Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Finite Elements in Analysis and Design 42 (2006) 1187 – 1197

www.elsevier.com/locate/finel

Finite element simulation of the simple tension test in metals


Carlos García-Garino a , Felipe Gabaldón b,∗ , José M. Goicolea b
a Laboratorio para la Producción Integrada por Computadora (LAPIC), Instituto Tecnológico Universitario, Universidad Nacional de Cuyo & CONICET,
Casilla de Correo 947, 5500 Mendoza, Argentina
b Grupo de Mecánica Computacional. E.T.S. Ingenieros de Caminos, Canales y Puertos, Universidad Politécnica de Madrid, 28040 Madrid, España

Received 4 November 2005; received in revised form 18 May 2006; accepted 28 May 2006
Available online 18 July 2006

Abstract
In this work the finite element simulation of the simple tension test in metals is discussed. Results from Bridgman as well as experimental
measurements obtained with aluminium specimens are compared with numerical simulations. The computational analyses carried out consider
several finite element technologies as the Q1/P 0 mixed formulation, and various enhanced assumed strain formulations from Simo and co-
workers. The constitutive model considered includes large strain plasticity effects via the multiplicative decomposition of deformation gradient
tensor. A key objective of this work is to perform a detailed discussion for strain and stress distribution at necking zone. The simulations
performed show good agreement with the analytical and experimental results. Analytical results due to Bridgman are confirmed by the
computational simulations. This test can be considered as an adequate benchmark in order to calibrate large strain plasticity finite element codes.
䉷 2006 Elsevier B.V. All rights reserved.

Keywords: Simple tension test; Necking; Plasticity; Finite deformation; Enhanced assumed strain elements

1. Introduction numerical results, and proposed a methodology in order to cal-


ibrate constitutive models. Norris et al. [6] tested a steel spec-
The simple tension test is a valuable tool in order to charac- imen, and compared their experimental results with numerical
terise constitutive equations for metals in presence of plasticity simulations and some disagreements with the analytical results
and large deformation. Bridgman [1,2] has proposed the clas- proposed by Bridgman were reported. Goicolea [7] performed
sical analytical results for this problem taking into account the experimental tests using aluminium specimens that are detailed
nonuniform strain and stress distribution for the necking zone. in Section 3 and studied numerically the problem, reporting a
Another important result from this author is the factor that re- good agreement with Bridgman results. More recently Cabezas
lates the average axial stress zz with the yield stress Y values and Celentano [8] have analysed the fit of constitutive equations
in the necking zone. Davidenkov and Spiridonova [3] based from experimental and numerical results, using cylindrical and
on metallurgical studies supported some empirical results from sheet steel specimens.
Bridgman about the strain distribution at the necking section. In the computational mechanics literature the problem has
In the next section of this paper the more relevant analytical been widely studied, in order to calibrate finite element codes
results are summarised. It is important to point out that these and the corresponding constitutive models. Since the initial
results have not been validated through computational mechan- work of Wilkins [5], Chen [9] and Needleman [10], further
ics studies so far. studies include those of Hallquist [11], Simo [12] and Ponthot
Different works that compare numerical and experimental [13,14]. In most of these contributions the applied load and
results for this test can be found in the literature. Wilkins [4,5] necking ratio evolutions were compared with Norris et al. re-
appears to be the first author who compared experimental and sults [6], but no references to Bridgman results were reported.
Valiente [15] has studied blunt notched tension bars in order to
obtain a better insight on Bridgman’s results.
∗ Corresponding author. Tel.: +34 913366696; fax: +34 913366702. An overview of the problem can be found in an internal
E-mail address: felipe@mecanica.upm.es (F. Gabaldón). report of the authors [16] and the references therein. In this
0168-874X/$ - see front matter 䉷 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.finel.2006.05.004
1188 C. García-Garino et al. / Finite Elements in Analysis and Design 42 (2006) 1187 – 1197

paper Bridgman’s and finite element results are compared, and From the first hypothesis results (see Appendix A)
good agreement is found. From a numerical point of view, in
the authors’ opinion the problem can be adequately simulated Ḋ
drr = d = , (4)
using different finite elements codes with plasticity and large D
strain capabilities [17]. In this paper the study of nonuniform
stress distribution at necking zone in the simple tension test can where D is the current diameter of the specimen at necking
be considered an adequate benchmark in order to calibrate large section. From symmetry conditions follow:
strain plasticity finite element codes, using the displacement,
dr  = dz = drz = 0. (5)
strain and stress fields.
Finally, the incompressibility of plastic flow leads to
2. Analytical results
Ḋ D
In the simple tension test two well defined stages are recog- dzz = −2drr = −2 ⇒ p = zz = −2 ln . (6)
D D0
nised. The first stage takes place before the load reaches the
maximum value, and a homogeneous response is found. After Consequently, p and hence the size of the isotropic yield func-
the maximum load, the post-critical path is characterised for a tion Y (p ) may be considered uniform over the neck section.
well marked necking in the central zone of the bar.
In the first stage of the process the response is defined in 2.2. Stress distribution in the necking section
terms of the axial stress zz , that coincides with the average
longitudinal stress zz for the precritical load path, and the axial From the kinematics of the problem, defined in Eqs. (4)–(6),
natural strain zz : it follows that components r  , rz and z vanish at the neck-
 ing section, and rr =  .
P l dl l
zz = zz = , zz = = ln , (1) After some geometrical assumptions due to Bridgman [1,2]
A l0 l l0 (see Appendix A for details), the distribution of radial and cir-
cumferential stress rr and  results
where l, l0 are the current and original lengths of the bar, re- 
spectively. In this stage zz and zz are coincident with the Von   1 r2

rr =  = Y ln 1 + zz − 0.1 −2 2 . (7)
Mises stress and effective plastic strain, respectively: 2 D

def From the Von Mises criterion the yield stress Y results
zz = eq = 3
2s : s, p = zz (2)
Y = zz − rr . (8)
def
being s =  − tr()1 the deviatoric stress. The effective
1
3
plastic strain is defined from the rate of deformation ten- The distribution of axial stress zz can be found replacing Eq.
def (7) in the Von Mises criterion (8)
sor d = sym jẋ/jx and from its additive decomposition into
elastic and plastic components
   1 
r2
  zz = Y 1 + ln 1 + zz − 0.1 −2 2 . (9)
2 D
d = d e + dp , p = 2 p
3d : dp dt. (3)
The expression of the axial stress zz in (9) can be interpreted
as a uniform axial tension Y plus a superposed variable tension
In the second stage of the process after the maximum load is
rr =  that is maximum at r = 0 and vanishes at the outer
reached, necking occurs at central zone where a nonuniform
surface. The mean value of longitudinal stress zz is
distribution of strains and stresses is found.
zz = Y + rr > Y . (10)
2.1. Strain distribution in the necking section
From Eq. (10) it follows that the yield stress Y no longer can
Bridgman [1,2] and Davidenkov and Spiridonova [3] have be obtained directly from the test measurements via the aver-
proposed an expression for the strain field, derived from the age axial stress zz . Based on experimental results, Bridgman
following assumptions: [1,2] proposed a correction factor that relates zz with the yield
(1) The radial strains are uniform along the necked section. stress Y:
This assumption is based on metallographic studies of grain ⎧    √ 
2 zz − 0.1
⎪ zz
size due to Davidenkov and Spiridonova [3]. ⎪ Y = 1 + √zz − 0.1 ln 1 +

⎪ 2
(2) Incompressibility of plastic flow: drr +d +dzz =0, where ⎨
(zz > 0.1), (11)
(·)rr , (·) and (·)zz account for the radial, circumferential ⎪
⎪ zz

⎪ = 1
and axial components, respectively. ⎩ Y
(3) Symmetry conditions about the r = 0 and z = 0 axes. (zz 0.1).
C. García-Garino et al. / Finite Elements in Analysis and Design 42 (2006) 1187 – 1197 1189

1.5 30
CT1
50 13.2 37.5 37.5 13.2 50 CT2
18.5 25 CT4
15 16.2 15.9 15.9 CT5
20 CT6

P (kN)
15
12.9
10
Fig. 1. Geometry of the specimens. Measurements in mm.
5

Table 1 0
0 0.2 0.4 0.6 0.8 1 1.2
Annealing time for the tension test specimens zz
Specimen Annealing time (h)
Fig. 2. Experimental results. Axial load versus logarithmic strain.
CT1 5
CT2 7
CT4 3
CT1
CT5 5 250 CT2
CT6 5 CT4
CT5
200 CT6
zz (MPa)
3. Experimental tests of aluminium specimens 150

We shall consider here the experimental data obtained in [7] 100


for necking of annealed aluminium bars.
The material used in the tests was HE30 (BS1474) extruded 50

aluminium. The mechanical properties of the material provided


0
by the manufacturers were elastic modulus: E = 67 000 MPa, 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
zz
Poisson’s ratio:  = 0.3, 0.1% proof stress: 239–270 MPa and
ultimate tensile stress: 278–293 MPa for ultimate elongation
Fig. 3. Experimental results. Average axial stress versus natural strain.
7%.10%.
The material was machined for obtaining the specimens
whose geometry is shown in Fig. 1. A slight taper was given
in order to control the necking position, with the smallest
diameter being located at the mid section.
Five tension tests were carried out for specimens exposed
to an annealing process at 350 ◦ C. The time of this thermal
treatment is shown in Table 1.
The tests were carried out applying the load with stroke con-
trol, at cross-head velocities 0.5–10 mm/min, up to specimen
fracture. Frequent measurements of the neck diameter were
taken with a calliper of precision ±0.05 mm. Further details
about the test procedure can be found in [7].
The results of the tests are summarised in Figs. 2 and 3.
Fig. 2 shows the applied axial load P versus natural strain
zz at the minimum neck section. A peak value of the load is
reached when the necking instability arises, varying between
22 and 25 kN. Onset of necking occurs within a range between
zz =0.15 and 0.20, and maximum strains values before fracture
are zz = 1.15.1.25. These results seem independent of the
annealing time. In Fig. 3 the average axial stress at necking
section (zz = P /A) is plotted versus the axial strain. Although Fig. 4. Tensile specimens after fracture.
the applied load decreases after the onset of necking due to the
area reduction, at local level the material hardens all along the
test. polished in order to perform microhardness measurements.
After the cup and cone fracture of the specimens (Fig. 4) After polishing, microhardness indentations were performed
the nearest intact zone to the failure section was cut and then at several locations along two normal radii. Details of the
1190 C. García-Garino et al. / Finite Elements in Analysis and Design 42 (2006) 1187 – 1197

Table 2
Microhardness tests results

Point Indentation Microhardness


diameter (m) Hm (MPa)

1 53 640

2 53 640

3 54 630

4 52 660

5 52 660

6 53 640

7 53 650

microhardness testing procedure are described in [7]. The re- Table 3


sults are presented in Table 2, showing no significant variation Results of power-law fits to tension test results
of the hardness across the section. Specimen A (MPa) n
From simple considerations of plasticity theory [18], the
CT1 169.1 0.190
hardness H can be related linearly to the material strength Y,
CT2 172.5 0.177
itself a function of the effective plastic strain CT4 184.7 0.196
CT5 191.5 0.170
H = C Y (p ), (12) CT6 190.6 0.178

where C is a constant whose value for aluminium is between Average 181.7 0.182
2.5 and 3.0 [7]. This expression and the results shown in
Table 2 support the Bridgman assumption of uniformity of Y
and p across the minimum neck section. is one of the assumptions in plasticity of metals. Therefore, in
In order to model the behaviour of aluminium a power law order to obtain a better fit for low strains the exponent was
is considered: corrected adopting the final value of n = 0.159.
 = An , (13)
4. Finite element formulation
 being the uniaxial Cauchy stress,  the uniaxial natural strain
and A, n material constants. Taking into account the equiva- The kinematics of the constitutive model used is based on
lence between the uniaxial stress and the equivalent stress Y, a the multiplicative decomposition of the deformation gradient
convenient form of expressing the hardening law is [19–21]
 
Y n F = F e Fp . (15)
Y = Anzz = A p + ≈ A(p )n , (14)
E
The constitutive law is developed in the framework of irre-
where Y /E is the elastic part of the strains that can be neglected versible thermodynamics of solids via an uncoupled free energy
when compared to the plastic part. function [22]
The average axial stress zz plotted in Fig. 3 were corrected
following Eq. (11) to obtain the values of Y. Introducing each  = e (ee ) + p (), (16)
one of these values and the corresponding zz in (14), the pa-
ee being the elastic part of the Almansi strain tensor, and  a set
rameters A and n were adjusted following a linear regression in
of internal variables. The Cauchy stress tensor can be obtained
the logarithmic values for all the tests. In Table 3, the computed
from elastic free energy as
values are shown, yielding mean values of A = 181.7 MPa. and
n = 0.182. je (ee )
Although the average power law obtained represents a good = . (17)
jee
overall fit, some discrepancies do exist in the region of low
strains (z < 0.2), due to the simplification assumptions of the For the case of metals the elastic strains are negligible and
Bridgman adjustment [1,2]. In this range of deformation elastic in this case the distinction between intermediate and current
strains are not negligible with respect to plastic ones, which configurations have no meaning and elastic strains are small.
C. García-Garino et al. / Finite Elements in Analysis and Design 42 (2006) 1187 – 1197 1191

Then it is possible to write the elastic component of free en- proposed in [24] and the enhanced assumed strain (EAS) for-
ergy function as a quadratic function of elastic component of mulation for finite deformation problems [25,26].
Almansi strain tensor ee and material constants  and : The key ingredient for the Q1/P 0 element formulation is
the smoothing of the pressure field averaging its value at the
e = 21  tr 2 (ee ) + (ee : ee ). (18) element level. This element has been implemented in the fi-
From Eqs. (17) and (18) the Cauchy stress tensor results nite element code SOGDE [22] following the details discussed
in [27].
 =  tr(ee )1 + 2ee . (19) The key ingredient of EAS formulations is to enrich the solu-
tion as compared to a purely displacement-based formulation,
An associated plasticity model based on the Von Mises yield using additional assumed strain modes. For finite deformation
criterion is used problems the starting point is an additive decomposition of the
deformation gradient into a compatible part and an enhanced
Y (p ) = eq , (20)
part [25]:
where eq is the equivalent stress defined previously in (2), 
F = ∇X  + 
F . (27)
and Y is the yield stress. An isotropic hardening is considered   
verifying the potential law (13), with the values A = 181.7 and comp. enh.

n = 0.159. The enhanced part of the deformation gradient is expressed,


The numerical integration of the model is based on the at the element level, as a linear combination of the enhanced
predictor–corrector algorithm detailed in [22,23]. The trial in- modes FI :
termediate configuration is defined in terms of the elastic part
n
of the Finger tensor be−1 = Fe−T Fe−1 . e
enh.
F = eI FI , (28)
The elastic predictor is computed in the deformed configu-
I =1
ration, in terms of the incremental deformation gradient
nenh. being the number of enhanced modes, eI internal de-
t+t jt+t x grees of freedom of element e, and FI the basic enhanced
f= (21)
jt x modes of the deformation gradient. Several expressions of FI
may be employed to define different formulations of enhanced
resulting in
elements [25].
t+t
(be )−1 |(trial) = f−T t (be )−1 f−1 . (22) In this paper the enhanced elements named Q1/E5, Q1/ES5
and Q1/ET 5 from previous work of Armero and Glaser [28]
The plastic corrector is obtained via a backward Euler integra- are considered. These elements have five enhanced modes and
tion scheme, both for Q1/P 0 as well as EAS elements, leading axisymmetric formulation. They have been implemented by the
to authors in the finite element code FEAP [29].
For the EAS elements the enhanced component of deforma-
t+t
(be )−1 = t+t (be )−1 |(trial) + 2t+t n, (23) tion gradient tensor Fenh. is computed in the reference configu-
where the corrector term t+t n is computed using a radial ration. Consequently the trial elastic component Fe is obtained
return mapping algorithm [20,21]. From the value of the elastic from Eq. (15):
Finger tensor computed in (23), the elastic Almansi strain tensor t+t e (trial)
 −1
F| = t+t F t Fp . (29)
is obtained
Then, the trial elastic part of the Finger tensor is computed as
t+t e
e = 21 [1 − t+t (be )−1 ]. (24)
t+t
(be )−1 |(trial) = (t+t Fe |(trial) )−T (t+t Fe |(trial) )−1 . (30)
Substituting (23) in (24), the Almansi strain is expressed via the
additive contribution of a predictor term and a corrector term Plastic corrector step for EAS elements is exactly the same
t+t e e −1 (trial)
one discussed for Q1/P 0 element.
e = 21 [1 − t+t (b ) | ] − t+t n. (25)

Using Eq. (25) in (19) the expression of the Cauchy stress 5. Computational simulation of simple tension test
in the context of the classical theory of plasticity is recovered,
with the additive decomposition of an elastic predictor and a In this section the computational simulation of simple tension
plastic corrector test of a circular cylindrical bar is described. The finite element
codes used in the analysis are SOGDE [22] and FEAP [29].
t+t
 =  tr(t+t ee )1 + [1 − (t+t be )−1 |trial ] − 2t+t n The constitutive model and the finite element approximations
have been discussed in Section 4. The experimental data as well
= t+t |(trial) − 2t+t n. (26)
as the adjustment of the non-linear hardening law are due to
In order to carry out the computational simulation of the ten- Goicolea [7] and have been discussed in Section 3.
sion test, two finite element formulations have been considered: The radius and the height of the specimen are r = 8.1 mm
the well known mixed pressure–displacement Q1/P 0 element and H =75.0 mm, respectively. The mesh covers only a quarter
1192 C. García-Garino et al. / Finite Elements in Analysis and Design 42 (2006) 1187 – 1197

Fig. 5. Finite element mesh, transition and refined zones used.

Fig. 6. Deformed mesh.

of the bar due to symmetry conditions, and it is discretised are fixed (ur = 0) and at the central section (z = 0) longitudi-
with 360 quadrilateral axisymmetric elements and 412 nodes. nal displacements are fixed (uz = 0). The deformed bar after a
Fig. 5 shows the mesh as well as a larger picture of central longitudinal displacement of 10 mm, applied at the end of the
zone in order to show the mesh refinement adopted. In order to bar, is shown in Fig. 6.
ensure location of necking at the central part of the specimen,
the radius at z=0 is 1.85% lower than the end ones, considering
5.1. Evolution of necking
a linear variation along the height of the specimen.
The mesh sensitivity for this problem has been analysed in
The evolution of necking is shown in Fig. 7 where the ratio
detail in [30] and Ref. [16, Chapter 6]. Different structured
D/D0 is plotted in terms of engineering strain l/ l0 . The nu-
meshes, ranging from 50 to 1600 elements, adequately refined
merical results obtained with EAS elements using five integra-
in the necking zone were tested for all the element formulations
tion points are identical and slightly larger than that obtained
considered in this paper. All the meshes provided reasonably
with Q1/P 0 element.
good agreement with experimental results. Meshes with 200
elements seem to be adequate for the objectives of the work.
The sensitivity of the results with respect to the mesh refinement 5.2. Evolution of axial load P
is small.
Boundary conditions are imposed in order to satisfy symme- Fig. 8 shows the evolution of axial load P in terms of
try conditions: at the symmetry axis (r =0) radial displacements logarithmic strain zz for the finite element formulations
C. García-Garino et al. / Finite Elements in Analysis and Design 42 (2006) 1187 – 1197 1193

1 CT1
Q1/E5 250 CT2
Q1/ES5
0.9 CT4
Q1/ET5
CT5
Q1/P0 200 CT6
0.8 Q1/P0

zz (MPa)
150 Q1/E5
Q1/ES5
D/D0

0.7 Q1/ET5
100
0.6
50
0.5

0
0.4 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
0 0.05 0.1 0.15 0.2 0.25 0.3 zz
∆l/l0
Fig. 9. Evolution of zz versus zz . Comparison of experimental and numerical
Fig. 7. Necking versus engineering strain. Q1/E5, Q1/ES5, Q1/ET 5 and results.
Q1/P 0 elements.

2
rr /Y
30 rr,FEM /Y
CT1 rr /Y
CT2 1.5
25 zz,FEM /Y
CT4
CT5 rr /Y, zz /Y
20 CT6 1
Q1/P0
P (kN)

Q1/E5
15 Q1/ES5
Q1/ET5 0.5
10
0
5 0 0.5 1 1.5 2 2.5 3 3.5 4
r
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 Fig. 10. Stress distribution at necking zone. Comparison of analytical and
zz numerical results.

Fig. 8. Evolution of P versus zz . Comparison of experimental and numerical


results.
strain zz (Pmax ), as well as the computational fracture load Pfrac
computed for a logarithmic strain zz = 1.15.
Table 4
Results included in Table 4 lead to the following remarks:
Evolution of Applied load P in terms of logarithmic strain zz , using quadra-
tures with five and nine Gauss points • Element Q1/P 0 behaves slightly stiffer than EAS elements.
Pmax (kN) zz (Pmax ) Pfrac (kN) • All the EAS elements show practically the same response.
Before the peak load is reached the behaviour is identical for
Q1/P 0 23.21 0.172 14.87 all the elements studied.
Q1/E5 (5 G.P.) 23.15 0.176 14.71
Q1/E5 (9 G.P.) 23.15 0.176 14.71
• A mesh sensitivity analysis [16,30] reports a small variation
Q1/ES5 (5 G.P.) 23.15 0.176 14.71 of the results for meshes with a number of elements between
Q1/ES5 (9 G.P.) 23.16 0.179 14.72 200 and 1600 and an adequate mesh design.
Q1/ET 5 (5 G.P.) 23.15 0.176 14.72
Q1/ET 5 (9 G.P.) 23.15 0.176 14.71 5.3. Evolution of mean axial stress zz

In Fig. 9 the evolution of zz in terms of logarithmic strain


considered in this work, and a good agreement with the exper- zz is shown, and numerical results obtained for Q1/P 0 and
imental results is obtained. EAS elements are compared with the experimental results ob-
As explained before in Section 3, experimental peak loads tained with all the specimens. As can be seen experimental and
range between 21 and 25 kN, onset of necking between zz = numerical results adjust adequately.
0.15 and 0.20, and maximum strains before fracture between
zz = 1.15 and 1.25 being the applied load Pfrac = 14.85 ± 5.4. Stress distribution at the necking section
0.84 kN.
In order to provide a better insight on the evolution of applied The numerical simulation of stress distribution at necking
load P, Table 4 lists the values corresponding to the maximum section has been studied in detail by the authors in [16]. Fig. 10
load level Pmax and its companion values of the logarithmic shows a comparison of analytical results due to Bridgman [1]
1194 C. García-Garino et al. / Finite Elements in Analysis and Design 42 (2006) 1187 – 1197

Fig. 11. Effective plastic strain contours at the necking zone.

2
FEM
FEM 240 Bridgman
Bridgman

1.5 220
Y (MPa)

200
p

180
0.5
160

0 0 0.5 1 1.5 2 2.5 3 3.5 4


0 0.5 1 1.5 2 2.5 3 3.5 4 r
r

Fig. 13. Yield stress profile at necking section. Comparison of numerical


Fig. 12. Effective plastic strain profile at necking section. Comparison of
results (FEM) with Bridgman’s ones.
numerical results (FEM) with the ones due to Bridgman.

puted from finite element analyses, and from Bridgman ex-


with the numerical ones obtained using the Q1/ET 5 element pressions. As follows from Eqs. (6) and (14), Bridgman pro-
with a 9 point Gauss quadrature. posed a uniform distribution for both effective plastic strain
In order to compare the finite element results with the Bridg- and yield stress at necking zone, that are quite well adjusted
man ones, the current diameter D computed via finite element by finite element results as it is shown in the aforementioned
method is replaced in Eqs. (6), (7) and (9). figures.

5.5. Yield stress and effective plastic strain distribution at 5.6. Yield stress Y versus mean axial stress zz
necking section
In order to obtain a better insight about the correction factor
The values of yield stress Y and effective plastic strain p at that relates the yield stress Y with the mean axial stress zz
necking zone are required in order to calibrate the constitutive due to Bridgman [1], the theoretical values given in Eq. (11)
equation. Fig. 11 shows the effective plastic strain contours su- are compared with the numerical ones obtained in this work.
perimposed on the deformed mesh for EAS and Q1/P 0 ele- In Fig. 14 it can be seen that the numerical values of zz /Y
ments. are lower than the values obtained from Eq. (11) for zz < 0.8.
Figs. 12 and 13 show the profile of effective plastic strain Computational and Bridgman values match quite well for the
p and yield stress Y, respectively. These values are both com- interval 0.8 < zz < 1.2. The numerical values of zz are greater
C. García-Garino et al. / Finite Elements in Analysis and Design 42 (2006) 1187 – 1197 1195

1.6 ferent FEM technologies. Both the displacement field and its
FEM
1.5 Bridgman derivatives (stress and strain) can be contrasted with experi-
mental and analytical results.
1.4
• The numerical codes used in this paper reproduce well avail-
1.3 able results and can be a valuable tool in order to calibrate
zz / Y

1.2
constitutive equations for metals under large strain plasticity.

1.1 Acknowledgements
1
The financial support granted from the argentinian Agencia
0.9
Nacional de Promoción Científica y Tecnológica, Projects PICT
0.8 12-03268 and PICTR 184 is gratefully acknowledged.
0.2 0.4 0.6 0.8 1 1.2 1.4
z
Appendix A. Strain and stress distribution at the necking
Fig. 14. zz /Y versus zz . Comparison of computational (FEM) and Bridgman zone
results.
A.1. Strain distribution at necking section
than Bridgman’s ones for zz > 1.2. However, the last situation
The deformation rates for a revolution solid can be written
is out of practical interest because the specimen is fractured
in cylindrical coordinates r,  and z [31] as
before zz reaches the value 1.2. In a recent work Cabezas and
Celentano [8] have found similar results about the discussed ju̇ u̇ jẇ
˙rr = , ˙ = , ˙zz = , (A.1)
correction factor comparing numerical and experimental results jr r jz
of a SAE 1045 steel specimen.
ju̇ jẇ
˙r  = ˙z = 0, ˙rz = + , (A.2)
jz jr
5.7. Discussion of results
u̇ and ẇ being the radial and axial components of velocity,
From the results showed in previous sections it can be pointed respectively.
out: From symmetry considerations at z = 0, results
˙rz = 0. (A.3)
• The evolution of applied load P and average axial stress
zz show good agreement with experimental results (Figs. 8 The hypothesis of uniformity of radial strains has been validated
and 9). via metallographic analyses of grain size [3] and microhardness
• Effective plastic strain and yield stress profiles at necking tests [7], resulting
(Figs. 12 and 13) are in good agreement with results due to
Bridgman from Eqs. (6) and (14). ju̇ u̇ ṙ Ḋ
= = = , (A.4)
• Stress distribution at necking zone obtained from finite ele- jr r r D
ment analyses agree with the distribution reported by Bridg- where D is the current diameter at necking section. Substitut-
man (Fig. 10). ing (A.4) in (A.1) the strain distribution at necking section is
• Correcting factor in Eq. (11) proposed by Bridgman must obtained:
be further discussed because it is not in agreement with the
finite element results. Ḋ jẇ
˙rr = ˙ = , ˙zz = . (A.5)
D jz
6. Conclusions
A.2. Stress distribution at necking section
A computational simulation of tension test with aluminium
cylindrical bars has been carried out using several element tech- Taking into account the strain distribution at necking section,
nologies, and the numerical results obtained have been com- the Cauchy stress tensor can be written as
 
pared with experimental tests. From these analyses the follow- rr 0 0
ing conclusions can be remarked:  = 0 rr 0 . (A.6)
0 0 zz
• Finite element technologies and large strain elastoplastic con-
The equilibrium equations in terms of cylindrical coordinates
stitutive models available from the state of the art are able
[31] are considered in order to compute the stress distribution
to simulate properly the simple tension test. In particular,
at necking section
Bridgman’s results are confirmed using FEM simulations.
• Simple tension test appears to be an adequate benchmark in jrr jrz rr − 
+ + = 0, (A.7)
order to calibrate large strain plasticity codes as well as dif- jr jz r
1196 C. García-Garino et al. / Finite Elements in Analysis and Design 42 (2006) 1187 – 1197

isostatic line The curvature radius is not easy to measure and Bridgman [1]
proposed to express it as a function of the radial coordinate r:
 
3 1 D2
= + DR − r 2 , (A.18)
2r 4
z

R being the curvature radius of the outer isostatic line (which


1 is the contour of the specimen). Other expressions of are


reported in [15].
Substituting (A.18) in (A.17), and integrating results
r   
r2 D
rr = Y ln 1 − + . (A.19)
Fig. A1. Principal stresses diagram. DR 4R
In order to avoid the measurement of R, Bridgman [1] suggested
jrz jzz rz the simplification
+ + = 0. (A.8)
jr jz r
D 
Substituting the components of (A.6) in Eqs. (A.7) and (A.8) = 2 zz − 0.1 (zz > 0.1). (A.20)
R
jrr jrz Then, replacing (A.20) in (A.19) the distribution of rr and
+ = 0, (A.9)
jr jz zz at necking section can be obtained using the neck diameter
jzz value D:
= 0. (A.10)   
jz  1 2r 2
rr = Y ln 1 + zz − 0.1 − 2 , (A.21)
Being
the angle of the isostatic lines with the axial direction 2 D
(see Fig. A1), the tangential stress rz is written in terms of
  
 1 2r 2
principal stresses: zz = Y 1 + ln 1 + zz − 0.1 − 2 . (A.22)
2 D
rz = (3 − 1 ) sin
cos
. (A.11)
References
Nearing to necking section (z ≈ 0) the angle
is very small
and 1 = rr , 3 = zz . Therefore, rz can be rewritten as [1] P. Bridgman, The stress distribution at the neck of a tensile specimen,
Trans. Am. Soc. Met. 32 (1944) 553–574.
rz = (zz − rr )
. (A.12) [2] P. Bridgman, Studies in Large Plastic and Fracture, McGraw-Hill, New
York, 1952.
Substituting (8) in (A.12) [3] N. Davidenkov, N. Spiridonova, Analysis of the state of stress in the
neck of a tension test specimen, in: Proceedings of American Society
rz = Y
. (A.13) of Testing Materials, vol. 46, 1946, pp. 1147–1158.
[4] M. Wilkins, Mechanics of penetration and perforation, Int. J. Eng. Sci.
Then, the partial derivative of rz with respect z at z = 0 is 16 (1978) 793–807.
  [5] M. Wilkins, Third progress report of light armor program, Technical
jrz  j
 Y
= Y = , (A.14) Report, Lawrence Livermore National Laboratory, University of
jz z=0 jz z=0 California, rept. UCRL-50460, 1968.
[6] D. Norris, B. Moran, J. Scudder, D. Quiñones, A computer simulation
being the curvature radius of the isostatic line at z = 0. The of the tension test, J. Mech. Phys. Solids 26 (1978) 1–19.
equilibrium Eq. (A.9) results [7] J. Goicolea, Numerical modelling in large strain plasticity with
application to tube collapse analysis, Ph.D. Thesis, University of London,
jrr Y 1985.
+ = 0. (A.15)
jr [8] E. Cabezas, D. Celentano, Experimental and numerical analysis of the
tensile test using sheet specimens, Finite Elem. Anal. Des. 40 (2004)
This differential equation can be integrated in order to obtain 555–575.
the radial stress at necking zone, considering that rr = 0 at [9] W. Chen, Necking of a bar, Int. J. Solids Struc. 7 (1971) 685.
[10] A. Needleman, A numerical study of necking in circular cylindrical bars,
r = D/2
J. Mech. Phys. Solids 20 (1972) 111–127.
 D/2 [11] J. Hallquist, User’s Manual for DYNA2D—an explicit two-dimensional
dr
rr = Y . (A.16) hydrodynamic finite element code with interactive rezoning, Lawrence
r Livermore National Laboratory, University of California, Technical
Report UCID-18756, Revision 1, 1982.
From Eqs. (8) and (A.16), a similar expression is obtained for [12] J. Simó, S. Rifai, A class of mixed assumed methods and the method of
the distribution of axial stress at necking zone: incompatible modes, Int. J. Numer. Methods Eng. 29 (1990) 1595–1638.
  D/2  [13] J. Ponthot, Traitment unifie de la mecanique des milieux continus solides
dr en grandes transformations par le methode des elements finis, Ph.D.
zz = Y 1 + . (A.17)
r Thesis, Universidad de Lieja, Bélgica, 1994.
C. García-Garino et al. / Finite Elements in Analysis and Design 42 (2006) 1187 – 1197 1197

[14] J. Ponthot, Unified stress update algorithms for the numerical simulation [22] C. García-Garino, Un modélo numérico para el análisis de sólidos
of large deformation elasto-plastic and elasto-viscoplastic processes, J. elastoplásticos sometidos a grandes deformaciones, Ph.D. Thesis, E.T.S.
Plasticity 18 (1) (2002) 91–126. Ingenieros de Caminos, Universidad Politécnica de Catalunya, Barcelona,
[15] A. Valiente, On bridgman’s stress solution for a tensile neck applied to 1993.
axisymmetrical blunt notched tension bars, J. Appl. Mech. 68 (2001) [23] F. Gabaldón, Métodos de elementos finitos mixtos con deformaciones
412–419. supuestas en elastoplasticidad, Ph.D. Thesis, E.T.S. Ingenieros de
[16] C. García-Garino, F. Gabaldón, J. Goicolea, A. Mirasso, S. Raichman, Caminos, Universidad Politécnica de Madrid, Madrid, 1999.
Simulación computacional del ensayo de tracción simple con estricción [24] J. Nagtegaal, D. Parks, J. Rice, On numerically accurate finite element
(in Spanish). http://w3.mecanica.upm.es/papers/infor solutions in the fully plastic range, Comput. Methods Appl. Mech. Eng.
me-hyper.pdf, Proyecto PICT-12-03268, FONCYT, ANPCyT, 2004. 4 (1974) 153–177.
[17] J. Goicolea, F. Gabaldón, C. García-Garino, Interpretación de la [25] J. Simó, F. Armero, Geometrically nonlinear enhanced strain mixed
estricción en el ensayo de tracción empleando modelos hipoelásticos e methods and the method of incompatible modes, Int. J. Numer. Methods
hiperelásticos, in: M. Doblaré, J. Correas, E. Alarcón, L. Gavete, M. Eng. 110 (1993) 359–386.
Pastor (Eds.), Memorias del III Congreso de Métodos Numéricos en [26] J. Simó, F. Armero, R. Taylor, Improved versions of assumed enhanced
Ingeniería, SEMNI, Zaragoza, 1996. strain tri-linear elements for 3d finite deformation problems, Comput.
[18] R. Hill, The Mathematical Theory of Plasticity, Clarendon Press, Oxford, Methods Appl. Mech. Eng. 110 (1993) 359–386.
1950. [27] J. Simó, T. Hughes, Computational Inelasticity, Springer, Berlin, 1998.
[19] E. Lee, Elastic–plastic deformation at finite strains, J. Appl. Mech. 36 [28] F. Armero, S. Glaser, On the formulation of enhanced strain finite
(1969) 1–6. elements in finite deformations, Eng. Comput. 14 (1997) 759–791.
[20] J. Simó, A framework for finite strains elastoplasticity based on maximum [29] R. Taylor, http://www.ce.berkeley.edu/∼rlt, 2005.
plastic dissipation and the multiplicative decomposition, part i: continuum [30] C. García-Garino, F. Gabaldón, J. Goicolea, A. Mirasso, Simulación
formulation, Comput. Methods Appl. Mech. Eng. 66 (1988) 199–219. numérica del problema de la estricción de una probeta cilíndrica circular.
[21] J. Simó, A framework for finite strains elastoplasticity based on sensibilidad de los resultados frente a la malla de elementos finitos,
maximum plastic dissipation and the multiplicative decomposition, part in: M. Rosales, V. Cortínez, D. Bambill (Eds.), Mecánica Computacional,
ii: computational aspects, Comput. Methods Appl. Mech. Eng. 68 (1988) vol. XXII, AMCA, Santa Fé, Argentina, 2003, iSSN 1666-6070.
1–31. [31] S. Timoshenko, J. Goodier, Theory of Elasticity, third ed., McGraw-Hill,
New York, 1970.

You might also like