You are on page 1of 16

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

net/publication/266293518

SHAKEDOWN, DISSIPATION AND FATIGUE OF METALS

Conference Paper · July 2015


DOI: 10.13140/RG.2.1.1238.0003

CITATIONS READS

0 663

3 authors, including:

Eric Charkaluk Rian Seghir


French National Centre for Scientific Research Research Institute for Civil and Mechanical Engineering
148 PUBLICATIONS   2,005 CITATIONS    38 PUBLICATIONS   354 CITATIONS   

SEE PROFILE SEE PROFILE

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

Fatigue View project

LEAF: Laser procEssing plAtform for multiFunctional electronics on Flex View project

All content following this page was uploaded by Rian Seghir on 25 July 2015.

The user has requested enhancement of the downloaded file.


7th EUROMECH Solid Mechanics Conference
J. Ambrosio et.al. (eds.)
Lisbon, Portugal, 7–11 September 2009

SHAKEDOWN, DISSIPATION AND FATIGUE OF METALS

Eric Charkaluk1 , Laurence Bodelot2 , and Rian Seghir1


1 Laboratoirede Mécanique de Lille – CNRS
Ecole Centrale de Lille, 59651 Villeneuve d’Ascq, France
eric.charkaluk@univ-lille1.fr; rian.seghir@ec-lille.fr

2 Graduate Aeronautical Laboratories


California Institute of Technology, Pasadena, USA
lbodelot@caltech.edu

Keywords: Fatigue, Dissipation, Shakedown, Digital Image Correlation, Infrared Thermogra-


phy.

Abstract. The crack initiation phenomenon in a metallic material under cyclic loadings is due
to the localization of the deformation in some grains at the microstructure scale which induces
damage. However, the engineering way to design structures conduct to establish macroscopic
fatigue criteria which finally consider mean value of the mechanical fields, at the scale of a
representative volume of material. In the last decades, a lot of phenomenological models have
been proposed in this way, based on macroscopic stress or strain tensors but none are com-
pletely satisfactory, essentially due to the lack of physical basis. An alternative way consists
in considering energy based approach, in particular by studying the evolution of stored energy
during cycling.
In this aim, a first step consists in defining a theoretical framework based on dissipation
which enables to explain previous experimental works studying the evolution of temperature
of specimen during cyclic loading. Such a framework has been proposed by Charkaluk and
Constantinescu and is based on shakedown concepts and a micro-macro transition.
However, due to the heterogeneous behavior of polycristalline metallic aggregates, some
grains can exhibit plastic strains whereas other still undergo a pure elastic response and, in
fatigue, this can conduct to crack initiation in such plastic grains even if the macroscopic be-
havior of the specimen remains elastic. Therefore, the second step consists in the development
of a method enabling to access to a full-field measurement of both kinematical and thermal
fields of a same zone at the grains scale, in order to be able, in future work, to realize energy
balance at the microstructure scale. The principle of such an experimental method supposes
i) the determination of thermal fields thanks to an infrared camera equipped with a high mag-
nification lens, ii) the computation of kinematical fields based on images coming from a CCD
camera and analyzed with an image correlation technique and, iii) the use of a special coating
and a dichroic lens to realise both measurements at the same time, in the same area.

1
Eric Charkaluk, Laurence Bodelot, and Rian Seghir

1 INTRODUCTION
It is now well known that the fatigue phenomenon in metals is essentially associated to the
gliding of dislocations and the creation of slip bands which characterize irreversible damaging
mechanisms at the grains scale. A review of these mechanisms can be founded for example, in
the monograph of [34], in the pioneering works of [28] or in the recent synthesis of [24, 25, 26].
The behavior of mono- and polycrystals is dependent of the loading level and other factors such
as environment, temperature, . . . . For a Representative Elementary Volume of material (REV),

Σ(t), E(t) Σ(t), E(t)

well-oriented
slip plane

Generalized plasticity Localized plasticity plastic inclusion

Figure 1: Plasticity developed at the macroscopic scale in LCF and at the mesoscopic scale in HCF.

they can be summarized using the following scheme (see figure 1):

• if the loading level is very low, only a single slip system can be considered in each grain
with a shear stress below the resolved yield stress on this slip plane. An elastic response is
therefore obtained in all the grains and no slip bands can be observed. The macroscopic
response, on the REV boundaries, is then also elastic.

• if the loading level is increased but is still low, a single slip system can be considered
in each grain. The shear stress might exceeds resolved yield stress in some grains pos-
sessing a properly oriented crystallographic system with respect to the active slip system
and the loading direction. In such grains, the mechanisms of dislocations gliding and
slip bands creation can be activated. At the grain surface, such bands can be observed.
This heterogeneous plastic behavior, depending on the grains orientations, has recently
been quantified by [22] by diffraction measurements. Some observed grains display a
quasi elastic response while others cyclically plastify. However, if this plastic behavior
is confined in a few grains, the macroscopic response, on the REV boundaries, remains
quasi-elastic.

• if the loading level becomes more important, it can induce two effects. On the one hand,
the quantity of plastic grains becomes more important [33] and, on the other hand, mul-
tiple slip systems can be activated in the same grain. Plastic behavior is then no more
confined, is generalized in all the REV, and the macroscopic behavior becomes elasto-
plastic.

The three schemes correspond to three different fatigue domains, which can easily be described
by a Wöhler’s curve, schematically represented on the figure 2. The first case, presenting an

2
Eric Charkaluk, Laurence Bodelot, and Rian Seghir

elastic shakedown at all physical scales, can be easily associated to infinite lifetimes and unlim-
ited endurance. The second case, corresponding to a localized cyclic plastic behavior in some
well oriented grains, conducts to High Cycle Fatigue (HCF) and a limited endurance. Finally,
the last case, associated to a generalized plastic behavior in the REV, conducts to Low Cycle
Fatigue (LCF) phenomenon. As it can be understand, the fatigue mechanisms are the same in all

Figure 2: Schematic representation of the Wöhler’s curve defining the different fatigue domains (HCF and LCF).

the cases, the fatigue domains depending essentially on the spatial expansion of the plasticity.
Therefore, these domains can rather be related to different scales: the grain in a REV in HCF
and the complete material volume in LCF. In order to establish a common approach in HCF and
LCF, Charkaluk et al. [11, 6, 7, 5] proposed to analyze the link between the dissipation gener-
ated by the cyclic plastic activity and the crack initiation; this conducts to consider this plastic
dissipation as a damage indicator, as proposed a long time ago by Halford [15]. However, since
the pioneering work of Farren and Taylor [13], it is well known that, even if plastic work has a
great importance during the metal deformation, the real challenge is to understand and quantify
the partition between plastic work and stored energy in metals. This is also a crucial issue in
fatigue.
Commonly, since the pioneering work of Taylor and Quinney [36], the ratio between the
intrinsic dissipation and the rate of plastic work Φ is often considered as a constant taking
usually values between 0.8 and 0.9. But more recent results show that this ratio is loading
dependant and, under monotonic loading, a threshold is generally observed before a decrease
of the stored part [10].
In fatigue, only a few works are dedicated to this aspect. The main question concerns the ex-
istence or not of a critical cumulated energy before crack initiation and propagation. In order to
answer this question, Wong and Kirby [38] realised low cycle fatigue tests under strain reversed
tension-compression loading on a 6061-T6 aluminium alloy. The plastic work is estimated by
measuring the experimental hysteresis loop and temperatures measurements are realised in or-
der to estimate the mean dissipated energy. The ratio Φ is then calculated at each cycle. Its
evolution show a first short increase before obtaining a quasi-stable value during all the test
but, due to the observed scatter, it is very difficult to conclude to a constant critical value of
cumulated stored energy for all the tests.

3
Eric Charkaluk, Laurence Bodelot, and Rian Seghir

Following the same experimental principle, some fatigue tests on different materials were
realised by Harvey et al. [19, 18]. The same constancy of the cyclic stored energy is observed
from cycle to cycle but the cumulated stored energy seems to be loading dependent. However,
the results obtained on an aluminium alloy [17] show that the cumulated stored energy is pro-
portional to the evolution of the porosity in the material, due to damage evolution. The main
drawback of these previous experimental works is to consider global mean values on specimens
while damage initiation takes place at the grain scale. In fact, many isotropic metallic materi-
als are made of an aggregate of grains following a distribution of crystallographic orientations.
Among them, some are favourably oriented for plastic gliding with respect to the loading axes.
Under mechanical loading, this creates local heterogeneities as these grains can exhibit plastic
strains whereas others still exhibit a pure elastic response. As plasticity triggers the appearance
of slip bands at the surface of the material, heterogeneities of surface’s relief are also visible at
the scale of grains thanks to high resolution imagery techniques. Moreover, plasticity is associ-
ated with a thermal dissipation which can be detected at the structure scale. Hence the material
behaviour under fatigue may be thermally and mechanically different from one grain to another.
In order to realise energy balance in fatigue, thermal and kinematic measurements have also
been performed on both sides of a DP600 steel sample submitted to cyclic loading [9, 8]. The
experimental approach involves two quantitative imaging techniques: digital image correlation
and infrared thermography. By using a variational method, stress fields are deduced from the
displacement fields coming from the digital image correlation. Patterns of deformation energy
per cycle are then determined on the basis of stress and strain data. Then, a local form of the
heat equation is used to derive separately the thermoelastic and dissipative sources. Energy bal-
ances show that around 50 per cent of the deformation energy associated with the mechanical
hysteresis loop is dissipated while the rest corresponds to stored energy variations. Nevertheless
these results are obtained at a relatively macroscopic scale (a few millimetres of spatial reso-
lution) whereas phenomena arise at the grain scale. There is thus a lack of simultaneous data
about the kinematic and thermal behaviours of the material at the mesoscopic scale of damage.
Then, the objective of this communication is the following. After the presentation of the the-
oretical framework conducted to the study of dissipation evolution in fatigue, an experimental
setup enabling an access to fully-coupled measurements of both thermal and kinematic fields
of the same zone at the grain scale will be presented. The full-field measurements methods are
the same as those presented in [9, 8] but high magnification lenses are used in order to increase
the spatial resolution and a special coating and a dichroic mirror are used to be able to realise
simultaneously the measurements at the same time and in the same area [3]. The first results
obtained on monotonic and fatigue cyclic tests will then be presented and discussed.

2 THEORETICAL FRAMEWORK
A straightforward link between mechanical dissipation and fatigue damage stems from the
high cycle metal fatigue theory of Orowan [30], and the followers model of Dang Van [12] and
Papadopoulos [31]. Starting from the observation that the cyclic evolution of isolated grains in
polycrystals submitted to complex loadings can thereby be resumed by the creation of localized
slip bands, a cyclic plastic activity in slip bands and the nucleation of microcracks until the
creation of a macrocrack, Dang Van [12] and Papadopoulos [31] based their models on the next
framework of assumptions: (i) the fatigue damage is controlled by mechanisms at the grain
scale and therefore a description at this mesoscopic scale is necessary; (ii) at this scale, most
of the metallic materials are aggregates of cubic crystals with a random distributed crystallo-
graphic orientations, which can be considered isotropic and homogeneous at the macroscopic

4
Eric Charkaluk, Laurence Bodelot, and Rian Seghir

scale; (iii) among all grains and possible slip planes, only some well oriented slip planes, max-
imizing the shear stress for a given loading path, will develop plasticity and create localized
slip bands inducing crack initiation; (iv) below the fatigue limit, microscopic plastic strains ho-
mogenize to negligible macroscopic plastic strains, which matches the fact that macroscopic
stresses are small with respect to the yield limit; (v) mesoscopic plasticity is determined by
isotropic and kinematic hardening rules and the evolution of the mesoscopic yield limit τs will
initially evolves and then saturates under an imposed cyclic strain loading.
As a consequence, the natural framework of the multiscale approach proposed by Dang Van
et al. [12, 31], is a representative elementary volume (REV) with an elastic macroscopic be-
havior and some plasticity localized only in some grains. The simplest mesoscopic model is
a plastic Eshelby inclusion in an elastic matrix. As described before, the behavior of the plas-
tic inclusion can be resumed to a two stage evolution of its yield limit. Then, the macro- and
mesoscopic stress tensors denoted by Σ and σ, as well as the macro- and mesoscopic strain
tensors denoted by E and ε are linked using a homogenization scheme [4]. One possible and
well adapted is the Self-Consistent (SC) scheme of Kröner. The unlimited endurance condition,
i.e. conducting to infinite lifetimes, corresponds then, in this framework [12, 31], to an elastic
shakedown state in the plastic inclusion. However, since the works of Dang Van [12] and Pa-
padopoulos [31], many investigations were realized on the link between crack initiation and the
concept of fatigue limit: the endurance limit is rather close to the stress amplitude below which a
short crack initiated in an individual grain is unable to propagate toward the neighboring grains.
Then, the definition of the endurance limit proposed by Dang Van [12] and Papadopoulos [31],
corresponding to the non initiation of cracks in individual grains, is much more a lower bound
of the real endurance limit of the material.
The Dang Van and the Papadopoulos criteria are both based on a Lin-Taylor homogenization
assumption [23, 35]. The Dang Van criterion is defined in terms of the hydrostatic pressure, σh ,
and of the mesoscopic resolved shear stress τ . The Papadopoulos criterion depends also on p
and on the radius k of the smallest hypersphere enclosing the deviatoric stress path. In this last
case, the crack nucleation condition is deduced from the shakedown theorem of Mandel [27].
These criteria can be written as:

max [kτ (n, t) − τm (n)k + αdv σh (t)] ≤ βdv Dang Van (1)
n,t
max (k(t) + αpa σh (t)) ≤ βpa Papadopoulos (2)
t

The material parameters αdv , αpa , βdv and βpa are defined in terms of the endurance limits in
reversal torsion tests σDt and bending experiments σDb . n is the normal to the shear plane.
τm (n) is the vector which points to the center of the smallest circle circumscribing the path
described by the tip of the shear stress vector τ (n, t) on the plane defined by the unit normal
vector n. In the equation (2), k(t) is defined as:
r
3
k(t) = (s(t) − sm ) : (s(t) − sm )
2
where sm is the mean stress deviator, which corresponds to the co-ordinates of the center of the
smallest hyper-sphere enclosing the path described by the stress deviator s(t).
Let us only remark that a shear term in both expressions assures that no microcracks are
formed in the slip bands. Generalizing the observations of Winter [37] that PSB’s begin to
appear when the stress amplitude reaches a limit value denoted by ks , a crack initiation criterion

5
Eric Charkaluk, Laurence Bodelot, and Rian Seghir

is defined by Papadopoulos [31] as:


k ∗ = max k(t) ≤ ks
t

The ks limit value is larger than the yield limit σy of the crystal (ks > σy ), but smaller than
the macroscopic yield limit and the endurance limit. One can note that the pressure term takes
into account the local grain distribution and assure a good match with experiments but is not
deduced from a precise homogenization reasoning as the shear term. The plasticity developed

Figure 3: Theoretical evolution of the mesoscopic cumulated mechanical dissipation dcumul .

Figure 4: Theoretical evolution of the mesoscopic dissipated energy per cycle ∆w versus the previous defined
stress k ∗ .

in some grains will produce an instantaneous mesoscopic plastic dissipation dp . By using the
previous framework, dp can be computed from:
dp = σ : ε˙p (3)
As a consequence, the Papadopoulos fatigue criterion can be redefine in terms of dissipation. A
sketch of the ideas presented next is drawn on figure 3.

6
Eric Charkaluk, Laurence Bodelot, and Rian Seghir

1. If k ∗ ≤ σy , with σy the yield limit of the grain, no plasticity occurs at the mesoscopic
scale. As a consequence a fully elastic behavior is assured and no crack initiation will
occur. Moreover:
k ∗ ≤ σy if and only if dp = 0 (4)

2. If σy ≤ k ∗ ≤ ks , the grain plastifies but reaches an elastic shakedown state and, there-
fore, no crack initiation can be observed. As an elastic shakedowned state is attained,
the shakedown theorem assure that the cumulated plastic work Wp (∞) is bounded, as
demonstrated by [29]:
Z ∞ Z ∞
Wp (∞) = ˙p
σ(t) : ε (t)dt = dp (t)dt < ∞
0 0

Therefore dp leads to zero with the number of loading cycles and it can be stated that:

σy ≤ k ∗ ≤ ks if and only if lim dp (t) = 0 (5)


t→∞

3. ks < k ∗ , elastic shakedown is not reached and a crack will initiate in one of the slip bands
of the misoriented grains. The material assumptions of the fatigue criterion imply that the
isotropic hardening is saturated at ks and only kinematic hardening is activated. More-
over, applying Halphen’s shakedown theorem for elastoplastic structures with kinematic
hardening [16] indicates that under a periodic loading, the structure will have a periodic
solution for stress and strain tensors. As elastic shakedown is not possible anymore, the
misoriented grain reaches a plastic shakedown state. Therefore, in this case, it can simply
be proved that the plastic work over a loading cycle:
Z
∆wp = dp dt
cycle

is constant. Then,

k ∗ > ks if and only if lim ∆wp = constant (6)


N →+∞

The three types of evolution of the plastic dissipation: (4), (5) and (6) are schematically rep-
resented on figure 3 and 4. The plots represent the evolution of the cumulated plastic work
Wp (∞) versus the number of cycles of a fatigue test and the plastic work per cycle ∆wp versus
the applied load represented by k ∗ respectively. Figure 4 shows that the stabilized dissipative
regime depends on the loading: up to a critical load, denoted ks and related to the definition of
shakedown in the fatigue criterion, no dissipation can be observed. As loading is increased, a
plastic shakedown regime is obtained which conducts to limited endurance or low cycle fatigue.
Both figures 3 and 4 relate directly the fatigue regime with different plastic dissipative states.
Unfortunately, direct experimental measurements of the plastic dissipation at the mesoscopic
scale are not possible and, moreover, this measure has to be realise at the microstructure scale
of the REV. The following sections will then present a new original set-up developed in this
aim.

7
Eric Charkaluk, Laurence Bodelot, and Rian Seghir

3 EXPERIMENTAL PROCEDURE
Following the work of Chrysochoos and co-authors [9, 8], in order to determine intrinsic
dissipation and to realize energy balance, thermal and kinematic measurements have also to be
performed. Then, the proposed experimental approach involves two quantitative imaging tech-
niques: digital image correlation and infrared thermography. However, these previous results
are obtained at a relatively macroscopic scale (a few millimetres of spatial resolution) whereas
fatigue phenomena arise at the grain scale. Simultaneous thermal and kinematic measurements
in the same area have then to be performed at the microstructure scale. This is the objective of
the proposed set-up.

3.1 Kinematic and thermal measurements


Images of the sample are grabbed during the tests thanks to a CCD Camera Jai CV-M4+.
This camera is made of a matrix of 1368 by 1024 detectors sensitive to visible wavelengths
from 0.38 to 1 µm. A TAMRON 23FM50SP 50 mm lens associated to extension tubes enables
a high spatial resolution of 6.5 × 6.5µm2 per pixel leading to a working zone of approximately
8.9 × 6.7mm2 . Although no spatial distortion near to the borders is observed, the working zone
is reduced to 5 × 5mm2 , in order to match the infrared data.
Displacement and strain fields are then obtained through DIC performed on the images cap-
tured during the mechanical deformation of the sample. The displacement field of one image
is determined with respect to a reference image thanks to the DIC software CorreliLM T de-
veloped at LMT laboratory (Cachan, France). The principle is the following. The reference
two-dimensional image is divided into square regions (called Zone of Interest, ZOI); then, the
DIC software matches these ZOI between one image at a time ti and the reference image at a
time t0 , thanks to an algorithm based on Fast Fourier Transformation [20]. The result of the
DIC analysis is then a two-dimensional displacement field for each time ti . The principle of the
correlation is based on pattern recognition. The pattern has to be randomly obtained and has to
cover a wide range over the greyscales. In some particular cases, such a random pattern appears
naturally on the tested material so that correlation can be performed directly on images of the
raw material [1]. In more general cases, the pattern has to be created artificially: the sample is
usually covered with black and white paint in order to obtain a random speckle. In the present
case, a coating has to be applied since the sample is polished as a mirror.
Thermal measurements are performed thanks to a Cedip Jade III infrared camera. Its infrared
sensitive part is a Focal Plane Array (FPA) of 320 by 240 detectors working in the spectral
range from 3 to 5 µm. This camera is used with a high magnification lens known as G1 since
the spatial resolution matches the size of one detector, which is 30 × 30µm2 . This leads to a
9.6 × 7.2mm2 observation zone but as the G1 lens induces optical distortions, the working zone
is here narrowed to 5 × 5mm2 .
To perform accurate quantitative infrared thermography, a precise calibration must be per-
formed, the emissivity of the object has to be determined and some precautions must be taken.
Firstly, a calibration law, which gives a link between temperatures and thermal emissions, has to
be established by exposing the detector array of the camera in front of an extended blackbody
at different controlled temperatures. In this work the camera is not used conventionally: the
calibration law is not averaged over the whole array but is determined detector by detector [21]
to enhance the detection of local thermal heterogeneities. Moreover, as the calibration is influ-
enced by many parameters (distance to the object, lens used, working temperatue, . . . ), a new
calibration is necessary for each test configuration [32] and each time the camera is switched

8
Eric Charkaluk, Laurence Bodelot, and Rian Seghir

on. Secondly, the emissivity of the object has to be determined. This can be realized through
a classically used method which consists in a comparison between the radiation of the object
and the one of an extended blackbody, both being at the same temperature [14]. At last, for
a better accuracy, the camera is turned on more than 4 hours before beginning any measure-
ments including calibration in order to avoid facing with thermal drift. Every element close to
the camera is covered by black tissue or painted with high emissivity black paint to keep clear
of any reflection. The area around the camera is also prevented from any reflections from the
environment thanks to black tissue.

3.2 Principle of the coupling


As the objective of the tests is to catch simultaneously the kinematic and thermal fields at
the scale of the microstructure, it is not possible to realise kinematic measurements on one
face and thermal measurements on the other, as it was previously proposed [9, 8]. As the
microstructure is different from one face to another, measurements have to be done on the same
face and at the same instant. In order to realize a simultaneous observation of the same zone
by both cameras, a ”filter-mirror” is placed in front of the sample making an angle of 45◦ with
respect to the sample’s surface. On the one hand, this filter-mirror behaves like a filter since
it only lets the infrared radiation go towards the IR camera placed ahead of the sample. On
the other hand, it also behaves like a mirror since it reflects the rest of the radiation, among
which the visible radiation, in the direction of the CCD camera. The ”filter-mirror” allows
a simultaneous observation of the same zone and high magnification lenses give high spatial
resolution images. However, reaching fully-coupled measurements supposes that the coating
applied on the sample meets conditions for both DIC and IRT. As a matter of fact, DIC has to
be performed on optical images of random and contrasted aspect, covering a large range of grey
levels, whereas infrared measurements have to be performed on high emissivity objects, which
can be obtained thanks to a dark and homogeneous coating. As a result, the coating must be a
speckle when observed by the CCD camera as well as dark and homogeneous when observed
by the IR camera. As a consequence, a special coating has been developed; it is suitable,
at the working scale, for both techniques and thus permits fully-coupled measurements. The
emissivity of this coating has been characterized and its mean value is 0.927 with a standard
deviation equal to 0.009. The choice of the material is also a crucial issue. First of all, the
material’s thermal diffusivity (ratio of thermal conductivity to volumetric heat capacity) must
be as low as possible so that heat generation is not dissipated nor homogenized too quickly
with respect to the frame frequency of the IR camera. This is actually essential to be able
to capture the thermal field’s heterogeneities. Secondly, single phase steel is preferred as the
thermal properties should be considered identical in all grains. Finally, the grain size has to be
large enough so that several pixels of both IR and CCD matrixes can stare at one grain. This last
condition leads us to use a material having a mean grain size of roughly 100 µm. In order to
respect as well as possible all the above-mentioned constraints, the chosen material is an AISI
316L, whose thermal diffusivity is equal to α = 0, 4.105 m2 s−1 , since it meets the two first
criteria. Its initial grain size is of about 10 µm, which is not enough. However, in the case of
AISI 316L, this size can easily be raised up to about 130 µm (considering twins) thanks to a
heat treatment at 1,200◦ C for 2 hours, immediately followed by a water quench. Samples of
the material before and after the heat treatment are chemically etched and examined under a
microscope to underscore the grain growth. Specimens are thus AISI 316L 2 mm thin plates
having a gage area of 10 × 10µm2 . They are mechanically polished up to 1 µm on faces and up
to grit 4,000 SiC paper on the edges.

9
Eric Charkaluk, Laurence Bodelot, and Rian Seghir

4 CASE OF A MONOTONIC TENSION TEST


4.1 Test and mean fields results
Experiments are performed on an electro-mechanical Instron 4508 machine which enables
tests up to 300 kN on a speed range from 0.1 to 100 mm/min. A displacement controlled
tension test has been realised at a strain rate equal to ε̇ = 5.10−3 s−1 . The evolution of the
nominal stress in the specimen can be observed on figure 5. During the test, displacement, strain
and temperature fields are recorded in the gage zone thanks to the fully-coupled measurement
technique presented in the previous paragraphs. The load applied to the sample is also recorded
during the tests. In order to disturb as less as possible the thermal measurements, the working
zone is protected from its environment as it has already been explained; light is turned off
and the room temperature is kept steady thanks to air conditioned. The mean values of the
thermal fields from the initial instant and also of the strain fields in the loading direction are
first computed on a central zone of 5 × 5mm2 and their evolution are presented on figure 5.
At the beginning of the test, the evolution of the mean temperature of the central zone exhibits

Figure 5: Time evolutions, during the monotonic tension, of the nominal stress and of mean value of the fields
of thermal variation from the initial instant and of strain in the loading direction. The mean value of the fields is
determined on the central 5 × 5mm2 studied zone.

the effect of the thermoelastic coupling. This effect consists in a temperature decrease of the
specimen under tension. In the present case, this decrease is first linear until the end of the
elastic domain. At this particular instant, when strain hardening occurs, the rate of temperature
decrease and then, a temperature increase is observed during all the plastic part of the loading
curve which is the consequence of the dissipative character of the plastic mechanisms. Then, at
the end of the loading, the temperature increase stops and an equilibrium temperature is reached.

4.2 Kinematic and thermal fields study


At the end of the test, the coating has been removed by plunging the specimen in an alcohol
bath, then an ultrasound acetone one. The thermal and strain fields can then be compared
with this final microstructure. Concerning the displacement fields, correlation computations are
realised with CorreliLM T for a ZOI of 16 × 16 pixels. The strain fields are then obtained by
derivation. This superposition shows that high strained area correspond to grains which exhibits
many coarse visible slip bands [3]. On the opposite, area with low strained level correspond to
grains with no visible high plastic activity. The same principle is now applied to the thermal
fields. A link between thermal emission and plastic activity is more dlicate as temperature
information is instantaneous and therefore does note reflect the history of plastic deformation. In

10
Eric Charkaluk, Laurence Bodelot, and Rian Seghir

the present case, the final thermal fields correspond to the more deformed zone of the specimen.
In order to establish potentially more evident links, the temporal evolutions of zones with very
different strain levels can now be studied.

Figure 6: Evolutions of strain and temperature of both points of the studied zone, (A) highly deformed and (B)
lowly deformed.

4.3 Temporal analysis


Two points named A and B (see figure ??) are chosen in area which present at the instant
P2 (see figure 5) respectively high and low strain levels. The temporal evolutions of the strain
and thermal variation from the initial instant for points A and B are presented on figure 6.
The point A which present all over the loading the maximal strain level is situated in a zone
where the plasticity was very intense, localized in a small grains where slip lines are very
numerous. On the opposite, the point B whose deformation is lower, is situated in a zone
where plasticity activity seems to be negligible. From the thermal point of view, both points
present a temperature decrease very similar during the elastic part of the loading. Then, as
temperature increases at the beginning of the plastic activity, thermal evolution of points A
and B are diverging. However, the gap is less pronounced than with the strain curves which
is obviously due to the thermal diffusivity which induces an homogenization of the thermal
gradients. This type of analysis can now be applied in the case of a cyclic loading.

5 CASE OF A CYCLIC TENSION-TENSION TEST


5.1 Test and mean fields results
Cyclic experiment is realized on the same electro-mechanical machine and the test carried
out in this work is a 2.8 Hz cyclic test at a mean stress of 85 MPa and at Rσ = 0.1. The
stresses due to the cyclic loading applied on the sample are plotted on figure 7 as well as the
evolution on the mean values of the thermal fields from the initial instant and also of the strain
fields in the loading direction, computed on a central zone of 5 × 5mm2 On this figure, a
phase opposition can be observed between thermal and stress signals. This phenomenon is
linked to thermoelastic coupling. As a matter of fact, the temperature decreases as the sample is
submitted to an increasing tensile stress and then increases again when the stress is diminishing.
The different fields will be studied at point M 20 which corresponds to the maximal stress level
on the 20th cycle. A more precise observation of the mean thermal evolution show that the mean
value on a cycle increase from cycles to cycles. In fact, the test loading is always under tension
which induce a temperature decrease due to the thermoelastic coupling. Therefore, during the

11
Eric Charkaluk, Laurence Bodelot, and Rian Seghir

Figure 7: Evolutions of nominal stress, mean strain in the loading direction and mean temperature variation from
the initial instant in the central studied zone.

test, the temperature is coming from this state to an equilibrium temperature associated to the
room environnement.

5.2 Kinematic and thermal fields study


Strain field corresponding to the instant marked M 20 on figure 7 are computed by DIC on
the same working zone of 5 × 5mm2 in the middle of the sample’s gage area, with the same ZOI
of 16 × 16 pixels. At instant marked M 20 on figure 7, the mean deformation over the gage zone
reaches a value of 0.053%. The strain fields obtained at these instants exhibit heterogeneities:
some areas are under low deformation with a strain value of 0.025% whereas other areas are
submitted to higher strains up to more than 0.2%. This is the sign of the local plastic hetero-
geneity. The heterogeneities of thermal fields captured at instant M 20 can’t however be related
to the strain field when observed at the naked eye [2]. This is due to the high cycle fatigue
loading level which implies a very low thermal level which seems to decrease quickly due to
thermal diffusivity. This is confirmed by the temporal analysis in the next section.

5.3 Temporal analysis


Two points named A” and B” are chosen in the central area which present at the instant
M 20 respectively high and low strain levels. The temporal evolutions of the strain and thermal
variation from the initial instant for points A” and B” are presented on figure 8. The point A”
which present all over the loading the maximal strain level is situated in a zone which exhibits
little slip lines. On the opposite, the point B” whose deformation is lower, is situated in a
zone where no plastic activity seems to be observable. From the thermal point of view, both
points present quite the same temperature evolution. This conclusion is the same as in the
case of the monotonic tensile test but is reinforced as the loading level seems to be too low to
authorize precise conclusion on a link between plastic activity and thermal emission. However,
these first results show the feasibility of such measurements at the microstructure scale even if
some requirements are necessary: a sufficiently high loading level, cameras with high working
frequencies in order to be able to increase the applied strain rate. Under such conditions, it
seems possible to realise energy balance at the grain scale under monotonic or cyclic loadings.

6 CONCLUSION
The objective of these recent methods, developed in order to determine the stored part of
the energy, consists in studying the possible link between this energy and damage. The main

12
Eric Charkaluk, Laurence Bodelot, and Rian Seghir

Figure 8: Evolutions of strain and temperature of both points of the studied zone, (A”) more deformed than (B”).

interest is to be able to develop constitutive and damage laws thermodynamically admissible, i.e.
with respect to the energy balance presented in the introduction. Former and recent comparison
between models and experimental results show that classical models are generally not able
to verify such relations. This is therefore a new important challenge in the next years which
imply new set-up as the proposed one in order to be able to catch simultaneously kinematic
and thermal fields at the microstructure scale which is the scale where strain localisation and
damage initiate, in particular under cyclic loadings.

References
[1] A. El Bartali, V. Aubin, and S. Degallaix. Fatigue damage analysis in a duplex stainless
steel by digital image correlation technique. Fat. Frac. Engng. Mat. Struct., 31:137–151,
2007.
[2] L. Bodelot, L. Sabatier, E. Charkaluk, and P. Dufrénoy. Experimental determination of
fully-coupled kinematical and thermal fields at the scale of metal grains under cyclic load-
ing. Advanced Engineering Materials, 2009. in-press.
[3] L. Bodelot, L. Sabatier, E. Charkaluk, and P. Dufrénoy. Experimental set up for fully-
coupled kinematic and thermal measurements at the microstructure scale of an AISI 316L
steel. Mat. Sci. Eng. A, 501:52–60, 2009.
[4] M. Bornert, T. Bretheau, and P. Gilormini. Homogénéisation en mécanique des matériaux,
Tome 1, matériaux aléatoires élastiques et milieux périodiques. Hermès Sciences, 2001.

13
Eric Charkaluk, Laurence Bodelot, and Rian Seghir

[5] E. Charkaluk. Energy as a damage factor in fatigue. MP Materialprüfung, 51(5):269–275,


2009.

[6] E. Charkaluk and A. Constantinescu. Dissipation and fatigue damage. MP Materi-


alprüfung, 46(10):524–530, 2004.

[7] E. Charkaluk and A. Constantinescu. Dissipative aspects in High Cycle Fatigue. Mech.
Mat., 41:483–494, 2009.

[8] A. Chrysochoos, B. Berthel, F. Latourte, A. Galtier, S. Pagano, and B. Wattrisse. Local


energy analysis of high-cycle fatigue using digital image correlation and infrared thermog-
raphy. J. Strain Anal., 2008.

[9] A. Chrysochoos, B. Berthel, F. Latourte, S. Pagano, B. Wattrisse, and B. Weber. Local


energy approach to steel fatigue. Strain, 2008.

[10] A. Chrysochoos, O. Maisonneuve, G. Martin, H. Caumon, and J.C. Chezaux. Plastic and
dissipated work and stored energy. Nucl. Engng. Design, 114:323–333, 1989.

[11] A. Constantinescu, K. Dang Van, and H. Maı̈tournam. A unified approach for low and high
cycle fatigue based on shakedown concepts. Fat. Frac. Engng. Mat. Struct., 26(6):561–
568, 2003.

[12] K. Dang Van. Sur la résistance à la fatigue des métaux. Sciences Technique Armement,
47(3), 1973.

[13] W. S. Farren and G. I. Taylor. The heat developped during plastic extension of metals.
Proc. Royal Soc. A, 107:422–451, 1925.

[14] G. Gaussorgues and S. Chomet. Infrared Thermography. Chapman and Hall, 1994.

[15] G. R. Halford. The energy required for fatigue. J. Mater., 1(1):3–18, 1966.

[16] B. Halphen. L’accommodation des structures élastoplastiques à écrouissage cinématique.


C.R. Ac. des Sciences-série A, 283(10):799–802, 1976.

[17] D. P. Harvey II and R. J. Bonenberger. Detection of fatigue macrocracks in 1100 alu-


minium from thermomechanical data. Eng. Frac. Mech., 65:609–620, 2000.

[18] D. P. Harvey II and R. J. Bonenberger. Influence of flow stress on damage accumulation


in 1100 aluminium subjected to cyclic straining. J. Mat. Sci. Lett., 19:2189–2192, 2000.

[19] D. P. Harvey II, R. J. Bonenberger, and J. M. Wolla. Effects of sequencial cyclic and
monotonic loadings on damage accumulation in nickel 270. Int. J. Fatigue, 20(4):291–
300, 1998.

[20] F. Hild. CorreliLMT: a software for displacement field measurements by digital image
correlation. Internal report 254 - lmt, cachan, 2002.

[21] V. Honorat, S. Moreau, J.M. Muracciole, B. Wattrisse, and A. Chrysochoos. Calorimet-


ric analysis of polymer behaviour using a pixel calibration of an IRFPA camera. QIRT
Journal, 2:153–172, 2005.

14
Eric Charkaluk, Laurence Bodelot, and Rian Seghir

[22] A. M. Korsunsky, K. E. James, and M. R. Daymond. Intergranular stresses in polycrys-


talline fatigue: diffraction measurement and self-consistent modelling. Eng. Frac. Mech.,
71:805–812, 2004.

[23] T. H. Lin. Analysis of elastic and plastic straines of a FCC crystal. J. Mech. Phys. Solids,
5:143, 1957.

[24] P. Lukas and L. Kunz. Cyclic slip localisation and fatigue crack initiation in fcc single
cristals. Material Science and Engineering A, 314:75–80, 2001.

[25] P. Lukas and L. Kunz. Specific features of high-cycle and ultra-high-cycle fatigue. Fat.
Frac. Engng. Mat. Struct., 25:747–763, 2002.

[26] P. Lukas and L. Kunz. Role of persistent slip bands in fatigue. Phil. Mag., 84(3-5):317–
330, 2004.

[27] J. Mandel, J. Zarka, and B. Halphen. Adaptation d’une structure élastoplastique à


écrouissage cinématique. Mech. Res. Communications, 4(5), 1977.

[28] H. Mughrabi. Fatigue crack initiation by cyclic slip irreversibilities in high cycle fatigue.
In Fatigue mechanisms : advances in quantitative measurement of physical damage -
ASTM STP 811, pages 5–45, 1983.

[29] Q.S. Nguyen. On shakedown analysis in hardening plasticity. J. Mech. Phys. Solids,
(51):101–125, 2003.

[30] E. Orowan. Theory of the fatigue of metals. Proc. Royal Soc., 171:79–106, 1939.

[31] I. V. Papadopoulos. Fatigue polycyclique des métaux : une nouvelle approche. Phd thesis,
spécialité : Mécanique, Ecole des Ponts et Chaussées, France, 1987.

[32] H. Pron and C. Bissieux. Focal plane array infrared cameras as research tools. QIRT
Journal, 1:229–240, 2004.

[33] M. Sauzay. Effet de l’anisotropie élastique cristalline sur la distribution des facteurs de
Schmid à la surface des polycristaux. C.R. Méca, 334(6):353–361, 2006.

[34] S. Suresh. Fatigue of materials – second edition. Cambridge University Press, Cambridge,
1998.

[35] G. I. Taylor. Plastic strains in metals. J. Inst. Metals, 62:307, 1938.

[36] G. I. Taylor and H. Quinney. The latent heat remaining in a metal after cold working.
Proc. Royal Soc. London A, 143(849):307 – 326, 1934.

[37] A. T. Winter. A model for the fatigue of copper at low plastic strain amplitude. Phil. Mag.,
30(4):719–738, 1974.

[38] A. K. Wong and G. C. Kirby III. A hybrid numerical/experimental technique for deter-
mining the heat dissipated during low cycle fatigue. Engng. Fract. Mech., 37(3):493–504,
1990.

15

View publication stats

You might also like