Butterworth–Heinemann) p 233
[4] Davidson D L 1987 Report N00014-85-C-0206 06-8602/3 Southwest Research Institute
[5] You C P, Thompson A W and Bernstein I M 1987 Scr. Metall. 21 181
[6] Christman T and Suresh S 1988 Mater. Sci. Eng. A 102 211
[7] Llorca J, Needleman A and Suresh S 1991 Acta Metall. Mater. 3 2317
[8] Zok F, Embury J D, Ashby M F and Richmond O 1988 Mechanical and Physical Behaviour of Metallic and
Ceramic Composites ed S I Andersen et al (Roskilde, Denmark: Risø National Laboratory) p 517
[9] Lagacé H and Lloyd D J 1989 Can. Metall. Q 28 145
[10] Mummery P and Derby B 1991 Mater. Sci. Eng. A 135 221
[11] Llorca J, Suresh S and Needleman A 1992 Metall. Trans. A 23 919
[12] Gurson A L 1977 J. Eng. Mater. Technol. 99 2
3D analyses of ductile failure in metal reinforced by staggered fibres 155

[13] Tvergaard V and Needleman A 1995 Int. J. Solids Structures 32 1063


[14] Tvergaard V 1998 Acta Mater. 46 3637
[15] Tvergaard V 1998 Modelling of Structure and Mechanics of Materials from Microscale to Product
ed J V Carstensen et al (Risø, Denmark: Risø National Laboratory) p 523
[16] Bao G, Hutchinson J W and McMeeking R M 1991 Acta Metall. Mater. 39 1871
[17] Sørensen N J, Needleman A and Tvergaard V 1992 Mater. Sci. Eng. A 158 129
[18] Sørensen N J, Suresh S, Tvergaard V and Needleman A 1995 Mater. Sci. Eng. A 197 1
[19] Tvergaard V 1990 Mater. Sci. Eng. A 125 203
[20] Tvergaard V 1993 J. Mech. Phys. Solids 41 1309
[21] Mear M E and Hutchinson J W 1985 Mech. Mater. 4 395
[22] Tvergaard V 1987 J. Mech. Phys. Solids 35 43
[23] Tvergaard V 1981 Int. J. Fracture 17 389
[24] Tvergaard V 1982 Int. J. Fracture 18 237
[25] Tvergaard V and Needleman A 1984 Acta Metall. 32 157
[26] Needleman A and Rice J R 1978 Mechanics of Sheet Metal Forming ed D P Koistinen et al (New York: Plenum)
p 237
[27] Chu C C and Needleman A 1980 J. Eng. Mater. Technol. 102 249
[28] Tvergaard V and van der Giessen E 1991 J. Mech. Phys. Solids 39 763
[29] Tvergaard V 1976 J. Mech. Phys. Solids 24 291
[30] Tvergaard V 1990 Acta Metall. 38 185
[31] Fleck N A and Hutchinson J W 1997 Adv. Appl. Mech. 33 295

You might also like