Sie sind auf Seite 1von 41

1.

INTRODUCTION
The long time duration of structures at elevated temperatures is governed by creep
deformations. To design structures for an expected life of several years, it appears crucial
to understand how fracture is governed by plastic flow under nominal stresses and
temperatures that are maintained constant. In the early fifties, creep experiments on
metallic and polymeric materials have demonstrated the complex dependence of fracture
mechanism with respect to the applied stress. The experimental observations are
summarized in Figure 1 drawn from the work of Larson and Miller (1952). A thin bar of
stainless steel at elevated temperature is subjected to a constant tensile force. Figure 1
shows the dependence between the nominal stress and the time to rupture in log-log
scale. Similar results are given in the case of brittle low alloy steel at 500 oC in the work
of Richard (1955) (reported by Odqvist, 1971). Two linear branches are observed in these
diagrams, with two markedly different slopes. The time to failure is strongly stress
sensitive on the upper branch while a smaller stress sensitivity is observed for the lower
branch. The change of slope in these log-log diagrams appears to be a salient feature of
the creep experiments. In addition, two different modes of fracture are observed
depending on the stress level (Odqvist, 1971)

Figure 1. Experimental results of Larson and Miller (1952) concerning a cylindrical bar under

Tensile creep at elevated temperatures.


The nominal applied stress is represented in terms of the time to rupture in a loglog diagram. The material is a stainless steel. Rupture time is in hours.
(a) A ductile mode of failure is associated to the large stresses on the upper branch; then,
fracture occurs by necking accompanied in general by an important damage.
(b) A brittle failure mode corresponds to the small values of the stress on the lower
branch. No necking develops; rather the fracture is initiated by crack propagation
preceded by an important damage development. The denomination brittle used in the
literature may be considered as misleading. The failure is actually accompanied by a
growth of damage which is controlled generally by plastic flow. Similar observations are
reported from creep tests on polymeric materials. For example, the work of Barton and
Cherry (1979), experimental results are obtained concerning the time to failure of a
polyethylene pipe sustaining constant internal pressures; see also Richard (1959).
Different failure modes are again observed depending on which branch is considered. For

large pressures, fracture of the tube by development of an aneurysm is observed (ductile


failure). On the contrary, for low pressures the tube fails by slow crack growth with no
appearance of an aneurysm (brittle failure).
1.1 Initial and secondary creep
The common ground of all dislocation creep theories is the knowledge that the
material is hardened with the deformation and is softened with time (while heated). These
two procedures take place simultaneously and define the strain. This idea was first
formulated by Bailey and Orowan and was successively developed by a large number of
scientists. At a high temperature, usually about the one-third of the absolute melting
temperature, the dislocations acquire a new degree of freedom. Except for gliding, there
is also climbing, and therefore the dislocations are not obliged to move only on their slip
planes. These results in the gradual release of dislocations previously created with the
strain. The structure of the dislocations is subjected to the so-called recovery. This means
that if a dislocation is held by an obstacle, the recovery procedure will release it, allow it
to slide down to the next obstacle, the glide step of the dislocation being responsible for
almost the total strain. The mechanisms which are based on this succession of glide climb
of dislocations are referred to as hardening recovery mechanisms. The characteristic
difference which distinguishes these mechanisms from those of plastic flow under lower
temperatures (which mechanisms may also be thermally activated) is that the procedure,
at an atomic level, is rather the diffusive movement of the atomic voids towards or from
the dislocation, which glides more, than the gliding of the dislocation as a whole . The
modern unified way of describing the gliding phenomena and of dislocation hardening
and recovery, is the theory of a dislocations network and of an internal back stress.
In accordance with the experimental observations, it has been assumed that the
dislocations are arranged to form a network. The creep procedure consists of continuous
events of recovery and hardening. The coherence of the network is ensured by the
repulsive and attractive forces among the dislocations. As a result of the applied stress
and of the thermal fluctuation, some of the dislocations will escape from the network and
will slide a certain distance down to the point where they encounter some obstacle
3

(dislocation, second-phase particle, etc.). During their movement the strain and hardening
increase because the dislocations subjected to stress, increase their length and, therefore,
their density. The recovery procedure takes place simultaneously. The force for the creep
of the dislocations results from the linear stress of the dislocations which have been bent
by the obstacles. After the climbing, the procedure is repeated. At the beginning of the
loading (initial creep), many of the dislocations loosely connected to the network will
move, and therefore the creep rate will, initially, be very high. But the number of easily
escaping dislocations decreases, gradually, with time. This results in the fact that the
creep strain increases with a diminishing rate. At this stage, the hardening is dominant
and the dislocation density increases. However, the recovery trend increases with the
increase in the dislocation density, therefore, the recovery rate increases and finally, a
situation results where the two procedures balance one another while the dislocation
density and the creep rate remain constant (secondary creep). The time required for a
dislocation to overcome an obstacle has to do with the flow of atomic voids at jogs of its
length. What is characteristic is that the higher the stress and the temperature, the faster is
the climb. The time required for a dislocation to slide depends on the relationship
between the applied stress and the stress coming from the dislocations network (elastic
dislocation field). Dissolved atoms (foreign atoms) which are attracted to the elastic
dislocation field and which attempt to be diffused while the dislocation slides create a
friction force with the lattice and can decelerate its movement. In a material that was
strengthened with second-phase particles and the ratio of the particles volume with
respect the total volume is large enough, the opposing stress consists of a component that
is due to the particles. The model of a dislocation sliding on the sliding plane is
represented by a load sliding on a plane, r& is the friction stress in the lattice during the
movement of the dislocation because of foreign (interstitial) atoms.

is the internal

stress of particles coming from the elastic field of the distributed second-phase particles.
The stress that is applied on the dislocation and that is due to the network of the rest
dislocations, is called the internal back stress and is represented by the symbol
general, it can be assumed that

. In

when the atomic size of the interstitial atoms is not

considerably different from that of the lattice atoms. While


4

is taken into account only

in metallic materials heavily hardened with particles. In order for strain to take place, the
dislocation has to glide under the applied external stress which is herein represented by
I

, glide will not take place for a limited period of time until the internal back stress

is self adjusted to new values. The above formulation does not include existing inelastic
phenomena related to the grain boundary. But as shown in a previous study, these
phenomena can be neglected. At the microscopic scale, the hardening is controlled by the
rate at which the dislocations approach each other, while the recovery takes place through
rearrangement (and counterbalancing) of the dislocations. Therefore, the applied stress is
composed of two terms: the mean internal back stress, which is related to the recovery
hardening equilibrium, and the mean effective stress that is responsible for the glide. The
concept of the mean internal back stress (and, respectively, of the mean effective stress)
has been proven to be a very important phenomenological structure parameter because it
is directly related to the dominant procedure of dislocations movement.

1.2 Tertiary creep


The creep acceleration at the third stage is due to the creation and joining of
cavities at the boundaries of grains, that is, the fracture in creep is, generally,
intercrystalline The cavities may be created at the beginning of creep, even in the first
stage. Initially, their effect on the strain rate is negligible but, as their number and size
increase, this effect becomes definitive. The accelerating rate of creep can also be caused
by the collapse of the materials microstructure. Many metallic materials contain secondphase particles that act as obstacles in the movement of dislocations and improve the
strength to creep. Under a straining over a long period of time the possible growth of the
larger of these particles and the disappearance of the smaller, leads to increasing rates of
creep. But Dyson and McLean showed that, even in nickel super alloys (with a large
number of particles), the tertiary creep cannot be explained by the development of
particles. The main part of the accelerating rate of creep should be attributed to the
cavities of the grain boundary. Another possible cause (not definite) for the accelerating
rates of the third stage of creep is the corrosion on or below the surface, as, for example,

the internal oxidation (mainly at the boundary grain) and the following creation of a
crack. Under relatively medium stresses and high temperatures the metallic materials are
broken with a relatively low ductility. The decrease is due to the intercrystalline
development of cavities. Isolated cavities have been tracked at the second stage of creep
and, in some cases, at the first stage. At the later stage, the
Voids start to become unified at the sides of the grain boundaries, forming small cracks
(micro cracks). The unification of the micro cracks leads to the characteristic fibrousporous surface of the intercrystalline fracture. The general cause of the transition from the
intercrystalline failure (under low temperatures) to the intercrystalline failure is that the
atomic voids are rendered agile under high temperatures. The atomic voids that are
dispersed at the boundaries of grains can be concentrated to form cavity nuclei.
Moreover, the grain boundaries are active sources of voids so that they supply the
creation of cavities. In the alloys that usually contain second-phase particles at the limits
of grains, the cavities are, usually, created at the place of the particles. Numerous studies
have shown that in the low alloyed steels, the austenitic steels as well as the nickel super
alloys, the cavities are related to carbides, as well as to sulphides, silicates and oxide
existing in the main material. The development of creep damage can be expressed in
terms of two mechanisms. The one mechanism is the creation of cavities and provides a
measure for the rate at which the number of voids increases and the other mechanism has
to do with the development of cavities and provides a measure of the magnification of
cavities with time. The rate of creation of voids is represented as

and is measured

with the number of cavities created per unit of time on the unit area of the grains
boundary. Let
interval

be the rate of creation of cavities in a time

the number of new cavities created is

Therefore, in the time


In the later time, t, these

cavities will be magnified and the assumed rate of development of the cross-sectional
area of the cavities created in time
The total area, S, of the voids in time

will
is therefore

(1.1)
6

where S0 is the total cross-section. The above formulation is general but shows that, in
order to define a measure of damage, the effect of both functions M and W should
necessarily be known.
Rabotnov and Kachanov simplifying the analysis at this point, considered the
ratio

as a phenomenological (state) variable, and used the condition

as

the failure condition. They assumed that the following relation is valid

..(1.2)
where k is a constant. Moreover, similar relations have been proposed for the
development of damage .It is important to emphasize that, using the effect of damage on
the tertiary creep, it is possible to measure the quantity
method of measuring

In fact, in other studies, a

is suggested. Moreover, Equation is in perfect accordance with

the experimental data. Therefore, the variable


structure parameter.

concept is a fully defined macroscopic

2. LITERATURE REVIEW
The metallic materials creep behavior has been described and a complete model is
presented. The basic constitutive equation, as well as the structure parameters, has been
derived from a mathematical analysis that represents the dominant physical procedures
and mechanisms. The model is very general because it is referred to all stages of creep
and describes the creep behavior of all metallic materials, including those strengthened
by a dispersion of second-phase particles. A creep function has been derived from the
constitutive equation describing all three stages of creep under constant loading. The
function has the minimum possible number of fitting, parameters. The dependence of the
fitting parameters on the loading conditions has been described using very simple
mathematical relations. Applications and predictions have been carried out in a wide
range of metallic materials. Good agreement has been shown by a comparison made also
between the creep curves determined experimentally, and those obtained from creep
function and determined fitting parameters. [1]
Life time and failure modes are predicted for metallic bars sustaining tensile
creep. Experimental results show that a ductile or a brittle mode of fracture occurs
depending respectively on whether the nominal applied stress is large or small. The
analysis is based on a modeling of void nucleation and growth in which damage
evolution is controlled by two mechanisms of plastic flow in the matrix material. Fracture
is supposed to occur when the porosity attains a critical value which depends on the mode
of fracture considered. Experimental results are explained and described in terms of the
proposed model. [2]
From the equations of motion within the field theory of defects, creep curves are
derived and a relationship between the applied stress and the time to rupture under
different deformation conditions is obtained. The creep duration as a function of the
applied stress and the initial strain rate, as well as the ultimate strain, specifying the
material rupture, are found. [3]

Structural materials used in sodium cooled fast reactors (SFRs) shall have good
high temperature low cycle fatigue and creep properties, adequate weldability to fabricate
large size components and shall be compatible with the liquid sodium environment in
service. Austenitic stainless steels have been the natural choice for structural components
of SFRs worldwide. The creep design life of SFR component is very long and is of the
order of 40 years. This call for robust creep life rediction models to convert short and
medium term laboratory rupture data to design life. This seminar discusses the
application of creep dissipation energy concepts to predict creep rupture life of four
nitrogen alloyed grades of 316 LN SS. [4]

3. CONSTITUTIVE EQUATIONS
The basic equations of the problem are presented in this section. The material is
assumed to be porous and elastic-viscoplastic. Because elastic effects will be neglected in
this work, the total strain rate is equal to the plastic strain rate d * dp. Porosity is
accounted for by using the formulation of Xu and Needleman (1991). They worked along
the lines developed by Gurson (1977) to analyze ductile fracture by void nucleation and
growth. The generalization of Gursons approach to a rate-dependent matrix material was
made by Pan et al. (1983). There is no yield function in their approach. The viscoplastic
strain rate dp is given by the flow rule in terms of a viscoplastic potential .
.
.(2.1)
Where

is a positive scalar determined later. The viscoplastic potential has the

form
.(2.2)
Where the mean stress
macroscopic Cauchy stress,

and the effective stress


and Cauchy stress deviator

are defined in terms of the


as

..(2.3)
Equation (2) defines the effective matrix stress
potential

the porosity f is introduced In the

through the function f *(f) as proposed by Tvergaard and Needleman (1984)

to model the loss of stress carrying capacity due to void coalescence:

(2.4)
10

Where

differs from f, when the porosity f exceeds a certain critical value

fc. f * is equal to

when

stress-carrying

capacity

in that case the material suffers a complete loss of


as

can

be

The positive factor

seen

from

Equation

(2.2)

where

in the flow rule (1) is determined from

the equivalence between the macroscopic dissipation and the dissipation in the matrix
material (see Xu and Needleman, 1991):

.....
(2.5)
Where is the matrix effective plastic strain rate defined by
......
(2.6)
To complete the set of equations, the viscoplastic flow of the matrix remains to be given.
For metals at elevated temperatures, and for the level of stresses explored here, two basic
deformation mechanisms will be considered:
(I) thermally activated dislocation glide at large stresses, and
(II) Diffusion controlled dislocation creep at lower stresses.
To account for these two mechanisms, a phenomenological formulation of the Matrix-material
flow law is used:

In this relationship the effective plastic strain rate in the matrix

..(2.7)
is a function of the

effective matrix stress . Two strain rate sensitivity parameters m1 and m2 are introduced.
They are respectively related to mechanisms I and II discussed above. An important fact
is that the strain rate sensitivities associated with these mechanisms differ by an order of
magnitude. Mechanism I, associated with thermally activated dislocation glide, is
assumed to be sensitive to strain hardening. The strain dependence of the material flow is
11

described by the function

introduced in Equation (2.7), with

being the cumulated

plastic strain obtained by time integration of. The following form is taken:
.
(2.8)
where

is a reference stress,

is the strain hardening exponent and

is

the Young modulus.


The evolution of the porosity f is due to void nucleation and growth (see Xu and
Needleman, 1991):
.(2.9)
Where

..(2.10)

With

(2.11)

Table 2.1. Values of the material parameters typical of an alloy steel at 500 degree Celsius
12

..(2.12)
In the following, we consider that void nucleation is mainly strain controlled; therefore,
B will be kept equal to zero. By combining (2.92.11), and using the flow law (1) along
with the expression

...(2.13)
The evolution law for the volume fraction of voids can be written as

..(2.14)

The discussion in the next sections will be based on values of material parameters given
in Table I, which are typical of alloy steel at 500 degree Celsius. Values of parameters
related to void nucleation and growth, have been inspired by the work of Xu and
Needleman (1991). The strain rate sensitivities associated to mechanisms I and II are

13

taken as m1 = 0:02 and m2 = 0:25. For the parametric analysis, variations with respect to
the values given in Table I will be considered. The stress vs. strain-rate relationship (2.7)
for the matrix material is illustrated, in Figure 3, using values of Table I. The effective
matrix stress

is represented in terms of

on a logarithmic scale. Figure 3 shows the

activation of mechanisms I and II, respectively, for large and small values of the stress
Note the slopes m1 and m2 corresponding to these two mechanisms. Note the two Linear
branches corresponding to mechanisms I and II of viscoplastic flow with strain rate Sensitivity
m1 and m2, respectively.

Figure 3.1. Stress-strain rate response of the matrix material (log-log diagram).

14

4. CREEP ANALYSIS WHEN DAMAGE IS NEGLECTED

The effect of porosity is excluded in this section by keeping f = 0 through all the
calculations. We just have to solve Equations (2.7) and (3.4). Evolution of elastic
deformations during the process is neglected as stated in the beginning of the paper and
values of the material parameters are those of Table I. This analysis is first made for an
ideal specimen with uniform cross-section. The solution of this problem will be
designated as the fundamental uniform solution. In the next section, the development of
heterogeneous deformations due to an initial cross-section defect is analyzed. The
nominal stress is defined by
(3.1)
Where S0 is the initial cross-section

Figure 4.1. Numerical simulations of the evolution of the axial strain


time, in a creep test on a cylindrical bar.
15

with respect to

Here porosity is not taken into account: f =0. Values of the material parameters
are defined in Table I. These values are typical of steels at elevated temperatures. For the
applied nominal stress

, it appears that the early stages of the process are

dominated by mechanism II of diffusion controlled dislocation creep. At the later stages,


mechanism I of thermally activated dislocation glide is activated, leading to an explosive
growth of the deformation. The dashed line corresponds to a simulation where
mechanism I is not considered.
The evolution with respect to time of the axial deformation
(continuous line) for

. By

explosive growth of the deformation

is shown in Figure 4

we designate the time at which an

is observed. Until times close to

= 1.68 X 10^6

s, the plastic flow appears to be governed by mechanism II (m2 = 0.25). This point is
made clear in Figure 4 by considering the evolution of , when mechanism I is neglected
(

; dashed line). While mechanism I appears to be negligible for

most times t <

, it becomes important near

Thus the critical time

and leads finally to the explosion of

seems to be governed in general by the interplay of both

mechanisms I and II. These results can be related to the contribution of the two terms of
the constitutive flow law. Since f = 0, it follows from that
have

. Then from (6) we

where is defined. Thus the flow law has the form

(3.2)
to Because f = 0, the material is incompressible, and therefore the cross-section area is

given by
S = S exp. (- ). The true stress

is then related to the nominal stress

by

It , leads to

.
(3.3)

16

To evaluate the respective contributions of the two mechanisms controlling the material
flow, the ratio of the two last terms of Equation (24) is considered:

..
(3.4)
Due to the condition n1 > n2 (m1 < m2), the evolution of
. For large values of

in terms of

vanishes. If

has a maximum
, mechanism II is

dominant in some range of deformation. However, because

mechanism I shall eventually overcome mechanism II as observed in the calculations


reported in Figure 4. An interesting point concerns the role of the nominal stress. When
is increased,
Assume now that

is made smaller, meaning that the mechanism I is more prominent.


is large enough so that

. Then the contribution of

mechanism II is always negligible. On the other hand, when

is small enough,

mechanism II is dominant during the process. A relationship between time and


deformation can easily be obtained from Equation (3.5). After separation of variables and
integration, it yields

...
(3.5)
The time t as defined in Figure 4, for which S 0, is obtained by making

in

(3.5).
Let us now consider the case of large values of

. Then from the preceding discussion,

the contribution of mechanism II can be neglected, and we obtain from (3.6).

.(3.6)

With,

17

(3.7)

Figure 4.2. Nominal stress

vs. critical time to.

This time corresponds to the moment where an explosive growth of the


deformation is observed in the fundamental homogeneous solution. Note that the two
slopes (m1 = 0:02) and (m2 = 0:25) are associated to mechanisms I and II of the
plastic flow. Therefore, a linear relationship, with slope m1, exists between ln

and ln

t.

.(3.8)
For small values of , a similar relationship is obtained with m1 replaced by m2.
These results have to be put in perspective with experimental data where the same type of

18

relationship is found between logarithms of

and

is the time to rupture

observed in the experiments).


In Figure 5 the nominal stress

is reported in terms of the time t in logarithmic

scale. The calculated curve shows two linear branches with slopes m1 and m2.
Mechanism I (respectively II) is dominant for large (small) values of

. A transient zone,

where both mechanisms have a contribution, makes the link between the linear branches.
A typical value of the nominal stress corresponding to this transient zone is
. That value was used for the calculations represented , where it appeared
that mechanisms II and I were successively activated during the process.
The experimental results shown in Figure 1 for metals and in Figure 2 for
polyethylene are qualitatively in agreement with those of Figure 5. As the material
parameters used in the calculations are related to steel, a quantitative agreement can be
found with the results of Richard et al. shown in Odqvists paper.
The main conclusion of this section is that the two slopes observed in the stresstime to rupture diagram can be related to different mechanisms of viscoplastic flow with
a low and a large strain rate sensitivity (here m1 = 0:02, m2 = 0:25). Damage evolution,
neglected in this section, is accounted for in the following and will be shown to have an
important effect on the failure modes.

19

5. DAMAGE AND FAILURE MODE ANALYSIS

5.1. Damage Influence


The influence of damage on fracture is now examined. Porosity evolution is taken
into account by Equation (3.1). Cavities may interact when their size becomes large or
when their number is multiplied. As a result, cracks are initiated and propagate across the
section, leading to failure. A simple criterion is used here to account for the interaction
between cavities. We shall consider that fracture is initiated when a critical value fR of the
porosity is reached.
Different values of fR will be introduced, depending on the fracture mode that is activated.
At large stresses, fracture is the result of void nucleation and growth within grains
(ductile fracture or transgranular creep fracture). At lower stresses, intergranular creep
fracture can be observed with grain boundary cavitations. This is a brittle damage
mechanism. Two different values of the critical porosity fD and fB are attributed to the
ductile and brittle failure respectively. In ICF, voids are located at the grain boundaries.
Therefore at a given value of the overall porosity, interaction between voids is enhanced,
compared to TCF where voids are spread uniformly in the material. Somehow in the case
of ICF, the notion of an effective local porosity Q f at the level of grain boundaries can be
introduced. Due to the Concentration of voids within thin layers, the effective porosity f
is much larger than the overall porosity f. For that reason, fracture is initiated in the case
of ICF, at values of the global porosity significantly smaller than for TCF. Therefore it is
assumed that fD > fB.
In a first step, the cross-section is assumed uniform. Two different characteristic
times are introduced. As before, t0 is the time needed for a uniform bar to have its crosssection reduced to zero; tR denotes the time for a uniform bar to develop the critical value
of the porosity f = fD, resp. f = fB, associated to a given fracture mechanism (ductile
fracture or TCF, resp. ICF). In the following, the time to rupture tR is calculated in terms
of the nominal stress

. The initial porosity is f0 = 0:04.

20

In Figure 5.1, we report the evolution with respect to time of the porosity and of
the cross-section area for an ideal uniform bar. A small nominal stress is applied,

100 MPa. An explosive growth of the porosity and a drastic reduction of the crosssection area are reached at the end of the process.

Figure 5.1. Evolution of: (a) porosity (b) cross-section area; with respect to time for a
nominal stress = 100 MPa.
21

Note that the critical porosity fR= 0.06, at which rupture is initiated, is reached at the
time tR. In the present case, this time is smaller than the time t0 for which the crosssection area vanishes

Figure 5.2. Effect of the nominal stress

Note that tR = t0 when

on the time to rupture tR and on t0.

is large, and that tR < t0 when

is small. The transition

from the deformation mechanism II to mechanism I is indicated by arrows.


For discussion, a given value fR = 0.06 is considered. It is of interest to note that
this critical value of the porosity (indicative of the onset of rupture) is attained before the
vanishing of the cross section area (tR < t0). As discussed in the previous section, the
process is controlled by mechanism II, providing that the nominal stress is small enough.
In fact no change in the result is found when calculations are made by switching off
22

mechanism I ( = 0). The protuberance that appears in the curve of Figure 4.3a at the
time t = 8 x 10^7 s, results from the nucleation of voids activated at the strain
introduced in the relationship.
Different values of the nominal stress

are considered in Figure 7, where the

evolution of porosity with respect to time is shown. The values chosen for the critical
porosity are fD = 0.1 for ductile fracture at large stresses (here
0.06 for brittle fracture at low stresses (
4, that for increasing values of

> 300 MPa) and fB =

< 300 MPa). It is found, as discussed in Section

, the deformation mechanism I is taking a more

prominent role. In the early stage, the process is controlled by the deformation
mechanism II. At a certain moment marked by an arrow in Figure 7 mechanism I is
activated. The absence of an arrow for the small stress

= 100 MPa, means that the

process is entirely controlled by mechanism II. The porosity at which the mechanism I is
activated, is larger when

is decreased (see arrows in Figure 7). As a consequence, the

time difference t0 tR tends to zero when the nominal stress is increased. Indeed for
400 MPa, we have tR = t0 since ffR and S 0 simultaneously.

23

Figure 5.3. Influence of the damage evolution on the time to rupture

Note that by choosing smaller values of fB, the deformation at fracture in the
brittle mode would be reduced. In that way, brittle fracture at strains less than 5%
observed in certain engineering alloys would be described. In Figure 8, the nominal stress
is given in terms of the time to rupture on a logarithmic scale. The continuous line
refers to the situation where damage evolution is disregarded. Then, the time to rupture is
considered as being the time at which the section is reduced to zero (trup = t0). The case
where damage evolution is accounted for corresponds to the dashed lines. Two values of
the critical porosity are considered fR = 0.1 and fR = 0.06. The time to rupture is now the
moment where the porosity reaches the value fR (trup =tR). Two different slopes are
associated with the mechanisms of deformation I and II. Actually, scaling laws similar to
(2.8) are still valid for t0 and tR when the evolution of porosity is controlled by the Gurson

24

model. This point is demonstrated in Appendix A. We can remark that the lower branch
of the curve in Figure 8 is affected by the value of fR while the upper branch remains the
same for the two values of fR considered. We have to consider now that the value of fR
depends on the fracture mode. We have chosen fR = fD = 0.1 (ductile fracture) and fR = fB
= 0.06 (brittle fracture).

5.2 Failure Mode Analysis


So far, the discussion has dealt with the idealistic case where the flow remains
homogeneous along the bar. This case is referred to as the fundamental homogeneous
solution. The actual response observed in experiments is different. Ductile fracture is
accompanied by the development of an inhomogeneous deformation or necking. As
presented in the Introduction, mechanical tests indicate that for large values of the
nominal stress

, failure is preceded and accompanied by neck development (ductile

failure). On the other hand, when

is made smaller, the importance of necking is

reduced (sometimes a neck is hardly observed), while damage plays a major role in the
rupture process. Damage growth triggers rupture by crack initiation and crack
propagation. These experimental observations can be interpreted using features related to
the homogeneous fundamental solution.
Let us consider a bar with a cross-section defect. Denote by SA and SB the largest
and smallest cross-section area. In Figure 9, an example is shown of a model with two
zones A and B, the cross-section area being uniform in each zone. For a large value of

the material flow is governed by the small strain rate sensitivity m1, as discussed before.
Alternately, for small values of

the dominant mechanism is controlled by the large-

strain rate sensitivity m2. It is known that plastic instability and neck development is
favored by a small-strain rate sensitivity (here m1). This is schematically illustrated in
Figure 9 where the evolution of the axial deformation

(zone B) in terms of

(zone

A) is reported for two values of the strain rate sensitivity (m1 < m2). At the beginning of
the process, the material flow is stable and the deformations
identical. Localization of the deformation occurs later in zone B;

25

and

are almost

increases to large

values, while

saturates at some critical value

1977) that for a given geometrical defect


rate sensitivity.

. It is known (Hutchinson and Neale,

depends strongly on the value of the strain

is smaller for a process controlled by the small strain rate sensitivity

(m1) than for a process governed by the large strain rate sensitivity (m2), . The schematic
evolution of the deformation

in the thinner zone B, in terms of the deformation

zone A, is shown in Figure 9, for two values of the nominal stress


and

(dashed line), such that

>>

..(4.1)
Strain localization for the small stress

in

(continuous line)

. Therefore, we have

occurs later than for the large stress

Figure 5.4. Schematic representation of strain localization in a cylindrical bar with a two
zone model.
Two levels of the applied stress are considered with
>>
...

26

Depending on whether the stress

or

is applied, mechanisms I or II of the plastic

flow are respectively activated. Note that the localization strain

in zone A depends on

the strain rate sensitivities of the deformation mechanisms.

Let us denote by

the strain at which the critical value fR of the porosity is

reached. It can be checked by numerical calculation that

is weakly dependent on the

strain rate sensitivity. This can be related to the fact that the evolution of porosity is
driven by plastic deformation.
The value of fR depends on the fracture mode considered: ductile fracture fR = fD,
; brittle fracture fR = fB ,

. The values of

and

shown in Figure 9 correspond to the onset of rupture. It appears that for the small stress
, rupture occurs before strain localization. The damage failure mode is favored here
because of the large strain rate sensitivity m2 which refrains strain localization and neck
formation. On the other hand, for

, strain localization is initiated much earlier, due to

the small strain rate sensitivity m1. Not enough time is left for the porosity to reach the
critical value fR = fD before neck development. Rather, rupture is triggered during
necking. This description, although only qualitative, is consistent with experiments.
Quantitative investigations could be performed with a three-dimensional finite element
numerical model. However, the main results and trends are certainly described by the
simple model presented here.

27

Figure 5.5. Influence of the damage evolution on the time to rupture

28

6. CREEP DURATION ANALYSIS IN TERMS OF THE FIELD


THEORY
The operation of equipment under severe conditions (high stress and
temperature) has resulted in the discovery of the creep effect and the development of
creep theory. Engineers have centered on creep analysis, i.e., on the evaluation of the
time period within which the strain reaches the ultimate value. This problem remains of
practical importance nowadays. At its incipient stage, creep theory was developed as an
engineering science. Later, it evolved into a branch of continuum mechanics.
Simultaneously, physical mechanisms responsible for the creeping effect were studied.
Because of their complexity, comprehensive physical description of creep is lacking.
Some progress in creep physics has been achieved by invoking the concepts of the
dislocation theory. A number of dislocation models describing different creep stages and
conditions have been constructed. It is argued that, at moderate temperatures, elementary
creep events in solids are attributed primarily to the motion of dislocations. Therefore,
creep mechanisms will be considered in terms of the field theory, which involves the
dynamics of translational defect. Note that the field theory of defects deals with defect
ensembles and, according to, describes a system on the mesoscopic scale. In contrast, the
classical dislocation theory has to do with individual defects and their interactions and
thus, implies microscopic methods of description.
The dynamic equations in the field theory of defects have the form,

....
(5.1)
From these equations, the equation relating the tensor of the dislocation flux density I to
the applied stress

was obtained:
29

.(5.2)
Here,

is the dislocation density tensor, V is the elastic displacement rate,

material density,

is the viscosity, B and S are the theoretical constants, and

is the
is the

Kronecker delta. The symbols (X) and () stand for the vector and scalar products,
respectively, and(X.) designates the vector product with respect for the first subscripts of
the dyad and the scalar product with respect for the second ones. Equation (2.2), which is
helpful in studying the creep process, is written for the uniform distribution of defects,
i.e., for space-independent field strengths

and I. It is believed that these assumptions

valid near the yield point, where defects are distributed randomly and do not form spatial
structures. A great body of experimental data on the creep effect has been obtained from
tensile tests of rods. Therefore, in our previous study, we considered uniaxial
deformation, for which Eq. (2.2), written in the dimensionless variables,
,

and

becomes,

..(5.3)
Where vis the plastic strain rate

The special attention has been given to the analysis of the functions v (t), which specifies
the
creep curves

under constant stress. In the following, we will carefully

investigate the relationship between the applied stress and the time to rupture of a system.
It has been found that there are two creep modes depending on the applied stress S: stable
at S < 1/2 and unstable at S > 1/2. The corresponding expressions for creep rates are
.(5.4)
at S 1/2,

(5.5)

30

at S 1/2,

Then, the creep curves are given by,


..
(5.6)
Where,
is the initial creep rate, and p, q=

are the steady-state rates given by

dv/d= 0.
The critical applied stress S= 1/2 corresponds to the bifurcation point where the
creep behavior changes. With this parameter introduced, one can define the stable creep
limit
* = /2B, which depends on the material parameters. Expressions (2.4)(2.7) yield the
condition
...(5.7)
From which the creep life of a real system at S > Sis obtained. Under condition (2.8),
creep rate (2.5) becomes infinite and the time to rupture is given by
.(5.8)
Curves plotted in Fig. 1 relate the time to rupture of the system to the applied stress at S >
S* and v0 = (1) 0.5 and (2) 0.9.
According to, when S < S*, solutions of Eq. (5.7) are strongly dependent on the
initial value of v0. The range of v0 can be divided into the intervals 0 < v0 < q, q < v0 < p,
and v0 > p. At v0 > p, the creep rate becomes infinite when the denominator of (4)
vanishes; hence, the time to rupture is
..
..(5.9)
Expression (2.10) is illustrated in Fig. 1 by curves 3 and 4 for v0 = 1.6 and 1.9,
respectively. If the initial creep rate lies in the range q < v0 < p, v () q at ; i.e.,
the stationary creep mode is observed. In this situation, creep duration analysis turns to
31

the conventional problem (see the first paragraph). If the critical strain * at which the
material fails is known, Eq. (5.9) with yields the relation of interest between the
stress and the time to rupture:

This dependence is demonstrated in Fig. 2 for * = (1) 0.65 and (2) 0.35. In both
cases, the initial strain 0 is taken to be equal to 0.01%. This value is in qualitative
agreement with Figs. 1 and 3. The latter depicts experimental data for a number of
materials tested at various temperatures. The test duration was up to 100 days. Curves 1
and 4 were obtained for carbon steel (0.5% C and 0.24% C) at temperatures of 300 and
432C, respectively; curve 2, for nickel steel at 400C; curves 3 and 5, for high-alloy
nickelchromium steel at 600 and 700C, respectively; 7 and 11, for high-speed steel at
593 and 732C, respectively; 6, for cast heat resistant steel at 800C; 8, for lead at room

32

temperature; 9, for nickelcopper alloy at 600C; and (2.10), for magnesium alloy at
150C. Following the conclusions made .we leave without consideration the initial range
0 < v0 < q, where attempts to describe the early portion of the creep curve with decreasing
rate have failed.

33

7. CASE STUDY
Detection of life of structural materials used in sodium cooled fast
reactors

(SFRs)

Creep dissipation energy concepts to predict creep rupture


life:Notations
E = Youngs Modulus in N /mm2
= Uniaxial equivalent creep strain rate (/hr)
= Uniaxial equivalent deviatoric stress or Mises equivalent stress (N/mm2)
= Mises equivalent stress at timet (N/mm2)
t= Total time (hr)
A, m, n Material parameters
Structural materials used in sodium cooled fast reactors (SFRs) shall have good
high temperature low cycle fatigue and creep properties, adequate weldability to fabricate
large size components and shall be compatible with the liquid sodium environment in
service. Austenitic stainless steels have been the natural choice for structural components
of SFRs worldwide. The creep design life of SFR component is very long and is of the
order of 40 years. These calls for robust creep life rediction models to convert short and
medium term laboratory rupture data to design life. This paper discusses the application
of creep dissipation energy concepts to predict creep rupture life of four nitrogen alloyed
grades of 316 LN SS. Austenitic stainless steels have been chosen world wide as the
structural material for high temperature components of SFRs. The choice has been based
on their good high temperature mechanical properties, compatibility with the coolant
sodium and adequate weldability. The material chosen for present study was 316(L) N
stainless steel with different nitrogen contents (0.07 Wt. % N, 0.11 Wt. % N, 0.14 Wt. %
N and 0.22 Wt. % N). The creep design life of SFR components is very long and is of the
34

order of 40 years. Structural analysis under creep conditions requires reliable constitutive
models which reflect time dependent creep deformation. In the present investigation, an
attempt has been made to predict creep curves and creep rupture life on the basis of creep
dissipation energy. The Bailey-Norton form of creep law is derived from the experimental
data from uniaxial creep tests for 316 (L) N stainless steel with different nitrogen
contents. A strain hardening form of creep law is used in ABAQUS FEM code to evaluate
the equivalent creep constants.
The datas for 316(L) N SS with 0.07% Wt. % N, 0.11% Wt. % N, 0.14 Wt. % N
and 0.22 Wt. % N at 175 MPa and 200 MPa at 923 K from the uniaxial creep test are
fitted to the Bailey-Norton form of creep law and equivalent strain hardening ABAQUS
creep constants were evaluated accordingly.
The Bailey Norton law,
.
(6.1)
Taking log on both the side of the equation
..(6.2)
Equation (2.2) can be compared to the standard plane equation of
...................(6.3)
The best fit plane for the above mentioned plane is found using the least square method.
The equation of plane can be written as follows.
Z (x, y; a, b, c) = c + ax + by...(6.4)
Minimizing the sum of the squares of the vertical distances between the data and the
plane and which is called as

For extreme values the gradients must be zero,

35


(6.5)

Solving the above set of linear equation we get the values of the constants a, b and c
which are equivalent to the values of p, q and log (K) respectively of Bailey-Norton
equation. The value of K is obtained by taking antilog (exponential).
7.1 Derivation of equivalent creep constants of strain hardening:-

Strain hardening creep is modeled using ABAQUS software. The available time
hardening creep subroutine in ABAQUS can be used for simulating strain hardening
creep by deriving its equivalent constants. The ABAQUS time hardening is calculated
using the formula,
..(6.6)
Integrating Eq. (2.13) w.r.tot, creep strain can be written as

....(6.7)
From the above Eq. (2.14).,.

.(6.8)
36

Substitutingt in Eq. (2.13) we get the ABAQUS equivalent strain hardening form of
power law and is given by [4-5]
.. (6.9)
The Bailey Norton law given in Eq. (2.1)
(6.10)
Differentiating Eq. (2.1) with respect tot we get
...... (6.11)
Solving fort from Eq. (2.1)

. .. (6.12)
Substituting Eq. (3.4) in Eq. (3.3) we get

.... (6.13)
Comparing Eq. (3.5) with the equivalent strain hardening form used in ABAQUS- Eq.
(16) we get
(6.14)
... (6.15)
... (6.16)
The creep constants for 316 (L) N SS for the Eq. (2.13) or for Eq. (3.2) are given in Table
1 for different loads which were calculated from experimental data fitting in the Eq. 1012.

37

Specimen. No
1(0.07 % N)
2(0.11 % N)
3(0.14 % N)
4(0.22% N)

A
N
Applied load 175 MPa at 923 K
5.8417 X 10^-14
4.5607
5.972 X 10^-17
5.7606
5.898 X 10^-17
5.7501
5.547 X 10^-21
7.141
Table 6.1: Creep constants for 316(L) N SS

m
-0.3152
-0.3351
-0.3251
-0.3558

Creep dissipation energy can be found using the formula:

. (6.17)
Where o is the Applied stress, cr is found from the graph for experimental value of time
versus strain, V is total volume = A.L andt is the experimental creep life.
Creep dissipation energy for 316(L) SS with different nitrogen content (0.07% N, 0.11%
N, 0.14% N and 0.22% N) have been tabulated in Table 4. These values are used to find
out simulated creep life from the creep dissipation curve generated from analysis result.
Creep dissipation energy for different specimens was calculated from the experimental
data using Eq. (4.1). Creep dissipation energy vs time has been generated for different
cases and creep life in terms no. of hours were given in Table 4.

38

Fig. 1: Creep dissipation energy for the specimen 1 (0.07% N) 316(L) N SS of load 175MPa

Specimen

Stress()

Strain

No.

MPa

(Per hr)

rate Volume(mm^3)

Time(hrs)

Creep
dissipation
Energy(N-

1(0.07 % N)
2(0.11 % N)
3(0.14 % N)
4(0.22% N)

175
175
175
175

0.00045
0.00031
0.00022
0.00016

3544.10
3544.10
3544.10
3544.10

738
1484
1422
4412

Table 4: Experimental results for 316(L) N SS

RESULT:-

39

mm)
205974
285325
194029
437824

Hence the simulation results can help to find out the creep life for different loads
using the less number of specimen data.
From the results, it shows that increase in creep life was found for 316(L) N SS
by increasing the level of nitrogen content (quantitatively estimated and tabulated for
different specimens of 0.07% N, 0.11% N, 0.14% N and 0.22% N). Comparing specimen
of 0.22% N with different loading condition of 175 MPa load is increased by 1.14 times
and respective creep life is decreased by 0.71 times.

40

REFERENCES
[1]Review Mathematical description of the mechanical behavior of metallic materials
under creep conditions, by N. D. Batsoulas. Journal of material sciences 32
(1997) 2511 - 2527.
[2]Creep Failure Analysis of Bars Sustaining Constant Tensile Loads by Elf Atochem,
Cerdato. Mechanics of Time-Dependent Materials 4: 5779, 2000
[3]Creep Duration Analysis in Terms of the Field Theory of Defects by N. V. Chertova
and Yu. V. Grinyaev. Technical Physics, Vol. 46, No. 7, 2001, pp. 844846.
[4]Creep rupture life prediction based on creep dissipation energy analysis by Vela
Murali1,M.D. Mathew2, K. Bhanu Sankara Rao2, V. Ganesan2, S. Ravi2, S.
Baskar1 and N. Angaraj1. Vol 63, issues 2-3 , April-June 2010 , pp 635-639.

41

Das könnte Ihnen auch gefallen