Vous êtes sur la page 1sur 19

Engineering Fracture Mechanics 74 (2007) 1203–1221

www.elsevier.com/locate/engfracmech

Role of stress triaxiality in elastoplastic characterization


and ductile failure prediction
G. Mirone *

Dipartimento di Ingegneria Industriale e Meccanica, Università di Catania, Italy

Received 20 June 2005; received in revised form 27 May 2006; accepted 7 August 2006
Available online 11 December 2006

Abstract

The triaxiality of the stress state is known to greatly influence the amount of plastic strain which a material may
undergo before ductile failure occurs.
During tensile load histories, the necking induces significant stress triaxiality modifications which in turn affect the
experimental stress–strain measurements needed for the characterization of ductile metals.
In this paper, the recently proposed ‘‘MLR’’ model of necking effect is used to obtain the flow curves of various metals
by correcting the experimental data of tensile tests. Finite elements simulations of the experimental tests are performed to
calculate the stress triaxiality evolution on various notched and unnotched specimens. A ductile failure criterion, due to
Bao and Wierzbicki, is then applied to evaluate the material damage and predict failure. This procedure is applied to a
set of 20 specimens series made of six metals with 10 different notch shapes.
The damage calculations also indicate the material points where failure initiates. These predictions are confirmed by
micrographic observation of voids on polished fragments of the broken specimens.
 2006 Elsevier Ltd. All rights reserved.

Keywords: Necking; Flow curve; Stress triaxiality; Damage; Ductile failure

1. Introduction

The stress–strain characterization of elastoplastic ductile metals and the prediction of their failure are two
aspects which, tough being extensively investigated in the last decades, are still of great actuality because fur-
ther details are to be explained and a commonly accepted viewpoint is not yet achieved.
This paper describes the application of the MLR model for the solution of the necking problem and the
Bao and Wierzbicki model for ductile failure, used together with finite element analyses for simulating various
experimental tensile tests also performed.

*
Tel.: +39 95 7382418; fax: +39 95 330258.
E-mail address: gmirone@diim.unict.it

0013-7944/$ - see front matter  2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.engfracmech.2006.08.002
1204 G. Mirone / Engineering Fracture Mechanics 74 (2007) 1203–1221

Nomenclature

F, rTrue, eTrue current tensile load, True stress, True strain


a0, a initial and current radius of the minimum cross-section
r radial coordinate on the neck section
rz, rr, rh, rH, rEq axial, radial, hoop, hydrostatic and equivalent stress
ez, er, eh, eN, eEq axial, radial, hoop, necking-initiation and equivalent strain
eF, ef local failure strain, logarithmic (neck-averaged) failure strain
MLRr, MLRe corrective function for stresses, corrective function for strains
TF triaxiality ratio
D, Dcr damage function, critical value of the damage function

The MLR model applied to experimental data supplies the stress–strain curves needed to define the mate-
rials in the finite element analyses, and the Bao–Wierzbicki damage model uses the FE results for predicting
the failure of specimens.
The distributions of principal stresses and strains on the necked cross-section, calculated by the MLR
method and by finite elements, are also compared.
Damage model predictions about material points where failure initiates are then verified by micrographic
observation of polished fragments cut from the broken specimens.
The most known necking model is due to Bridgman [1], it is based on the approximate hypothesis of uni-
form strain distributions and relates the current Mises stress to the current true stress and the current necking
profile radius. It is the almost official model in this field, its accuracy ranges from 5% to 15% and requires
considerable experimental efforts in acquiring many necking profile shapes on a deforming specimen and cal-
culating their curvature radii. Other known models are due to Clausing [2] and to Alexandrov et al. [3] whose
model incorporate a damage variable. Earl et al. [4] derived the stress and strain distributions on the neck by
assuming a hyperbolic sine function for the radial displacements, then evaluating by experiments the approxi-
mations of the Bridgman method. Also Alves et al. [5] and La Rosa et al. [6], found that large errors, up to
15%, are possible when the Bridgman formulae are used to predict the stresses and strains on the minimum
cross-section of necked tensile specimens.
The MLR model of the necking effect, recently proposed in [7], is based on the material-independency of
the necking-induced modifications of the stress state, related to an opportunely reduced strain. This model
does not require any experimental measurement exceeding those typically required for the true stress–true
strain curve, and enables to calculate the Mises stress with an accuracy within 5%.
The other aspect of elastoplastic behavior of metals still needing a comprehensive description is the ductile
failure, for which a wide variety of explanations has been proposed in the literature. The basis of every failure
criterion is the evidence that failure depends on the evolution of the equivalent plastic strain and of the triax-
iality ratio. The importance of their combined effect was initially documented by McClintock [8] and Rice and
Tracey [9], who calculated the enlargement of cylindrical and spherical voids ordered in arrays within a
surrounding metal matrix. The great number of studies on ductile failure can be roughly subdivided in the
groups of the continuous damage mechanics, the Gurson-derived models and the uncoupled models based
mainly on energetic considerations. The first and probably most investigated approach is due to Lemaitre
[10–13] and Chaboche [14,15] and is based on a damage potential function expressing the loss of elastic
stiffness which occurs in a damaged-voided solid. Many CDM models have been proposed by incorporating
modifications of the Lemaitre’s potential, such as in [16–18] which have been compared each other and with
experimental data in [19]. Gurson is the precursor of the second group of ductile failure models and defined, in
[20], a yielding function for porous solids which includes the void volume fraction as a damage variable. It is a
fully coupled approach handled almost exclusively by finite elements, and allows to simulate the evolution of a
pre-existing porosity. Tvergaard and Needleman [21–23] presented an upgrade of the Gurson model, in which
the modeling of the void nucleation and coalescence is added by way of opportune functions. Other modifi-
cations of the Gurson model are those due to Ragab et al. [24], Saie et al. [25]. In the third group can be
G. Mirone / Engineering Fracture Mechanics 74 (2007) 1203–1221 1205

included the models based on the Rice-Tracey void growth rate such as [26–28], or those directly expressing
the constancy of the energy associated to the damage work as in [29–31]. Another point of view is proposed by
Brown, Hancock and Mackenzie in [35,36], where the failure domain is experimentally extrapolated as a func-
tion of the initial stress triaxiality and of the corresponding failure strain, for differently notched specimens.
The damage calculations in this work are performed according to the model of Bao and Wierzbicki [32,33],
based on the finding that the triaxiality ratio on every material point, integrated over the plastic strain up
to failure, is a material constant.

2. Experimental tensile tests

Tensile tests are conducted on unnotched and notched specimens made of different materials, according to
Fig. 1 and Table 1.
Each series of specimens is univocally identified by the name of its material and the curvature radius R of its
notch in the undeformed state; a number of identical specimens ranging from 2 to 4 is tested for each series.
Tensile tests are performed at nominal strain rates between 1 and 8 · 103 s1. The data for copper 99.97%
are extracted from Ref. [34].

L'
d D

R
L

Fig. 1. Specimen shape.

Table 1
Nominal dimensions of the specimens (see Fig. 1)
Material R (mm) d (mm) D (mm) L 0 (mm) L (mm)
Aluminium 2011 1 – 9 – 55
10 6 9 10.5 55
5 6 9 7.1 55
Copper 99.9% 1 – 9 – 55
10 6 9 10.5 55
5 6 9 7.1 55
Steel AISI T304 1 – 9 – 55
30 6 9 18.7 55
15 6 9 13.1 55
5 6 9 7.1 55
2 6 9 3.9 55
Steel AISI 1040 1 – 9 – 50
20 6 9 15.2 50
10 6 9 10.5 50
Steel ASTM A284 1 9 – 50
20 6 9 15.2 50
10 6 9 10.5 50
Copper 99.97% Ref. [34] 1 – 6 – 60
4 4 9 7.4 60
1 4 9 2 60
1206 G. Mirone / Engineering Fracture Mechanics 74 (2007) 1203–1221

The local strain rate in the necked section increases substantially with respect to the nominal value, but
results for the same metal and notch geometry pulled at different speeds confirmed that all rates are low
enough to justify neglection of strain rate effects. During each tensile test, a digital video sequence, centred
either on the notch region or on the entire gauge length is acquired. Single frames are extracted a posteriori
from the video sequences at fixed time intervals and processed by image analysis, for measuring the current
radius of the minimum cross-section, a = d/2.
By matching a load value F recorded by the testing machine with a radius a from the image at the same
instant of the test, true stress and true strain are calculated as in (1)
F
rTrue ¼
p  a2 ð1Þ
eTrue ¼ 2  Lnða0 =aÞ

where a0 is half the diameter of the initially undeformed cross-section.


This procedure is repeated many times for each specimen, obtaining as many points of its true curve as
reported in Fig. 2 for three of the six tested metals.
The continuous curves in Fig. 2, indicated as true curves, are plots of the least squares fitting polynomials
approximating the experimental true stress–true strain points for the smooth specimens. These polynomials,
also indicated as true stress functions or rTrue(eTrue), are reported in Table 2 for all metals tested, together with
other data from experiments, while the average failure strains for each specimen series are listed in Table 3.
The true stress function for the annealed copper 99.97% is obtained by backward-application of the MLR
correction method described in the next section to the flow curve proposed in [34]. In the small strain ranges
successive to the first yield and to necking, the piecewise laws of Table 2 are connected by short linear ramps.

3. Stress–strain characterization of metals

If cyclic loading does not occur and the deviation from proportional loading is not too large, the isotropic
hardening alone is usually considered for the stress–strain characterization of elastoplastic metals; it is
expressed as a scalar-to-scalar function relating the equivalent von Mises stress rEq to the equivalent plastic
strain eEq:
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
iffi
1h 2 2 2
rEq ¼ ðrz  rr Þ þ ðrz  r# Þ þ ðr#  rr Þ
2
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð2Þ
2 2 2 2
eEq ¼ ½e þ er þ e# 
3 z
where the subscripts z, r and h indicate the axial, radial and hoop directions in the principal coordinate system
of cylindrical tensile specimens.
The simplest way to calculate rEq and eEq from experiments requires the adoption of smooth cylindrical
specimens undergoing tensile loads, so that the stress is uniaxial and uniform all over the volume of interest
in the specimen.
The occurrence of necking modifies the stress state making it neither uniaxial nor uniform, thus, at post-
necking deformation levels, the flow curve may largely differ from the experimental true curve.
A method for the exact calculation of the equivalent stress and plastic strain under necking conditions does
not exist, but Bridgman’s approximate correction method [1] is currently the most widely used one in the lit-
erature. The MLR model [7], used in this paper for the post-necking characterization, is more accurate and
easily manageable than the former, it is based on the trends shown by the ratios between two couples of stres-
ses and strains, each averaged onto the current neck section. If the reduced post-necking strain eEq  eN is
taken as the variable governing these ratios, then the two relationships are found to be material-independent
for a wide variety of ductile metals, as shown in Fig. 3.
The fitting curves in Fig. 3 express the functions MLRr(eEq  eN) and MLRe(eEq  eN).
The MLRr function alone allows to derive the flow curve of a material by simply multiplying the true stress
polynomial of unnotched specimens by this function; If used in combination, the two MLR functions
G. Mirone / Engineering Fracture Mechanics 74 (2007) 1203–1221 1207

Aluminium 2011
600

500

True stress [MPa]


400
Unn
300
R10

200 R5

100 Fit

0
0 0.1 0.2 0.3 0.4 0.5 0.6
ε True = 2 Ln(a0 /a)

Copper 99.9%
600

500
True stress [MPa]

400

300 Unn
R10
200
R5
100
Fit

0
0 0.2 0.4 0.6 0.8 1 1.2
ε True = 2 Ln(a0 /a)

STEEL ASTM A284


1000
900
800
True stress [MPa]

700
600 Unn
500 R20
400
R10
300
Fit
200
100
0
0 0.2 0.4 0.6 0.8 1
ε True = 2 Ln(a0 /a)
Fig. 2. Samples of experimental true stress–true strain curves for three metals.

constitute two mathematical constraints which, together with the typical equations of plasticity and with two
more hypotheses on the axial strain and the hydrostatic stress, allow to calculate the complete distributions of
stress and strain on the current neck section of a smooth specimen.
1208 G. Mirone / Engineering Fracture Mechanics 74 (2007) 1203–1221

Table 2
Experimental data from unnotched specimen
True curve fitting (smooth specimens) (MPa) Yield stress rY (MPa) Elastic modulus (MPa) Necking strain eN
Aluminium 2011 545 e0.15; for 0 < e < 0.15 143 73000 0.15
383 + 265e; for e > 0.18
Copper 99.9% 327 + 210e; for e > 0.06 180 130,000 0.03
0.25
AISI T304 1183e ; for 0.02 < e < 0.25 190 209,000 0.25
693 + 592e; for e > 0.28
AISI 1040 417 + 351e; for e > 0.05 410 206,000 0.02
ASTM A284 534 + 389e; for e > 0.05 360 205,000 0.02
0.43
Copper 99.97% 517e ; for e < 0.2 45 121,000 0.32
450e0.32/LRr; for e > 0.23

Table 3
Experimental logarithmic strains at failure
Aluminium 2011 Copper 99.9% Steel AISI T304 Steel AISI 1040 Steel ASTM A284 Copper 99.97%
Unn 0.52 Unn 0.96 Unn 1.45 Unn 1.08 Unn 0.99 Unn 1.23
R10 0.35 R10 0.63 R30 1.35 R20 0.92 R20 0.82 R4 0.95
R5 0.28 R5 0.55 R15 1.29 R10 0.8 R10 0.73 R1 0.76
R5 1.14
R2 1.03

1 AISI 304 Steel DM 1


1035 Le Roy
0.9 0.9
C40 1045 Le Roy
0.8 0.8
Fe 36 1090 Le Roy
σeq Avg / σZ Avg

0.7 0.7
Allum. DM (400 p)
εr Avg / εθ Avg

1015 Le Roy
0.6 2
0.6 MLR ε(εeq, εΝ) = 1 - 0.265 (εeq- εN) +
MLR σ(εeq, εN) = 1 - 0.6058 (εeq- εN) + 1090 Le Roy HY 130
0.5 3 4
0.5 2 3
+0.6317 (εeq- εN) - 0.2107 (εeq- εN) 1045 Le Roy +0.241 (εeq- εN) - 0.074 (εeq- εN) Q1 LT
0.4 0.4
1035 Le Roy SWEDISH IRON
0.3 0.3
1015 Le Roy C 40
0.2 0.2
HY130 Fe 36
0.1 0.1
εeq-εN εeq-εN
Q1 LT D 98
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1

Fig. 3. Material-independent ratios between averaged neck stresses and strains.

Details about these calculations are reported in Appendix. Only the final expression of the equivalent stress
estimated by the MLRr correction is presented here as:
F
rEq ðeEq Þ ¼ rTrue ðeTrue Þ  MLRðeTrue  eN Þ ¼  MLRðeTrue  eN Þ ð3Þ
p  a2
Eq. (3), applied to the function rTrue(eTrue) from unnotched specimens, yields the equivalent stress function
rEq(eEq) for each material. The latter functions are plotted in Fig. 4 (material flow curves marked by symbols
for identification), together with the rTrue(eTrue) functions (true stress–true strain curves, unmarked) that origi-
nated them, in order to show how significant is the departure of each couple of these curves after necking
initiation.
Fig. 4 shows that, at failure, the equivalent stress may differ greatly from the true stress, so that it is impos-
sible to reasonably perform finite elements simulations at high strain levels or failure analyses, if the true
stress–true strain curve (expressing rTrue(eTrue)) is assumed to play the role of the material flow curve (express-
ing rEq(eEq)), without any correction.
The flow curves in Fig. 4 are then used to define each metal in FE analyses simulating all the experimental
tensile tests performed.
G. Mirone / Engineering Fracture Mechanics 74 (2007) 1203–1221 1209

True & Flow Curves


1600
Aluminium 2011 σTrue
1400 Aluminium 2011 σEq
Copper 99.9% σTrue
1200
Copper 99.9% σEq
Stress [MPa]

1000 AISI T304 σTrue


AISI T304 σEq
800
AISI1040 σTrue
600 AISI 1040 σEq
ASTM A284 σTrue
400
ASTM A284 σEq
200 Copper 99.97% σTrue
Copper 99.97% σEq
0
0 0.15 0.3 0.45 0.6 0.75 0.9 1.05 1.2 1.35 1.5
ε True = 2 Ln(a0 /a)

Fig. 4. Flow curves and true stress–true strain curves.

4. Finite element analyses

Finite element (FE) models of the tensile specimens are assembled with 3500–5000 eight-noded axisymmet-
ric elements for the different notch geometries. Only one quarter of each specimen is modelled due to symme-
try, as shown in Fig. 5.
Large displacements, finite plastic strain, updated Lagrangian formulation and additive decomposition
options are activated in the MARC code. The imposed specimens elongations are subdivided in 600–1500
steps, depending on material and geometry.
In order to verify the accuracy of the material flow curves and of the FE analyses, a comparison is per-
formed between the experimental true curves and the same curves obtained as an output from the FE results.
The reaction load on the rear cross-section of each FE model and the radial displacement of the outer node in
the minimum cross-section are read at each analysis step on the deforming mesh, so that true stress and true
strain are calculated by Eq. (1).
Samples of the comparison between experiments and finite elements results are shown in Fig. 6, for the
specimens made of three materials among those investigated.
The error in rTrue(eTrue) predicted by FE compared to experiments for unnotched specimens at failure is 4%,
<1%, 3%, <1%, <1% and 2% for Cu 99.9%, AISI T304, Aluminium 2011, AISI 1040, ASTM A284 and Cu
99.97%, respectively. For the sharper notches of the same materials, the FE error is 2%, <1%, 2%, <1%,
2% and 6%.
The agreement between the true curves from experiments and those from the FE results, allows taking the
numerical results as an acceptable accurate reference including those variables which cannot be directly
measured by experiments, such as the distributions of local stress and strain on the minimum cross-section.

Fig. 5. Mesh of R5 specimens and post-necking shape of AISI T304 unnotched model.
1210 G. Mirone / Engineering Fracture Mechanics 74 (2007) 1203–1221

AISI T304 F.E. - EXP. COMPARISON


1800

1600

1400

True stress [MPa]


1200

1000

800
Unn R 30
600
R 15 R5
400

200 R2 F.E.

0
0 0.3 0.6 0.9 1.2 1.5
εTrue = 2Ln(a0/a)
AISI 1040 F.E. - EXP. COMPARISON
900

800

700
True stress [MPa]

600
Unn
500
R 20
400
R 10
300
F.E.
200

100

0
0 0.2 0.4 0.6 0.8 1 1.2
εTrue = 2Ln(a0/a)
Copper 99.97% F.E. - EXP. COMPARISON
600

500
True stress [MPa]

400

300 Unn
R4
200
R1
F.E.
100

0
0 0.2 0.4 0.6 0.8 1 1.2
εTrue = 2Ln(a0/a)
Fig. 6. Validation of flow curves for AISI T304, AISI 1040 and copper 99.97%.
G. Mirone / Engineering Fracture Mechanics 74 (2007) 1203–1221 1211

5. Stress, strain and triaxiality ratio at the neck section

The stress and strains at the cross-section of the unnotched specimens, calculated according to the MLR
model with Eqs. (A1)–(A8) in Appendix, are compared with the FE results;

Aluminium 2011 Strain at failure Aluminium 2011 Stress at failure


0.60 700
FEM axial FEM axial
0.50 600
FEM radial
0.40 FEM radial
500 FEM hoop
0.30 FEM hoop
400 FEM Equiv

Stress [MPa]
Strain

0.20
FEM Equiv 300 FEM Hydr
0.10
r/a MLR Axial 200 MLR Axial
0.00
0.0 0.2 0.4 0.6 0.8 1.0 MLR Radial
MLR Radial 100
-0.10
MLR Hoop
-0.20 0
MLR Hoop 0.0 0.2 0.4 0.6 0.8 1.0 MLR Hydr.
-0.30 -100
MLR Equiv r/a MLR Equiv.
-0.40 -200

Cu 99.9% Strain at failure Cu 99.9% Stress at failure


1.20 800
FEM axial FEM axial
1.00 700
FEM radial
0.80 FEM radial 600
FEM hoop
0.60 FEM hoop 500
FEM Equiv
Stress [MPa]
Strain

0.40 400
FEM Equiv FEM Hydr
0.20 300
MLR Axial MLR Axial
0.00 r/a 200
0.0 0.2 0.4 0.6 0.8 1.0 MLR Radial
-0.20 MLR Radial 100
MLR Hoop
-0.40 MLR Hoop 0
0.0 0.2 0.4 0.6 0.8 1.0 MLR Hydr.
-0.60 -100
MLR Equiv r/a MLR Equiv.
-0.80 -200

AISI T304 Strain at failure AISI T304 Stress at failure


2.00 2500
FEM axial FEM axial

1.50 FEM radial


FEM radial 2000
FEM hoop
FEM hoop
1.00 1500 FEM Equiv
Stress [MPa]
Strain

FEM Equiv FEM Hydr


0.50 1000
MLR Axial MLR Axial
r/a
0.00 MLR Radial
MLR Radial 500
0.0 0.2 0.4 0.6 0.8 1.0
MLR Hoop
MLR Hoop
-0.50 0 MLR Equiv.
0.0 0.2 0.4 0.6 0.8 1.0
MLR Equiv r/a MLR Hydr.
-1.00 -500

AISI 1040 Strain at failure AISI 1040 Stress at failure


1.40 1200
FEM axial FEM axial
1.20
1000 FEM radial
1.00 FEM radial
FEM hoop
0.80 800
FEM hoop
FEM Equiv
Stress [MPa]

0.60
Strain

0.40 FEM Equiv 600 FEM Hydr

0.20 MLR Axial MLR Axial


400
0.00 r/a
MLR Radial
0.0 0.2 0.4 0.6 0.8 1.0 MLR Radial
-0.20 200
MLR Hoop
-0.40 MLR Hoop
0 MLR Equiv.
-0.60 0.0 0.2 0.4 0.6 0.8 1.0
MLR Equiv r/a MLR Hydr.
-0.80 -200

Fig. 7. Stresses and strains on the neck of AISI T304, AISI 1040, ASTM A284 and copper 99.97% unnotched specimens at failure.
1212 G. Mirone / Engineering Fracture Mechanics 74 (2007) 1203–1221

a b σz 2000
1500
1000
600 500
0
σz 400
1000

200
750
0 r=0
100
σθ 500
50 r=0
σθ r=a
0
250 r=a
-50
100 80 60 40 20 0 0
1000 750
σr 500 250 0
σr

Fig. 8. Stress paths on the neck center and the neck outer radius for Cu 99.97% unnotched and AISI T304 Notch R2.

The comparisons are carried out at the true strain values corresponding to the incipient failure, as shown in
Fig. 7 for aluminium, copper and two different steels. For all materials, the distributions of eZ are almost coin-
cident and with those of eEq and not visibly distinguishable from them, in agreement with the findings in [5,7].
The accuracy of the MLR model in predicting the strains and the von Mises stress rEq all over the neck
section is remarkable (average error lower than 2%, max. 4%).
The hydrostatic stress rH, on the neck center exhibits an error ranging from 1% to 9% (5.5% average), and
this induces similar error levels on the principal stresses but does not affect rEq, which is insensitive to rH.
On the outer circumference of neck section (r = a), rz and rh are prone to a lower degree of accuracy, most
likely because the radial equilibrium in Eq. (A8) is enforced just locally, by imposing that here the radial stress
is nil, rather than by way of an indefinite equation applying all over the radial abscissa.
It is worth noting that, for calculating stresses and strains at the neck, the MLR method requires just the
parameters fitting the experimental post-necking true stress (usually two as this part of the true curve is almost
linear), the necking strain and the current diameter reduction. Therefore, the MLR method gives a good
insight of what happens on the neck section, much more easily and rapidly than a finite elements analysis.
The knowledge of local stresses on the neck section from f.e. results, allows to evaluate the paths each mate-
rial point describes in the stress or strain spaces during the entire deformation history; now it is also possible to
evaluate how the stress state triaxiality, quantified by the ratio TF of Eq. (4), evolves after necking initiation.
rH
TFðrÞ ¼ ð4Þ
rEq
The stress paths are calculated at the center (r = 0) and at the outer radius (r = a) of the minimum cross-sec-
tion of all the specimens, as reported in Fig. 8 for the unnotched specimens of 99.97% Cu and the R2-notched
specimens of AISI T304.
In the unnotched specimen before necking occurrence, all material points are subjected to simple uniaxial
tension and the initial stress path up to the necking initiation is the straight segment common to the curves for
r = 0 and r = a in Fig. 8a. After necking initiates, the stress paths on these two material points start to diverge
each other, for r = 0 the stress state becomes increasingly triaxial, but, for r = a, the radial stress is constantly
null (biaxial stress state) and the stress path lies on the plane rr = 0.
For notched specimens, the stress and strain nonuniformity is a feature acting already in the elastic phase (lin-
ear part of the curves b), and the two paths have no points in common except for that indicating the initial zero-
load state. The stress path of points lying on the outer perimeter of the neck is again evolving on the plane rr = 0.

6. Cumulated triaxiality ratio and local failure

When tension is the prevailing stress, ductile failure initiates at a single material point where the void vol-
ume fraction or a related damage function firstly reaches a critical value.
G. Mirone / Engineering Fracture Mechanics 74 (2007) 1203–1221 1213

The function whose critical value triggers the local failure depends on TF and eEq, because these variables
are known to promote void growth and hence ductile failure; The main plots in Fig. 9 show the histories of TF
vs. eEq for the same two specimen series considered in Fig. 8. The values of triaxiality and strain are both local,
referred to the neck center and to a point on the outer perimeter of the neck section, and are calculated from
the FE results as previously validated.
The black round symbols in Fig. 9 indicate the FE predictions at an increment for which the minimum
cross-section diameter, measured on the deforming mesh, corresponds to the experimental value of logarith-
mic strain at failure. The smaller plots in Fig. 9 show the distributions of TF and eEq along the neck radius.
The data in Fig. 9a confirm the known experimental evidence that unnotched specimens always start failing
at the neck center, where the maximum values of both TF and eEq occur at each instant of the load history.
On the contrary, for notched specimens (Fig. 9b), the triaxiality ratio is maximum on the neck center while the
plastic strain is maximum at the peripheral points.
If one of the two parameters prevails on the other with regard to its void growth promoting effect, then
some predictions about the failure initiation site would have been possible for notched specimens, using data
such as that in Fig. 9b.
Usually this is not possible because, as stated in [5], for sharp notches ‘‘. . .there is a competition between
high triaxiality levels reached at the center (of the neck section), and the high plastic strains occurring at the
notch root, for determining the failure site.’’

a
TF TF History for Cu 99.97% Unnotched
0.8
0.7 Failure step
TF

0.5

r=a
0.55 r/a [mm]
0.2
0.00 0.20 0.40 0.60 0.80 1.00
r=0
1.40
Failure step

1.30
Eq. Strain
0.4
1.20

r/a [mm]
1.10
0.00 0.20 0.40 0.60 0.80 1.00

ε Eq
0.25
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8

b 1.3
TF History for AISI T304 Notch R2
TF

1.2
1.1 Failure step

0.9

TF
0.6
0.9
r/a [mm]
0.3
r=a 0 0.2 0.4 0.6 0.8 1

1.3
Failure step
0.7 1.2

r=0 1.1

1 Eq. Strain
0.9
0.5 r/a [mm]
0.8
0 0.2 0.4 0.6 0.8 1

ε Eq
0.3
0 0.2 0.4 0.6 0.8 1 1.2 1.4

Fig. 9. History of triaxiality ratio vs. equivalent plastic strain, for Cu 99.97% unnotched and AISI T304 R2 notched specimens.
1214 G. Mirone / Engineering Fracture Mechanics 74 (2007) 1203–1221

According to the damage model [32,33], the triaxiality ratio integrated over the equivalent plastic strain, for
each material point, is the key damage parameter triggering local failure when a critical value is reached,
according to Eqs. (5).
Z eEq
DðeEq Þ ¼ TFðeEq ÞdeEq
Z 0eF ð5Þ
DðeF Þ ¼ TFðeEq ÞdeEq ¼ DCr
0

The critical damage DCr is constant for each material within the TF range of interest in this work (TF > 0.33).
The strain eF is the local failure strain of the first point where rupture initiates, for the given component shape,
constraints and loads.
For TF variation within the range (1/3,1/3), such as in case of torsion with superimposed compressive or
tensile uniaxial load, the damage variable is still the same but its limiting domain differs from the constant
value DCr proposed by expression (5).
In principle, the damage variable has to be evaluated in all points within the volume of the specimen, so that
the first one where the limiting value DCr is reached may be identified as the failure initiation site.
For many specimens tested in this work, however, this multiple evaluation was not necessary since, for
unnotched and bluntly notched specimens, the neck center is the point where both TF and eEq are maximum
at every load step up to failure.
For the sharper notches, such as that for AISI-T304-R2, this certainty was not the case (Fig. 9b). However
the damage variable D calculated all over the minimum cross-section resulted again to be greater at the neck
center than elsewhere; then, failure should have initiated at r = 0 also for the sharper notches of Table 1.
This finding is in agreement with the fracture surfaces observed for all the specimens, in fact the cup-cone
shapes visible in Fig. 10 are typical indicators of early failure at the center of the cross-section and successive
propagation toward the peripheral areas.
In Ref. [5], it is also reported that failure of specimens with similar notch severity initiated at the neck cen-
ter, and only one of the notches tested in that work, the smallest-radius one, was found to initially fail on the
external surface.
It might be said that failure initiate at outer points of the neck section only for very sharp notches, whose
initial undeformed geometry induces strain nonuniformity much greater than the TF nonuniformity. In order
to allow failure localisation at the outer surface, the material should also be not too ductile because, as the
strain increases, eEq distributions become more uniform while TF ones become more nonuniform; so, the ini-
tial ‘‘elastic’’ advantage of external points of sharp notches as candidates for failure initiation, decreases as the
plastic deformation goes on.
In Figs. 11 and 12 are reported the histories of the TF and of the damage D, calculated by applying Eq. (5)
to the FE results at the neck center for all the specimens.
The local values of eEq, TF and D corresponding to experimental failure are marked by round black
symbols.
It is worth noting that local values of eEq at the neck center differ from the overall logarithmic strain at each
load step which does not consider nonuniformity of strains across the neck section.
The damage variable D is well suitable for predicting failure because its critical value is almost constant
for all the different geometries of a given material, as visible in Fig. 12 and reported in Table 4, where
Dcr_Avg is the average critical damage found for each metal. The brittlest among the materials considered

Fig. 10. Fracture surfaces of R5 aluminium, R5 copper and R2 AISI T304.


G. Mirone / Engineering Fracture Mechanics 74 (2007) 1203–1221 1215

0.9 0.3
TF AL2011 - neck center D AL2011 - neck center
0.8
0.25 Dcr
0.7 R5
0.6 R 10 0.2 R5
R 10 Smooth
0.5
0.15
0.4 Smooth
0.3 0.1
0.2
0.05
0.1
εEq ε Eq
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7

1.2 0.8
TF Cu 99.9% - neck center D Cu 99.9% - neck center
R5 0.7 Dcr
1
0.6
0.8 R 10 R5
0.5 R 10 Smooth
0.6 Smooth 0.4
0.3
0.4
0.2
0.2
0.1
εEq ε Eq
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 1.1 1.2 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 1.1 1.2

1.4 1.2
TF AISI T 304 - neck center D AISI T 304 - neck center R 30
1.2 R2 1
R 30 Dcr
1 R5 R 15
R5 0.8 R2
Smooth
0.8 R 15
0.6
0.6 Smooth
0.4
0.4

0.2 0.2
ε Eq ε Eq
0 0
0 0.4 0.8 1.2 1.6 0 0.4 0.8 1.2 1.6

Fig. 11. Triaxiality ratio and damage history at the neck center, specimens of aluminium 2011, Cu 99.9% and AISI T304.

is the Aluminium alloy showing the lower value of Dcr_Avg. The two types of almost-pure copper have cumu-
lated triaxiality ratio values very close to each other, while the three steels exhibit Dcr_Avg values ranging
from slightly above those of the copper to a quite higher value, for the AISI T304 which is the tougher steel
tested.
The damage model is then used to predict failure of specimens, by imposing that rupture initiates when the
local condition D = Dcr_Avg is firstly achieved at a node.
The logarithmic strains ef predicted by FE when D = Dcr_Avg and their error with respect to experimental
logarithmic strains at failure, are reported in Table 5.
Obviously, the greater error levels are found for those metals (aluminium and copper 99.9%) which exhib-
ited scattered Dcr values of different notch geometries around the material-averaged value Dcr_Avg. This scatter
in failure predictions is almost equally distributed around the average value of Dcr, so that no intrinsic ten-
dency at overestimating or underestimating the failure strains is found.
The cumulated TF is then estimated also for three specimen configurations tested in [5], for which Alves
and Jones plotted the TF vs. eEq at the material points r = 0 and r = a.
The triaxiality ratio shown in those curves is roughly approximated by linear functions of eEq because no
further data are supplied in [5] for better representation of these relationships. The limiting values of the
1216 G. Mirone / Engineering Fracture Mechanics 74 (2007) 1203–1221

1 0.9
0.9 TF AISI 1040 - neck center D AISI 1040 - neck center
0.8
0.8 Dcr
R 10 0.7
0.7 R 10 R 20
0.6
0.6 R 20 0.5
0.5
0.4 Smooth
0.4 Smooth
0.3 0.3
0.2 0.2
0.1 εEq
0.1 εEq
0 0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 0 0.2 0.4 0.6 0.8 1 1.2 1.4

1 0.8
TF ASTM A284 - neck center ASTM A284 - neck center
0.9 D
0.7
0.8 R 10 Dcr
0.6
0.7 R 20 R 10 R 20 Smooth
0.6 0.5
Smooth
0.5 0.4
0.4
0.3
0.3
0.2
0.2
0.1 0.1
εEq εEq
0 0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 0 0.2 0.4 0.6 0.8 1 1.2

1.6 0.8
TF Cu 99.97% - neck center D Cu 99.97% - neck center
1.4 0.7
Dcr
1.2 0.6
1 R1 0.5
R1 R4 Smooth
0.8 R4 Smooth 0.4
0.6 0.3
0.4 0.2
0.2 0.1
εEq εEq
0 0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 0 0.2 0.4 0.6 0.8 1 1.2 1.4

Fig. 12. Triaxiality ratio and damage history at the neck center, specimens of AISI 1040, ASTM A284 and Cu 99.97%.

Table 4
Values of Dcr
Dcr Aluminum 2011 Copper 99.9% Steel AISI T304 Steel AISI 1040 Steel ASTM A284 Copper 99.97%
Unnotched 0.239 0.674 1.01 0.707 0.652 0.644
Notch R30 – – 0.99 – – –
Notch R20 – – – 0.722 0.649 –
Notch R15 – – 0.996 – – –
Notch R10 0.236 0.617 – 0.704 0.652 –
Notch R5 0.222 0.656 1.029 – – –
Notch R4 – – – – – 0.636
Notch R2 – – 0.993 – – –
Notch R1 – – – – – 0.625
Dcr_Avg 0.232 0.649 1.003 0.711 0.651 0.635
Max discrepancy 4.7% 5.2% 2.4% 1.5% 1.1% 1.6%
G. Mirone / Engineering Fracture Mechanics 74 (2007) 1203–1221 1217

Table 5
Estimated logarithmic strain at failure and error with respect to experiments
Aluminum 2011 Copper 99.9% Steel AISI T304 Steel AISI 1040 Steel ASTM A284 Copper 99.97%
Unnotched Estimated ef 0.50 0.92 1.44 1.08 0.99 1.22
Error 4% 4% <1% <1% <1% <1%
Notch R30 Estimated ef 1.37
Error 1%
Notch R20 Estimated ef 0.91 0.83
Error 1% 1%
Notch R15 Estimated ef 1.29
Error < 1%
Notch R10 Estimated ef 0.34 0.67 0.82 0.73
Error 3% 6% 2% <1%
Notch R5 Estimated ef 0.30 0.59 1.12
Error 7% <1% 2%
Notch R4 Estimated ef 0.95
Error <1%
Notch R2 Estimated ef 1.04
Error 1%
Notch R1 Estimated ef 0.76
Error <1%
jErrorjAvg 4.7% <3.3% <1.2% <1.3% <1.0% <1.0%
Max error 7% 6% 2% 2% 1% <1%

linearly varying TF(eEq) between the zero strain and the local failure strain, taken from Figs. 11 to 14 of Ref.
[5] are then:

• Specimen n0509 (nominal notch radius 0.5 mm): for r = 0: TF(0) = 0.8, TF(0.15) = 1.4; for r = a:
TF(0) = 0.95 and TF(1.2) = 0.55;
• Specimen n205 (nominal notch radius 2 mm): for r = 0: TF(0) = 0.5, TF(0.4) = 1.2; for r = a: TF(0) = 0.6
and TF(0.8) = 0.45;
• Specimen n452 (nominal notch radius 4 mm): for r = 0: TF(0) = 0.7, TF(1.05) = 0.95; for r = a:s
TF(0) = 0.4 and TF(0.95) = 0.4.

It is worth noting that, in [5], the local triaxiality values are plotted vs. the logarithmic strain (neck-averaged
current strain), but the above results refer to local strain.
The TF integrals expressing the damage variable D at failure are then obtained as trapezoid areas. This
calculations suggests that the critical damage for the metal FE430A tested in [5] ranges between 0.87 and
0.9, and these values are firstly reached at the neck outer points for the sharply notched specimen n0509
(Dcr = 0.9) and on the neck center of the bluntly notched specimen n452 (Dcr = 0.87). This is in perfect agree-
ment with the failure initiation sites experimentally found in Ref. [5].
The numerical data for the specimen n205 do not comply with experimental results because the variable D
at failure results to be 0.34 at the neck center and 0.42 at the neck outer radius. These values in both locations
are much less than what is needed for local failure, according to the data from the other two specimens of the
same metal.
However, some considerations arise about the history of TF distributions for this specimen. In fact, given
the three notch geometries, the TF on the neck center of specimen n205, at low deformation levels, should be
approximately halfway between the corresponding values for the two other specimen geometries. Instead, at
low strains it is reported to be the lowest in the entire set of three specimen geometries.
1218 G. Mirone / Engineering Fracture Mechanics 74 (2007) 1203–1221

7. Micrographic assessment of damage distributions

The damage parameter D, found to be greater at the neck center than everywhere else within the volume of
the specimens tested, indicates that the greater void growth and the consequent failure initiation occurred
there.
This is known and obvious for unnotched specimens because TF and eEq are both maximum at r = 0, but,
for notched specimens, it is not possible to say a priori where the greater void growth will develop. In fact, in
the latter case, the stress triaxiality is maximum on the neck center but the plastic strain may be maximum on
the outer points, depending on the notch severity.
A first confirmation that failure initiated at the neck center, also for the sharply notched specimens tested, is
given by the cup-cone shaped fracture surfaces in Fig. 10, but further evidence is sought using optical micro-
graphs taken for some of the ruptured specimens. Polished fragments are observed with a reflection micro-
scope at magnifications between 80 and 120X. Various images aligned along the neck section trace are
overlapped to obtain the entire profile of the fracture surface and its neighbors, as shown in Fig. 13 for a
smooth specimen of 99.9% copper and a specimen of AISI T304 having a 2 mm notch radius.
Observation of micrographs confirms the predictions of the damage model about failure initiation site, in
fact, larger voids, marked by red circles, are located at the central region of the fracture surface regardless of
the notch.
In case of smooth specimens this is obvious for the reasons already described. As for notched specimens,
larger voids located on the neck center indicate that the triaxiality effect (maximum at neck center) in promot-
ing voids enlargement, surmonted the strain effect (maximum at the outer points of the cross-section).
Only if the distribution of equivalent strain is extremely nonuniform as for the n0509 specimen in [5] (at
failure, eEq(r = a)/eEq(r = 0) > 10), the strain-induced growth prevails on the TF-induced one and failure
may initiate at the outer points of the neck.
For smooth specimens, the voids are spread almost uniformly along the neck cross-section and their shape
and size varies slightly from the neck center toward the external radius, because the non-uniformity in the dis-
tributions of eEq and TF is moderate.
On the contrary, for notched specimens, the nonuniformity of these variables is larger, the voids growth far
from the center is mainly promoted by higher axial strain and lower triaxialities (elongated shapes and small
size), while the growth of voids close to the neck center is due to lower eZ and higher TF (more elliptical/cir-
cular shapes and greater size). This is in agreement with the findings of Rice & Tracey [9] who demonstrated
that TF and eEq are mainly related to volume change of voids, while principal strains are directly responsible

Fig. 13. Micrographs from the unnotched specimen of Cu 99.9% and the R2 specimen of AISI T304.
G. Mirone / Engineering Fracture Mechanics 74 (2007) 1203–1221 1219

for the shape modifications of a void. Also in [37,38] there is reported evidence of triaxiality-induced dilatation
and -strain-induced distortion of voids.
The role of the material hardening on void growth is visible by comparing the outer zones of the neck sec-
tions in Fig. 13. For r close to a, the strain and the triaxiality in the notched steel specimen are only slightly
greater than they are in the smooth copper specimen. Fig. 9 gives the following ratios for these two metals at
r = a: TF is 0.27/0.33 and eEq is 1.15/1.25. Despite this small difference, in that area of the mildly-hardening
copper there are more voids, larger and with more circular shapes than in the same zone of the strongly-hard-
ening steel. Reduction of void growth rate due to higher hardening is also reported in [9,38].
Voids fraction measurements are attempted using a software for image analysis, but considerable inaccu-
racy arose because of the many pixels with different gray shades forming the boundary of each void. In order
to select the pixels belonging to a single void, the algorithm requires to specify a threshold gray level for dis-
criminating between the pixels within the voids and those of the metal matrix. The choice of this threshold
level is rather arbitrary and, hence, the void fraction measurements assume very different values. Though,
whatever threshold level was set, the voids distributions have shown their maximum values on the central area
of the failed sections.

8. Conclusions

Two recent methods are used for characterizing the elastoplastic response of six ductile metals and for pre-
dicting their failure under various stress triaxiality histories, each corresponding to a different notch geometry.
True stress–true strain curves for all the metals and the notches are derived from experimental tensile tests,
then the true curves from the unnotched specimens are corrected by the MLRr function accounting for the
necking effect. The resulting data approximate the flow curve for each material. The MLR necking model
is also applied for finding the distributions of stresses and plastic strains all over the neck section of the unnot-
ched specimens. Finite elements analyses simulating the tensile tests are then performed, using the previously
calculated flow curves as the input for defining the material behavior; the accuracy of these analyses is verified
by comparing the experimental true curves with those calculated from the FE results. The triaxiality ratio is
then obtained from the FE data, at every load step, at different points of the neck section, showing how this
stress state parameter varies remarkably during the post-necking plastic flow. The model by Bao and Wierzb-
icki is then applied by integrating the calculated triaxiality ratio over the range of the equivalent plastic strain
up to failure, at all the nodes of the minimum cross-section.
The maximum value this integral reaches at failure is confirmed to be a material constant, within an error
level of about 5%. The integral function is then used in a predictive way given that its critical value, triggering
the local failure, is determined for each metal by experimental tests on smooth specimens and MLR or FE
simulation.
The shapes of the fracture surfaces and the micrographic observation of voids, confirm the predictions of
the damage model which indicate the neck center as the site where failure initiates for all the specimens tested.

Appendix

The axial plastic strain at each instant of the tensile test is supposed to vary as a quadratic polynomial of
the radial abscissa r within the interval (0, a)
ez ðrÞ ¼ A þ B  r2 ðA1Þ
Plastic strains are deviatoric and do not induce volume change in every material point:
oqðrÞ qðrÞ
ez ðrÞ þ er ðrÞ þ e# ðrÞ ¼ A þ Br2 þ þ ¼0 ðA2Þ
or r
and solution of (A2) gives q(r), radial displacement:
C 1
qðrÞ ¼  ð2  A  r þ B  r3 Þ ðA3Þ
r 4
1220 G. Mirone / Engineering Fracture Mechanics 74 (2007) 1203–1221

The integration constant C results to be null by imposing that, q(0) = 0, then substitution of the final form of
q(r) into the radial and hoop strains yields:
1
er ðrÞ ¼ ð2A  3Br2 Þ
4 ðA4Þ
1
e# ðrÞ ¼ ð2A  Br2 Þ
4
The axial plastic strain averaged onto the current minimum cross-section must be equal to the logarithmic
plastic strain 2Ln(a0/a) by definition, and the ratio between radial and hoop strains, both averaged on the
same area, is imposed to be equal to the current value of the material-independent polynomial MLRe:
Z a a 
1 0
ezAvg ¼  ez ðrÞ  2pr  dr ¼ 2Ln
p  a2 0 a
Ra ðA5Þ
erAvg pa1 2 0 er ðrÞ  2p  r  dr
¼ 1 Ra ¼ MLRe ða0 ; a; eN Þ
e#Avg pa2 0 e# ðrÞ  2p  r  dr
Solution of Eqs. (7) for each current value of the cross-section radius a, returns the parameters A and B defin-
ing the current distributions of principal strains.
The normality rule, coupled to the known strains and to the material flow curve of Eq. (4), gives the devi-
atoric stresses in the principal reference system.
2 dez ðrÞ
r0z ðrÞ ¼ rEq ðrÞ 
3 deEq ðrÞ
2 der ðrÞ
r0r ðrÞ ¼ rEq ðrÞ  ðA6Þ
3 deEq ðrÞ
2 de# ðrÞ
r0# ðrÞ ¼ rEq ðrÞ 
3 deEq ðrÞ
Only the hydrostatic stress rH(r) is needed to complete the picture of the current stress–strain state on every
point of the minimum cross-section.
The second hypothesis introduced now approximates the distribution of hydrostatic stress as another qua-
dratic function of the radial abscissa, with no linear term in order to reproduce the first order continuity of the
distribution on the neck center:
rH ðrÞ ¼ H 0 þ H 2  r2 ðA7Þ
The axial and radial equilibrium are stated in Eqs. (A8), whose solution gives the parameters H0 and H2, able
to completely identify the current distribution of hydrostatic stress on the neck section.
Z a Z a
2p  r  rZ ðrÞdr ¼ 2p  r  ðr0z ðrÞ þ rH ðrÞÞdr ¼ LoadðaÞ
0 0 ðA8Þ
rr ðaÞ ¼ r0r ðaÞ þ rH ðaÞ ¼ 0
By adding the hydrostatic stress to the three principal deviatoric stresses, the complete stress state is
determined.

References

[1] Bridgman PW. Studies in Large Flow and Fracture. London: McGraw Hill; 1956.
[2] Clausing DP. Effect of plastic strain state on ductility and toughness. Int J Fract Mech 1970(6):71–85.
[3] Alexandrov SE, Goldstein RV. Distributions of stress and plastic strain in notched tensile bars. Int J Fract 1998(91):1–11.
[4] Earl JC, Brown KD. Distribution of stress and plastic strain in circumferentially notched tension specimens. Engng Fract Mech
1976:599–611.
[5] Alves M, Jones N. Influence of hydrostatic stress on failure of axisymmetric notched specimens. J Mech Phys of Solids
1999:643–67.
G. Mirone / Engineering Fracture Mechanics 74 (2007) 1203–1221 1221

[6] La Rosa G, Mirone G, Risitano A. Post-necking elastoplastic characterization: degree of approximation in the Bridgman method and
properties of the flow-stress/true-stress ratio. Met Mat Trans A 2003;34A(3):615–24.
[7] Mirone G. A new model for the elastoplastic characterization and the stress–strain determination on the necking section of a tensile
specimen. Int J Solid Struct 2004(41):3545–64.
[8] McClintock FA. A criterion for ductile fracture by growth of holes. J Appl Mech 1968(35):363–79.
[9] Rice JR, Tracey DM. On the ductile enlargement of voids in triaxial stress fields. J Mech Phys Solids 1969(17):201–17.
[10] Lemaitre J. A Course on Damage Mechanics. New York: Springer; 1996.
[11] Lemaitre J. A continuous damage mechanics model for ductile fracture. J Engng Mater Technol 1985(107):83–9.
[12] Lemaitre J. How to use damage mechanics. Nucl Engng Desi 1984(80):233–45.
[13] Lemaitre J. Local approach of fracture. Engng Fract Mech 1986(25):523–37.
[14] Chaboche JL. Continuum damage mechanics: Part I and II. J Appl Mech 1988(55):59–72.
[15] Chaboche JL. Continuum damage mechanics and its application to structural lifetime predictions. Rech Aerospatiale 1987(4):37–54.
[16] Bonora N. A nonlinear CDM model for ductile failure. Engng Fract Mech 1997(58):11–28.
[17] Wang TJ. Unified CDM model and local criterion for ductile fracture. Engng Fract Mech 1992(42):177–83.
[18] Chandrakanth S, Pandey PC. An isotropic damage model for ductile material. Engng Fract Mech 1995(50):457–65.
[19] La Rosa G, Mirone G, Risitano A. Effect of stress triaxiality corrected plastic flow on ductile damage evolution in the framework of
continuum damage mechanics. Engng Fract Mech 2001;68(4):417–34.
[20] Gurson AL. Continuum theory of ductile rupture by void nucleation and growth. ASME Trans, J Engng Mat Tech 1977(99):2–15.
[21] Tvergaard V, Needleman A. Analysis of the cup-cone fracture in a round tensile bar. Acta Metallurgica 1984;32(1):157–69.
[22] Needleman A, Tvergaard V. An analysis of ductile rupture in notched bars. J Mech Phys Solids 1984;32(6):461–90.
[23] Tvergaard V, Hutchinson JW. Two mechanisms of ductile fracture: void by void growth versus multiple void interaction. Int J Solid
Struct 2002(39):3581–97.
[24] Ragab AR, Saleh ARCh. Evaluation of constitutive models for voided solids. Int J Plasticity 1999(15):1041–65.
[25] Saie M, Pan J. Void nucleation effects on shear localization in porous plastic solids. Int J Fract 1982(19):163–82.
[26] Le Roy G, Embury JD, Edwards G, Ashby MF. A model of ductile fracture based on the nucleation and growth of voids. Acta
Metallurgica 1981(29):1509–22.
[27] Mirone G. Approximate model of the necking behaviour and application to the void growth prediction. Int J Damage Mech
2004:241–61.
[28] Straffelini G. Ductility of materials with ferritic matrix. Mater Sci Engng 2003(342):251–7.
[29] Chaouadi R, De Meester P, Vandermeulen W. Damage work as ductile fracture criterion. Int J Fract 1994(66):155–64.
[30] Schiffmann R, Heyer J, Dahl W. On the application of the damage work density as a new initiation criterion for ductile fracture.
Engng Fract Mech 2003(70):1543–51.
[31] Schiffmann R, Bleck W, Dahl W. The influence of strain history on ductile failure of steel. Comput Mater Sci 1998(13):142–7.
[32] Bao Y, Wierzbicki T. On fracture locus in the equivalent strain and stress triaxiality space. Int J Mech Sci 2004;46:81–98.
[33] Bao Y, Wierzbicki T. A comparative study on various ductile crack formation criteria. ASME J Engng Mater Technol
2004:314–24.
[34] Pardoen T, Doghri I, Delannay F. Experimental and numerical comparison of void growth models and void coalescence criteria for
the prediction of ductile fracture in copper bars. Acta Metallurgica 1998(46):541–52.
[35] Mackenzie AC, Hancock JW, Brown DK. On the influence of state of stress on ductile failure initiation in high strength steels. Engng
Fract Mech 1977(9):167–88.
[36] Hancock JW, Mackenzie AC. On the mechanism of ductile failure in high strength steels subjected to multi-axial stress states. J Mech
Phys Solid 1976(24):147–69.
[37] Benzerga AA, Besson J, Pineau A. Anisotropic ductile fracture Part I: experiments. Acta Materialia 2004(52):4623–38.
[38] Ragab AR. Application of an extended void growth model with strain hardening and void shape evolution to ductile fracture under
uniaxial axisimmetric tension. Engng Fract Mech 2004(71):1515–34.

Vous aimerez peut-être aussi