You are on page 1of 15

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/249505888

Procedure for the Determination of True Stress-Strain Curves From Tensile


Tests With Rectangular Cross-Section Specimens

Article  in  Journal of Engineering Materials and Technology · January 2004


DOI: 10.1115/1.1633573

CITATIONS READS

37 1,593

3 authors, including:

Ingo Scheider Wolfgang Brocks


Helmholtz-Zentrum Geesthacht Christian-Albrechts-Universität zu Kiel
69 PUBLICATIONS   1,224 CITATIONS    172 PUBLICATIONS   2,875 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Microstructure of Dental Enamel View project

All content following this page was uploaded by Ingo Scheider on 11 December 2014.

The user has requested enhancement of the downloaded file.


Procedure for the determination of true stress–strain curves
from tensile tests with rectangular cross section specimens

I. Scheider, W. Brocks, A. Cornec


GKSS-Forschungszentrum Geesthacht, Postfach 1160, D-21494 Geesthacht
Fax: 04152 - 872534, E-mail: ingo.scheider@gkss.de

Keywords: tensile test, true stress–strain curve, miniature specimen, digital image processing, grey value
correlation

Abstract

The problem of determining true stress–strain curves from flat tensile specimens beyond the
onset of necking has been investigated based on finite element analyses under consideration of
experimental accessible data using digital image correlation (DIC). The displacement field on the
specimen surface is determined by in-situ deformation field measurement. A 3D finite element
study with different stress-strain-curves has been carried out to develop a formula, with which it
is possible to calculate the true stress subject to the strain in the necking region. The method
has been used to evaluate the true stress–strain curve with a so-called micro flat tensile specimen,
which is normally used to determine the material properties in the material gradient around thin
weldments.

1 Introduction
Materials stress–strain curves necessary as input for finite element analyses cannot be determined
straight forward from global load versus elongation curves beyond the onset of necking after maximum
load, as the strain field becomes inhomogeneous and the stress state triaxial. An analytical solution
derived by Bridgman [3] is used for round tensile bars to determine the effective stress in the necked
section. It requires the continuous measurement of the reduction of diameter and the necking radius
during the test, which requires advanced testing techniques. Mostly, optical methods are applied.
The specimen shape is photographed and the images are either evaluated manually or by electronic
processing.
The Bridgman correction does not work for rectangular specimens as the respective solution assumes
an axisymmetric stress and strain field with constant plastic strain in the smallest necking section.
The strain gradients in the cross section increase with the aspect ratio of rectangular bars. Difficulties
also arise with the experimental determination of the cross section area. It cannot be evaluated
from photos as the rectangular cross section takes the shape of a cushion, see fig. 1. Some procedures
determining stress–strain curves from specimens with rectangular cross sections have been proposed in
literature, see e.g. [6, 10]. However, the first method proposed by Ling is based on an extrapolation of
the stress-strain curve before necking und thus not suitable for calculations with very high strains. The
method proposed by Zhang et al. uses the local thickness reduction in the necking area to determine
the actual cross section area, but the triaxiality of the stress state after occurence of local necking is
not taken into account. A comparison between the present method and that of Zhang et al. will be
given in section 3.
Flat tensile specimens are favourably used for characterizing sheet metals or graded materials like
weldments [11]. In these applications, standard round tensile bars cannot be used, since the smallest

1
I. Scheider, W. Brocks, A. Cornec: Procedure for the determination of true stress–strain curves 2

t = t0 - Δt

b = b0 - Δb

Fig. 1: Cross section of the minimum area plane after occurance of local necking. Dark grey: the
undeformed cross section area with finite element mesh. t, b are measurable quantities.

standardized specimen has a diameter of 2.5 mm, see [2], and the head of the specimen is about 4
mm. Flat specimens can be tested in such cases instead.
For weldments, micro-tensile specimens are applied since the material properties change over short
distances and as many specimens as possible have to be cut from the weldment to sufficiently capture
their gradients. The shape of the micro-tensile specimens, which is used at GKSS [7] for many years,
is displayed in fig. 2. Its width is 2 mm and its thickness just 0.5 mm. Different from conventional
testing, where the global elongation of the specimen is measured via the displacement of the specimen
grips, a digital camera is applied to determine the displacement fields from the specimen surface. The
increments of deformation during testing are evaluated by image analyses of photos from subsequent
load steps. Total deformation can be computed by integration with respect to time, and strain field
from deformation gradients.

thickness
5
2

0,5
9

21

Fig. 2: Geometry and size of the investigated rectangular shaped specimens.

The determination of the true stress would additionally require a finite element simulation of the test.
However, an approximate formula can be derived from numerical parameter studies which relates true
stresses and applied load.
After briefly describing the fundamentals of the present approach, the results of the numerical studies
and the deduction of an analytical relation are presented in section 3. Finally, the respective procedure
is applied to a micro-flat tensile specimen out of a ferritic steel plate.

2 Fundamentals
In a standard uniaxial tensile test, the applied force, F , and the elongation, Δl, over a measuring
length, L0 , are measured and these data are converted to ”nominal” or ”engineering” stress and strain
measures by

o F o ΔL L − L0
σ= ; ε= = , (1)
A0 L0 L0
I. Scheider, W. Brocks, A. Cornec: Procedure for the determination of true stress–strain curves 3

where A0 is the initial cross section of the specimen. These quantities are restricted to small uniform
deformations. For large uniform deformations, ”true” stress and strain measures are defined by

F o o
 L  o

σ= =σ 1+ ε ; ε = ln = ln 1+ ε , (2)
A L0

where A and L are the actual cross section and length of the specimen. For determining A from the
measured length L, incompressibility of plastic deformation is assumed,

AL = A0 L0 . (3)

Eq. 2 is correct as long as the elongation of the tensile bar is uniform. Beyond maximum load, however,
necking starts, deformation becomes non-uniform and the stress state tri-axial. Both effects have to
be accounted for to obtain material curves of true effective stress versus true strain which are required
for a large strain finite element analysis.
If the deformation in a round bar is assumed to be uniform across the necked section (but of course
not along the specimen axis, any more), the uniaxial effective strain can be determined from

D0
ε̄ = 2 ln , (4)
D

where D is the specimen diameter. This requires the measurement of the diameter reduction in the
necking section during the test.
The effective stress under the assumption of a von Mises yield condition is given by
   
3   1 2 2 2
σ̄ = σ σ = (σI − σII ) (σII − σIII ) (σIII − σI ) , (5)
2 ij ij 2


where σij are the deviatoric stresses and σI , σII , σIII the principal stresses, respectively. In a round
bar, we have σI = σz , σII = σIII = σr = σϕ , and hence

σ̄(r) = σz (r) − σr (r) . (6)

Bridgman derived an analytical solution for the stress state in a round necked bar assuming constant
plastic strain over the cross section,
  2

D /4 + ρD − r2
σz = σ̄(ε̄) 1 + ln , (7)
ρD

where ρ is the neck radius, which has to be measured during the test. By integration of eq. (7) over
the area,

D/2
F = 2π σz (r) dr , (8)
0

corr
a correction factor, fround , can be derived which relates the (true) effective stress to the measured
load

F corr 4F corr
σ̄(ε̄) = f = f (9)
A round πD2 round
I. Scheider, W. Brocks, A. Cornec: Procedure for the determination of true stress–strain curves 4

with

1
  
corr
fround = (10)
1+2ρ R
R ln 1 + 2ρ

The problem is now, how to apply this to tensile bars with rectangular cross section. Deformation
is uniform before maximum load so that eq. (2) can still be applied, and as the incompressibility
condition is independent of the geometry, the reduction of area results from eq. (3). However, beyond
uniform elongation, i.e. maximum load, eqs. (4) and (7) to (10) cannot be applied as they are restricted
to axisymmetric conditions.
In the presently proposed procedure the area A is used for the calculation of the true stress according
to eq. (2) only up to the maximum load. In an isotropic solid the measurement of the reduction of
one transverse direction is sufficient, since b/b0 = t/t0 holds under uniaxial tension und thus

t0
A = b2 . (11)
b0

Up to maximum load, the strain field is uniform so that in principle the true strain can be calculated
analogously to eq. (4) as

b 0 t0 b0
ε̄ = ln = 2 ln , (12)
b t b

where b, t are width and thickness of the cross section. In the presently proposed procedure, however,
local strain fields will be measured directly on the specimen surface in the mid section, where no shear
deformation is supposed to occur, as
 
∂ux ∂uy
εI = ln 1 + , εII = ln 1 + (13)
∂X ∂Y

with X and Y being the reference coordinates in length and width direction of the specimen, respec-
tively. The effective total strain is ε̄ = ε̄el + ε̄pl with

t t
pl pl 2 pl pl
ε̄ = ε̄˙ dτ = ε̇ ε̇ dτ , (14)
0 0 3 ij ij

which reduces to

t  
4 pl2 pl2
ε̄ pl
= ε̇I + ε̇pl pl
I ε̇II + ε̇II dτ (15)
0 3

because of the condition of incompressibility, ε̇pl pl pl


I + ε̇II + ε̇III = 0. Since elastic strains are small
compared to plastic strains, at least beyond uniform elongation, their contribution may be neglected.
Eq. (15) yields the ”accumulated” effective plastic strain, ε̄pl , as used in the von Mises theory.
Note, that due to the non-linearity of the deformation, in general the integration (or summation of
deformation increments) will yield a different result than calculating the effective strain from the total
deformation

4 2
ε̄ ≥ (ε + εI εII + ε2II ) . (16)
3 I
I. Scheider, W. Brocks, A. Cornec: Procedure for the determination of true stress–strain curves 5

Only in special cases, i.e. when the principal strain directions remain unchanged and the local stresses
increase monotonically, the equality sign holds in eq. (16). This is realized in the centre of the symme-
try plane of the specimen, with only negligible deviation resulting in practice from the fact that strain
values are measured on the specimens’ surface. In the proposed procedure, it is actually recommended
to use eq. (16) with the equality sign for calculating ε̄ as numerical studies have additionally shown
that the error due to summation of finite increments in the application of eq. (15) is larger than the
error resulting from taking total strains, if the number of load steps recorded in the test is too small,
i.e. load steps are too large. This procedure carries on also beyond maximum load.
After the onset of necking, the actual area of the necking section cannot be measured or calculated any
more due to the cushion like shape, see Fig. 1. Eq. (11) is still used, however, to define a ”nominal”
area

t0
à = b2 , (17)
b0

where b is a measurable quantity, and the idea of relating the true effective stress to the measured
load is still kept,

F corr
σ̄ = frect . (18)

Different from the Bridgman correction, eq. (10), the correction factor for rectangular cross sections,
corr
frect , does not only include the effect of triaxiality of the stress state but also the conversion of the
nominal area, Ã, to the actual area, A.

3 Numerical simulation
corr
A finite element parameter study has been conducted for determining the correction factor, frect ,
from eq. (18).

3.1 FE model

A series of simulations of the micro flat tensile specimen (dimensions see fig. 2) with rectangular cross
section has been performed. The threedimensional model was meshed using 2688 isoparametric 20-
noded elements with reduced integration. Due to symmetry conditions only one eighth of the specimen
was modelled. The load was applied by a prescribed displacement at the top of the model. Due to fine
meshing at the symmetry plane the necking of the model occured in that plane also without geometric
imperfection. The model had 38373 degrees of freedom all together. The mesh is shown in fig. 3.
The strain hardening behaviour of the material has been varied assuming a power law with hardening
exponents of n = 5, 7, 10 and 20. Additionally, the real stress–strain curve of a ferritic steel showing
a Luders plateau after yielding and a typical aluminum alloy Al2024 have been applied.

3.2 Results

The three quantities F, Ã and σ̄, appearing in eq. (18) can be determined from the finite element cal-
culation. F and à are finite element results, σ̄ is predefined by the elastic–plastic material behaviour.
The strain ε̄ does not appear explicitely, but it is nevertheless required, since à and F belong to a
I. Scheider, W. Brocks, A. Cornec: Procedure for the determination of true stress–strain curves 6

Fig. 3: Finite element mesh of the rectangular tensile specimen.

specific load step ”t”, which has to be equated with the stress σ̄ of the material curve via ε̄. Eq. (18)
can thus be rewritten with eq. (17) to


σ̄( ε̄|t ) b2 t t0
corr b0
frect = . (19)
F |t

corr
Fig. 4 shows the correction factor frect in dependence on the effective strain for varying material
hardening.

Fig. 4: Correction function frect


corr
for different material laws.

corr
All curves, frect , start at a value of one which is kept until maximum load occurring at the respective
limit values of uniform strain, ε̄u , that is the equivalent strain at maximum load, and decrease for
ε̄ > ε̄u in the beginning. The minimum value of the correction factor depends on the value of ε̄u . At
I. Scheider, W. Brocks, A. Cornec: Procedure for the determination of true stress–strain curves 7

ε̄ ≈ 0.8, frect
corr
takes a value of one again. For higher values the curves spread significantly. One reason
for this is that low hardening materials (n = 20) localize much stronger than materials with a high
hardening exponent. The localization becomes apparent when showing the width and the thickness
reduction, Δb and Δt, respectively for a specific strain. The relative deformations Δb/b0 and Δt/t0
for ε̄ = 1 are shown in table 1.
The finite element mesh used throughout this study is suitable for a thickness reduction up to Δt/t0 ≈
55%. For higher thickness reductions shear stresses at the symmetry plane rise significantly, which is
not in accordance with boundary conditions at that symmetry plane. The respective reduction value
of Δt/t0 is taken as validity limit for the calculation. The corresponding strain values according to
eq. (16) are also given in table 1.

Table 1: Thickness and width reductions for different materials.


Material Δb/b0 Δt/t0 ε̄(Δt/t0 = 0.55)
at ε̄ = 1 at ε̄ = 1 (validity limit)

n=5 29% 32% 1.4


n=7 27% 53% 1.05
n = 10 25% 56% 0.99
n = 20 23% 63% 0.90
St37 33% 45% 1.22
Al2024 31% 48% 1.12

The influence of the correction factor in eq. (18) is significant for strain values greater than 0.8, i.e.
corr corr
when frect deviates from 1 considerably. The error produced by neglecting frect in the stress
calculation is less than 4% up to ε̄ = 0.8 even for low hardening materials, n = 20. Beyond this limit,
the correction factor cannot be neglected anymore without significant loss of accuracy.

3.3 Approach
corr
A second order polynomial is used to fit the curves frect ( ε̄ ) in fig. 4,


corr 1 ε̄ < ε̄1
frect (ε̄) = , (20)
aε̄2 + bε̄ + c ε̄ > ε̄1

with a, b, c, ε̄1 being fit-parameters. One parameter of the set (a, b, c) can be expressed by ε̄1 through
the necessary condition

corr
frect (ε̄ = ε̄1 ) = 1 (21)

A fit with all three parameters shows that the curvature, expressed by the parameter a in eq. (20) is
approximately equal for all materials. With a = 0.22 and b, c substituted by ε̄1 and a new parameter,
ε̄2 , eq. (20) can be rewritten as


corr 1 ε̄ < ε̄1
frect (ε̄) = (22)
0.22(ε̄ − ε̄1 )(ε̄ − ε̄2 ) + 1 ε̄ > ε̄1

corr
Even though the correction function frect deviates from 1 directly after uniform strain, the error is
minor, when ε̄1 takes a value larger than ε̄u . A least square fit for ε̄1 and ε̄2 gives a linear dependence
of these parameters on the uniform strain ε̄u as shown in fig. 5.
I. Scheider, W. Brocks, A. Cornec: Procedure for the determination of true stress–strain curves 8

Fig. 5: Fitting of the two remaining parameters ε̄1 and ε̄2 to the results of the simulation for
each material law.

The simplest reasonable approach is given by taking a constant value for ε̄2 and a linear function for
ε̄1 , beginning at the origin. Therefore, these parameters can be approximated by ε̄1 = 1.42ε̄u and
ε̄2 = 0.78.
With the above conditions the function for the correction factor is given by

corr 1 ε̄ < ε̄1
frect (ε̄) = (23)
0.22((ε̄ − 1.42ε̄u)(ε̄ − 0.78)) + 1 ε̄ > ε̄1

With eq. (23) in combination with eq. (18) one can easily approximate a stress–strain curve for FE
simulation from the experimental data via


⎪ b0

⎨ F t02 ε̄ < ε̄1
σ̄(ε̄) = b (24)


b 0

⎩ F t02 [0.22((ε̄ − 1.42ε̄u )(ε̄ − 0.78)) + 1] ε̄ > ε̄1
b

The stress σ̄ determined by experiments depends on the equivalent strain in the middle of the minimum
cross section surface (under the assumption of plane stress conditions), σ true , on the minimum width,
b, and on the force, F . Therefore the measurement of the geometrical quantities can be reduced to
the local area of necking.

3.4 Accuracy and advantages of the procedure

Important aspects for the applicability of a procedure are its accuracy and validity limits. For an
evaluation of these aspects, reverse evaluations of the true stress–strain curves have been performed
for all simulations conducted. The comparison between the material curve primarily introduced to
the calculation and the approximated stress using eq. (24) is shown in fig. 6. The agreement between
these two curves is excellent as expected. The strains in fig. 6 reach values up to ε̄ = 1.4 for high
hardening materials, so a very broad range of strains is covered by the proposed method.
I. Scheider, W. Brocks, A. Cornec: Procedure for the determination of true stress–strain curves 9

Additionally, the stress-strain curves for n = 5, n = 20 and Al2024, based on a stress calculation
corr
without correction factor, i.e. frect = 1, have been plotted into fig. 6. As was already stated in
section 3.2, the stress–strain curves are well captured up to ε̄ = 0.8, only for higher strains the curves
without the correction factor deviate significantly from the real material curve.

Fig. 6: Comparison between the material input and the approximation with and without the
correction function. The solid lines refer to the input, the squares are the results of
corr
eq. (24), the dash-dotted lines are the calculations with frect = 1. The validity limits
are given in table 1.

Three advantages over the method proposed by Zhang et al. can be found, two of which are visible
in fig. 7.

1. The validity limit of Zhang’s et al. procedure is a relative thickness reduction of Δt/t0 = 0.5,
which results in a true strain of ε̄ ≈ 0.6 . . . 0.7, depending on the material and aspect ratio1 .
Therefore, the procedure given in this paper by eq. (24) can be used up to much higher strain
values.

2. The accuracy of the method proposed by Zhang et al. is not as high as the method proposed
here. The curves calculated according to their procedure deviate significantly from the material
curves put into the calculations. It is supposed that the inaccurate stress–strain curves result
from the coarse mesh with only about 3000 degrees of freedom (that is less than 10% of what is
used here), which probably led to the poor finite element results. A preliminary study performed
by the authors with a mesh consisting of more than 20000 degrees of freedom (instead of 38000)
showed that even this mesh was too coarse to give a high accuracy in the necking region at high
strain values.
3. The most important argument for the new procedure is that the thickness reduction Δt necessary
for Zhang’s method is a quantity which can hardly be measured practically, see fig. 1. A clip
cannot be fixed at the tensile specimen, since the region of local necking is not known in advance
and the application of optical methods is complicated, since the out-of-plane displacement has
to be measured using a two camera system.

To investigate the sensitivity of the input parameters of the present method to the specimen’s geom-
etry, a calculation has been performed with a slightly changed aspect ratio of 1 : 3.9. It turns out
1 The equivalent strain is calculated by ε̄ = ln(A /A). The actual area is determined from the thickness reduction
0
using fitting functions, which depend on the material and the aspect ratio
I. Scheider, W. Brocks, A. Cornec: Procedure for the determination of true stress–strain curves 10

1400

1200 n=5

1000
Approximation
800 acc. to Zhang et al.
4
σ Al202
600
n=20
400

200 Validity limits for Zhang et al.

0
0 0.5 1.0 1.5
ε
Fig. 7: Approximations of the stress–strain curve proposed by [10]. Solid lines denote the
material input for the calculations, the dashed lines denote the approximations produced
by Zhang’s method.

corr
that up to the validity limit given in table 1 the change in the correction function frect is less than
1%. It can therefore be stated that inaccuracies during machining of the specimen do not affect the
method. However, this does not mean that a complete different aspect ratio can be used with the
same correction function.

4 Experimental application

4.1 Setup and conduction

The material used is a plate of a structural steel (german designation St37), from which a micro-flat
tensile specimen is machined. In addition, round bars are made for comparing the stress–strain curves
developed from the rectangular and the round tensile bars. The testing procedure is explained only
for the rectangular specimen in the following.
The loading device of the testing machine consists of four pins (diameter 3 mm), which act two by
two at the shoulders of the specimen. The specimen and the holder is shown in fig. 8. With this
equipment no initial force or bending moment acts on the specimen, which is difficult to achieve by
clamps. An additional advantage of the device is that it gives free sight from one side to the specimen
surface for the optical gear.
The preparation of the specimens is rather simple. For a high contrast on the specimen surface a
graphite spray is used. An indirect light source prevents any shadow disturbing the initial grey value
distribution.
The deformation of the specimen during the test is recorded optically with a resolution of 1280x1024
pixels. A window of 7.5 · 5mm2 is scanned which guarantees that local necking of the specimen is
0
included. One picture has been taken every strain increment Δ ε= 1.5 · 10−3 , measured over a length
of L0 = 12 mm. With a loading rate of ε̇ = 3.0 · 10−4 , which is slow enough to ensure quasi-static
conditions, this gives one photo every five seconds. The photos of the specimen surface are analysed
by a programme correlating and identifying points of identical grey values in subsequent load steps.
I. Scheider, W. Brocks, A. Cornec: Procedure for the determination of true stress–strain curves 11

Specimen

Loading Pins Detail viewed by digital camera

Holder

Fig. 8: Schematic drawing of the specimen including the loading pins and holder.

4.2 Evaluation of test data and determination of the stress–strain curve

The general procedure of evaluating the test data is described in the following, and specific test
results are presented. The load-elongation curve of this specimen is shown in fig. 9. The values on the
abscissa are the displacements of the loading device, which are not used for any further evaluation.
The camera took 128 photos until failure of the specimen; only 18 of which are taken for the evaluation.
The points where photos have been taken are indicated by crosses, the pictures which have been used
for evaluation of the stress–strain curve are signed by diamonds.

400

300
Load (N)

200

100
Pictures exposed
Pictures evaluated
0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Displacement (mm)
Fig. 9: load–elongation curve of the small rectangular tensile specimen from base metal.

The evaluation of displacement quantities is done by grey value correlation, [9, 5]. This procedure
compares the grey values of one photo from one load step with those of another photo at the next
load step. In particular, it compares the grey values within a small area of e.g. 30 · 30 pixel around
defined node points of a virtual mesh of the first photo with the grey values within an area of e.g.
I. Scheider, W. Brocks, A. Cornec: Procedure for the determination of true stress–strain curves 12

100 · 100 pixel of the second photo. The position where the grey values fit best is taken as the new
position of the respective node determining its displacement from one load step to the next.
In a first step, a coarse virtual mesh of 20 · 40 nodes is spread over the whole visible range of the
specimen. It is used for the determination of the original coordinates of the necking section as well as
of global stretching,i.e. load line displacement. Note that this mesh is not applied mechanically but
only a posteriori by the evaluation programme. As the nodes of this mesh are moving, they represent
material points. The displacements of the nodes are calculated step by step. Fig. 10 displays the
specimen surface with the deformed mesh shortly before fracture of the specimen.

Fig. 10: Moving mesh of sampling points for the displacment evaluation, close to failure of the
specimen.

In a second step, a fine mesh is applied to the necking section of the specimen to get the width
reduction and the strain field. The respective meshes in the initial state and the final state before
fracture are shown in fig. 11. The width reduction is evaluated from a polynomial fitting of the
displacements of the edge nodes and determination of its extremum value. The displacement values
of all the nodes are is approximated by biquadratic smoothing functions from which the strain values
are calculated according to eq. (13).

a) b)
Fig. 11: Fine mesh of sampling points in the area of local necking for the evaluation of width
reduction and strain at the necking center — a) in the initial state, b) close to failure
of the specimen.
I. Scheider, W. Brocks, A. Cornec: Procedure for the determination of true stress–strain curves 13

Finally, a stress–strain curve is generated using eq. (24), see fig. 13. The curve reaches strain values
up to ε̄ = 1. That value lies within the valid range of the proposed procedure. This stress–strain
curve was then applied for a numerical simulation of the tensile test. The load–elongation curve of
this simulation shows good agreement with the experimental curve, see fig. 12. At Δb = 0.65 mm
the specimen has failed unstable, which can of course not be predicted by a purely elastic-plastic
simulation.

800
400

sim.
Fmax = 394 N
exp 600
300 Fmax = 386 N

σ (MPa)
Load (N)

400
200

micro bars: test data


200 micro bars: numerical data
100
experiment round bars: test data
simulation
0
0 0.00 0.25 0.50 0.75 1.00
0.0 0.1 0.2 0.3 0.4 0.5 0.6
Δb (mm) ε (-)

Fig. 12: Comparison of the load – width re- Fig. 13: stress–strain curve of a structural
duction curve between experiment steel, obtained from tests and sim-
and simulation ulation of small rectangular speci-
mens and from tested round tensile
bars.

Using the results of the simulation, the material’s stress–strain curve has been re-evaluated with the
proposed method, namely eq. (24). Fig. 13 shows that the two curves from experimental and numerical
data, the latter using the first one as input, lie closely together, thus giving an additional verification
of the procedure.
Experiments with round tensile bars (D = 6 mm) have also been conducted, but without diameter
reduction measure technique. A comparison between both stress–strain curves up to the maximum
load are shown in fig. 13. Higher strain values are not recorded for the round tensile bar, since no
width reduction has been measured. However, the fracture strain of the round tensile bar is similar
to that of the rectangular specimen.
The difference between micro-flat tensile specimen and round specimen is 20 MPa in Yield strength
and 30 MPa at maximum load. The reason for the difference, which already appears in the initial
yield strength, might be slight changes in the material properties due to the different machining of
the specimen. The spark eroding induced a high heat input to the micro-flat tensile specimen, which
may lead to phase transformations at the surface of the specimen. For future uses of the specimen
preparation this has to be taken into account.

5 Conclusions
Modern digital cameras and a digital image correlation software allow for determining the displacement
and strain fields on the surface of test specimens - or even of structures. Due to the high resolution of
digital images and accuracy of the software one can determine global quantities as well as local fields.
I. Scheider, W. Brocks, A. Cornec: Procedure for the determination of true stress–strain curves 14

A procedure has been proposed and verified to utilize local strain field measurements for the determi-
nation of true stress–strain curves. The method requires the following experimental data: load, width
reduction and the strain in the center of the neck. The method works up to very high strain values.
Even without the strain measurement, a sufficient accuracy up to an equivalent strain of ε̄ = 0.8 can
be achieved. The experimental verification shows that the procedure gives accurate results and is easy
to use. The accuracy depends mainly on the resolution of the digital camera. The number of points
defining the material curve can be chosen by the user.
The procedure has also been applied for the determination of the stress–strain curve in the gradient
of a laser welded plate (in fact, it is the same plate used in section 4, and the results are shown in [4]).
The extension of the procedure to other aspect ratios is easy to perform and will be investigated in
the future.

References
[1] ABAQUS (2002) ABAQUS User’s manual, Version 6.3 Hibbit, Karlson and Sorensen, Inc., 1080
Main Street, Pawtucket, RI, USA
[2] ASTM E8 (1999) Standard Test Methods for Tension Testing of Metallic Materials American
Society for Testing and Materials, West Conshohocken, PA, USA
[3] Bridgman, P.W. (1952) Studies in large plastic flow, McGraw Hill; New York
[4] Cornec, A.; Scheider, I. (2001) Modellierungskonzept für die Versagensbewertung von Laser-
schweißverbindungen / Failure assessment of laser weldments based on numerical modelling,
Mat.-wiss. u. Werkstofftech. 32, 316–328
[5] Kieselstein, E., Seiler, B., Dost, M. (1999) DAC - Deformation Analysis by Correlation. In:
Material Mechanics - Fracture Mechanics - Micro Mechanics (Eds. T. Winkler, A. Schubert,
Berlin/Chemnitz
[6] Ling, Y. (1996) Uniaxial true stress-strain after necking, AMP J. Technology 5, 37–48
[7] Kocak, M;, Cam, G;, Riekehr, S.; Torster, F.; Dos Santos, J. (1998) Microtensile test techniques
for weldments. In Weld Mis-match effect, IIW Doc. SC X-F, no. 079-98.
[8] Schmitt, W.; Keim, E.; Sun, D.-Z.; Blauel, J.G.; Nagel, G. (1996) Load-carrying capacity and
crack resistance of a cladding by the sigma-oscillating wire technique, Proceedings of the 22.
MPA-Seminar, Stuttgart
[9] Vogel, D., Kieselstein, E., Dost, M., Michel, B. (1996) Novel experimental methods for high
resolutions analysis by means of correlation-based algorithms Int. Symp. on Local Strain and
Temperature Measurements in non-uniform Fields at Elevated Temperatures, Berlin
[10] Zhang, Z.L.; Hauge, M.; Odegard, J.; Thaulow, C. (1999) Determining material true stress-strain
curve from tensile specimens with rectangular cross section, Int. J. Solids Struct. 36, 3497–3516
[11] Zhang, Z.L.; Odegard, J.; Thaulow, C. (2000) Novel methods for determining true stress strain
curves of weldments and homogenous materials, Proc. 13th European Conference on Fracture
Mechanics, San Sebastian

View publication stats

You might also like